paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
0909.0067
3
0909
2012-11-06T01:47:14
Bilinear biorthogonal expansions and the Dunkl kernel on the real line
[ "math.FA", "cs.IT", "cs.IT" ]
We study an extension of the classical Paley-Wiener space structure, which is based on bilinear expansions of integral kernels into biorthogonal sequences of functions. The structure includes both sampling expansions and Fourier-Neumann type series as special cases, and it also provides a bilinear expansion for the Dunkl kernel (in the rank 1 case) which is a Dunkl analogue of Gegenbauer's expansion of the plane wave and the corresponding sampling expansions. In fact, we show how to derive sampling and Fourier-Neumann type expansions from the results related to the bilinear expansion for the Dunkl kernel.
math.FA
math
the expansion in terms of the prolate spheroidal wavefunctions ϕn, (2) inλnϕn(x)ϕn(t), sin(x − πn) x − πn ∞Xn=−∞ e−ixt = √2π ∞Xn=0 BILINEAR BIORTHOGONAL EXPANSIONS AND THE DUNKL KERNEL ON THE REAL LINE LU´IS DANIEL ABREU, ´OSCAR CIAURRI, AND JUAN LUIS VARONA Abstract. We study an extension of the classical Paley-Wiener space structure, which is based on bilinear expansions of integral kernels into biorthogonal sequences of functions. The structure includes both sampling expansions and Fourier-Neumann type series as special cases, and it also provides a bilinear expansion for the Dunkl kernel (in the rank 1 case) which is a Dunkl analogue of Gegenbauer's expansion of the plane wave and the corresponding sampling expansions. In fact, we show how to derive sampling and Fourier-Neumann type expansions from the results related to the bilinear expansion for the Dunkl kernel. The function K(x, t) = eixt has several well-known bilinear expansion formulas, for example: the Fourier series expansion, 1. Introduction (1) eixt = eiπnt, t ∈ [−1, 1]; where λn are the square roots of the eigenvalues arising from the time-band limiting integral equation (see the recent paper [28]); and Gegenbauer's expansion of the plane wave in Gegenbauer polynomials and Bessel functions (see [16, § 4.8, formula (4.8.3), p. 116]) (3) in(β + n)Jβ+n(x)C β n (t), t ∈ [−1, 1], eixt = Γ(β)(cid:16) x 2(cid:17)−β ∞Xn=0 (in the particular case β = 0, this formula is the so-called Jacobi-Anger identity). Here and in what follows in this paper we are using C β n to denote the Gegenbauer polynomials of order β and Jν to denote the Bessel functions of order ν. Each of the above expansions is associated with important developments in mathemat- ical analysis. The first one is equivalent to the Whittaker-Shannon-Kotel'nikov sampling theorem (see [27, Ch. 2]), the second one is the prototype of a Mercer kernel [28], and the third one has been the main tool in the diagonalization of certain integral operators [17]. Our main interest is in expansions of the type (3). To see why biorthogonality is required, recall the definition of the Paley-Wiener space P W , P W =(cid:26)f ∈ L2(R) : f (z) = (2π)−1/2Z 1 −1 u(t)eizt dt, u ∈ L2(−1, 1)(cid:27) . THIS PAPER HAS BEEN PUBLISHED IN Expo. Math. 30 (2012), 32 -- 48. 2000 Mathematics Subject Classification. Primary 94A20; Secondary 42A38, 42C10, 33D45. Key words and phrases. Bilinear expansion, biorthogonal expansion, plane wave expansion, sampling theorem, Fourier-Neumann expansion, Dunkl transform, special functions. Research of the first author supported by CMUC/FCT and FCT post-doctoral grant SFRH/BPD/26078/2005, POCI 2010 and FSE. Research of the second and third authors supported by grant MTM2009-12740-C03-03 of the DGI. 1 2 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA We immediately obtain orthogonal expansions for f ∈ P W by simply integrating equa- tions (1) and (2), by using the orthogonality of the exponentials and the prolate spheroidal functions. However, if we try to do the same thing in (3), we must restrict ourselves to the case β = 1/2, when the weight function of the Gegenbauer's polynomials (actually Le- gendre) is 1. Therefore, even in this simple case it is not clear how to expand Paley-Wiener functions into Gegenbauer polynomials or Bessel functions with general parameter β, and the biorthogonal formulation serve to introduce the parameter β of (3) in a natural way. To organize the presentation of our ideas, we first construct a structure involving biorthogonal expansions, from which the results are obtained, after explicit evaluation of some integrals. In particular (but not exclusively), we use this structure to analyze the solution of the above mentioned expansion problem and its extension to the Dunkl kernel Eα(ix) = 2αΓ(α + 1)(cid:18) Jα(x) xα + Jα+1(x) xα+1 xi(cid:19) (as we will see in subsection 4.1, the Dunkl kernel is used to define the Dunkl transform on the real line similarly to how the kernel eixt is used to define the Fourier transform), and so we expand functions in P W and its generalization studied in [6] in terms of Fourier- Neumann series. From the following extension of (3) to the Dunkl kernel Eα(ixt) = Γ(α + β + 1)(cid:16) x 2(cid:17)−α−β−1 ∞Xn=0 in(α + β + n + 1)Jα+β+n+1(x)C (β+1/2,α+1/2) n (t), where C (β+1/2,α+1/2) uniformly convergent Fourier-Neumann type expansions n are the so-called generalized Gegenbauer polynomials, we obtain f (x) = ∞Xn=0 an(f )(α + β + n + 1)x−α−β−1Jα+β+n+1(x), valid for f ∈ P Wα (the natural generalization of the Paley-Wiener space as in [6]), and where Moreover, in some cases, the coefficients an(f ) are identified as Fourier coefficients. The paper is organized as follows. In the second section we describe our problem in abstract terms. First we build the general setup for bilinear orthogonal expansions and, later, we modify it to consider biorthogonal sequences in the expansions (see Theorem 1). In the third section we describe the results which are obtained in the case of the Fourier kernel. The fourth section studies the expansion associated with the Dunkl kernel (see Theorem 2), that we think is a new and interesting result; moreover, we also show its consequences for the Hankel transform. In the last section we collect the evaluation of some integrals involving special functions which were essential for the paper but could not be found in the literature. 2. Structure 2.1. Orthogonal expansions. We begin with K(x, t), a function of two variables defined on Ω × Ω ⊂ R × R satisfying K(x, t) = K(t, x) almost everywhere for (x, t) ∈ Ω×Ω, and an interval I ⊂ Ω. Using this function as a kernel, define on L2(Ω, dµ), with dµ a non-negative real measure, an integral transformation by (4) (Kf )(t) =ZΩ f (x)K(x, t) dµ(x). an(f ) = 2α+β+1Γ(α + β + 1)ZR f (t) Jα+β+n+1(t) tα+β+1 dµα+β(t). BILINEAR BIORTHOGONAL EXPANSIONS 3 Moreover, we suppose that K is invertible and that the inverse is (5) g(t)K(x, t) dµ(t). Then, from Fubini's theorem we get the multiplication formula (eKg)(x) =ZΩ (Kf )g dµ =ZΩ ZΩ (eKf )g dµ =ZΩ ZΩ (Kg)f dµ (eKg)f dµ. (6) and also Moreover, if in the multiplication formula we take g = K(f ) and use K(h) = eK(h), we get kKfkL2(Ω,dµ) = kfkL2(Ω,dµ). As usual, it is enough to suppose that the operators K and eK, defined by (4) and (5), are defined on a suitable dense subset of L2(Ω, dµ), and later extended to the whole L2(Ω, dµ) in the standard way. Moreover, let us also assume that, as a function of t, K(x, t)χI (t) belongs to L2(Ω, dµ) (or, in other words, K(x,·) ∈ L2(I, dµ)). Here, χI stands for the characteristic function of I. Now, let N be a subset of the integers, {φn}n∈N be an orthonormal basis of the space L2(I, dµ) and {Sn}n∈N be a sequence of functions in L2(Ω, dµ) such that, for every n ∈ N , (7) (KSn)(t) = χI (t)φn(t) (notice the small abuse of notation in the use of χIφn; here, φn is a function that is defined only on I, and by χI φn we mean that we extend this function to Ω by having it be the null function on Ω \ I; we will use this kind of notation often in this paper). Consider the subspace P of L2(Ω, dµ) constituted by those functions f such that Kf vanishes outside of I. This can also be written as P =nf ∈ L2(Ω) : f (x) =ZI u(t)K(x, t) dµ(t), u ∈ L2(I, dµ)o. On the one hand, by using that K is an isometry, it follows that Sn is a complete or- thonormal sequence in P. This implies that every function f in P has an expansion of the form On the other hand, from (7) we have Consequently, the Fourier coefficients of K(x, t), as a function of t, in the basis {φn}n∈N on L2(I, dµ) are Sn(x). As a result, K(x, t) has the following bilinear expansion formula: (8) (9) cnSn(x). f (x) = Xn∈N eK(χI φn) = eK(KSn) = Sn. K(x, t) = Xn∈N Sn(x)φn(t), that must be understood as in L2(I, dµ(t)) for every x ∈ Ω. Remark 1. The reader familiar with sampling theory, in particular with the generaliza- tion due to Kramer (see [19] or, also, [27, Theorem 3.5]), has probably noticed strong resemblances. Indeed, Kramer's lemma corresponds to a particular case of the above situ- ation when there exists a sequence of points xk such that Sn(xk) = δn,k. This implies that {K(xn,·)} = {φn} is an orthogonal basis of L2(I, dµ) and that P has an orthogonal basis given by {eK(χI K(xn,·))} = {eK(χI φn(·))} = {Sn}. The orthogonal expansion in the basis {Sn}n∈N is the sampling theorem. In [14] there is given a detailed exposition of a similar structure, which, although restricted to sampling theory, is in its essence equivalent to the one that we have described. 4 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA The objects that we are interested in here are mainly those expansions that fit in the above setup, but that are not sampling expansions. As we will see, there exist quite a few of these. We will see in this work a wealth of situations where explicit computations of certain integrals yield new expansion formulas of the type (9), but in general they are special cases of the more general setting that we will provide in the next section. Remark 2. Perhaps the most remarkable feature that this setup inherits from sampling theory is the fact that, in many situations, uniform convergence can be guaranteed, once we know that the expansion converges in norm. This happens because P is a Hilbert space with a reproducing kernel given by k(x, y) = Xn∈N Sn(x)Sn(y) =ZI K(x, t)K(y, t) dµ(t). This fact can be proved using Saitoh's theory of linear transformation in Hilbert space [22, 23] in a way similar to what has been done in [1] and also by the same arguments in [14]. The uniform convergence of the expansions (8) is now a consequence of the well- known fact that if the sequence fn converges to f in the norm of a Hilbert space with reproducing kernel k(·,·), then the convergence is pointwise to f and uniform in every set where kK(x,·)kL2(I,dµ) is bounded. 2.2. Biorthogonal expansions. We now consider the same setup as in subsection 2.1 (in particular, the notation for the operators Kf and its inverse eKg in terms of a kernel K(x, t) that satisfies the multiplication formula (6)), but instead of the orthonormal basis {φn}n∈N of the space L2(I, dµ), we assume that we have a pair of complete biorthonormal sequences of functions in L2(I, dµ), {Pn}n∈N and {Qn}n∈N , that is, ZI Pn(x)Qm(x) dµ(x) = δn,m and every g ∈ L2(I, dµ) can be written, in a unique way, as cn(g) =ZI g(t) = Xn∈N cn(g)Pn(t), g(t)Qn(t) dµ(t). Let us also define, in L2(Ω, dµ), the sequences of functions {Sn}n∈N and {Tn}n∈N given by (10) and Sn(x) = eK(χI Qn)(x), Tn(x) = K(χI Pn)(x), x ∈ Ω, x ∈ Ω (note that if Pn = Qn then Sn = Tn). Our purpose is to prove the following theorem, which says that it is still possible to find a bilinear expansion in this context. Theorem 1. For each x ∈ Ω, the following expansion1 holds, with respect to t, in L2(I, dµ): (11) Pn(t)Sn(x). K(x, t) = Xn∈N Moreover, {Sn}n∈N and {Tn}n∈N are a pair of complete biorthogonal sequences in P such that every f ∈ P can be written as (12) cn(f )Sn(x), x ∈ Ω, f (x) = Xn∈N 1The condition t ∈ I in the identity (11) is not a mistake. Although K(x, t) is defined on Ω × Ω, the functions Pn(t) are defined, in general, only on I. BILINEAR BIORTHOGONAL EXPANSIONS 5 with The convergence is uniform in every set where kK(x,·)kL2(I,dµ) is bounded. Proof. Let us start by proving (11). Since K(x,·) ∈ L2(I, dµ) for every x ∈ Ω, as {Pn}n∈N and {Qn}n∈N are a complete biorthogonal system on L2(I, dµ), we can write cn(f ) =ZΩ f (t)Tn(t) dµ(t). K(x, t) = Xn∈N bn(x)Pn(t) with (by (10)) bn(x) =ZI K(x, t)Qn(t) dµ(t) = eK(χI Qn)(x) = Sn(x). Now, let us prove the biorthogonality of {Sn}n∈N and {Tn}n∈N . By definition and the multiplicative formula, we have ZΩ SnTm dµ =ZΩ SnK(χI Pm) dµ =ZΩ K(Sn)χI Pm dµ =ZI QnPm dµ = δn,m. Finally, for f ∈ P, by applying (11), interchanging the sum and the integral, and using the multiplicative formula, we have f (x) =ZI u(t)K(x, t) dµ(t) = Xn∈N(cid:16)ZI u(t)Pn(t) dµ(t)(cid:17)Sn(x) = Xn∈N(cid:16)ZΩ = Xn∈N(cid:16)ZΩ (Kf )(t)χI (t)Pn(t) dµ(t)(cid:17)Sn(x) f (t)K(χI Pn)(t) dµ(t)(cid:17)Sn(x) = Xn∈N(cid:16)ZΩ and the proof is finished. f (t)Tn(t) dµ(t)(cid:17)Sn(x) (cid:3) 3. The Fourier kernel As an example to clarify the technique, and to show how useful the use of the biorthogonal sequences is, let us look at (1) and (3) in the light of the above scheme. 3.1. The classical sampling formula. With the notation of the above section, take dµ(x) = dx, Ω = R, I = [−1, 1] and the kernel K(x, t) = 1√2π eixt, so the operator K is the Fourier transform. The space P becomes the classical Paley-Wiener space P W . Now, take N = Z and the functions Pn(t) = Qn(t) = φn(t) = 1 √2 eiπnt, n ∈ Z. Then, Sn(x) is eixt √2π From this expression, by using (11) we obtain (1). −1 e−iπnt √2 Sn(x) = eK(χI Qn)(x) =Z 1 πZR sin(x − πn) x − πn 1 Moreover, using the identity f (x) dx = f (πn), dt = sin(x − πn) √π(x − πn) . f ∈ P W, 6 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA and (12), we deduce the classical Whittaker-Shannon-Kotel'nikov sampling theorem f (x) = ∞Xn=0 sin(x − πn) x − πn f (πn). 3.2. Gegenbauer's plane wave expansion. As in the previous case, take dµ(x) = dx, Ω = R, I = [−1, 1], the kernel K(x, t) = 1√2π eixt, K the Fourier transform, and P = P W . n (t) to denote the Gegenbauer poly- nomial of order β > −1/2 (with the usual trick of employing the Chebyshev polynomials if β = 0), take the biorthonormal system But, this time, consider N = N ∪ {0} and, using C β n (t), Pn(t) = C β Qn(t) = (1 − t2)β−1/2C β n (t)/hn C β n (t)2(1 − t2)β−1/2 dt = π1/2Γ(β + 1/2)Γ(2β + n) Γ(β)Γ(2β)(n + β)n! . with hn =Z 1 −1 Using the integral ZR e−ixtx−βJβ+n(x) dx = 2−β+1π1/2Γ(2β)(−1)ninn! Γ(β + 1/2)Γ(2β + n) (1 − t2)β−1/2C β n (t)χ[−1,1](t) (see [11, Ch. 3.3, (9), p. 123]) we deduce that Sn(x) = 2β−1/2π−1/2inΓ(β)(β + n)x−β Jβ+n(x). Then, (11) becomes (3). Moreover, every function f ∈ P W admits an expansion in a uniformly convergent Fourier-Neumann series of the form f (x) = 2β−1/2π−1/2Γ(β) cn(f )in(β + n)x−β Jβ+n(x), ∞Xn=0 f (t)K(χ[−1,1]C β n )(t) dt, cn(f ) =ZR with corresponding to the expansion (12). A more explicit expression (see (32)) for the coeffi- cients cn(f ) will be given in the next section for some values of β. Finally, since kK(x,·)kL2(I,dx) = 1 √2π(cid:13)(cid:13)eix·(cid:13)(cid:13)L2([−1,1],dx) = 1 √π , Remark 2 automatically ensure that the above mentioned expansions converge uniformly on the real line. 4. The Dunkl kernel on the real line 4.1. The Dunkl transform. For α > −1, let Jα denote the Bessel function of order α and, for complex values of the variable z, let Iα(z) = 2αΓ(α + 1) Jα(iz) (iz)α = Γ(α + 1) (z/2)2n n! Γ(n + α + 1) ∞Xn=0 (Iα is a small variation of the so-called modified Bessel function of the first kind and order α, usually denoted by Iα; see [26]). Moreover, let us take Eα(z) = Iα(z) + z 2(α + 1) Iα+1(z), z ∈ C. BILINEAR BIORTHOGONAL EXPANSIONS 7 The Dunkl operators on Rn are differential-difference operators associated with some finite reflection groups (see [7]). We consider the Dunkl operator Λα, α ≥ −1/2, associated with the reflection group Z2 on R given by (13) Λαf (x) = d dx f (x) + 2α + 1 x (cid:18) f (x) − f (−x) 2 (cid:19) . For α ≥ −1/2 and λ ∈ C, the initial value problem (14) (Λαf (x) = λf (x), f (0) = 1 x ∈ R, (15) has Eα(λx) as its unique solution (see [8] and [18]); this function is called the Dunkl kernel. For α = −1/2, it is clear that Λ−1/2 = d/dx, and E−1/2(λx) = eλx. xα+1 xi(cid:19) . Let dµα(x) = (2α+1Γ(α + 1))−1x2α+1 dx and write Eα(ix) = 2αΓ(α + 1)(cid:18) Jα(x) xα + Jα+1(x) In a similar way to the Fourier transform (which is the particular case α = −1/2), the Dunkl transform of order α ≥ −1/2 is given by (16) f (x)Eα(−iyx) dµα(x), y ∈ R, Fαf (y) =ZR for f ∈ L1(R, dµα). By means of the Schwartz class S(R), the definition is extended to L2(R, dµα) in the usual way. In [18], it is shown that Fα is an isometric isomorphism on L2(R, dµα) and that for functions such that f,Fαf ∈ L1(R, dµα). formula From Fubini's theorem, it follows that the Dunkl transform satisfies the multiplication F−1 α f (y) = Fαf (−y) (17) ZR u(y)Fαv(y) dµα(y) =ZR Fαu(y)v(y) dµα(y). Finally, let us take into account that the Dunkl transform Fα can also be defined in L2(R, dµα) for −1 < α < −1/2, although the expression (16) is no longer valid for f ∈ L1(R, dµα) in general. However, it preserves the same properties in L2(R, dµα); see [20] for details. This allows us to extend our study to the case α > −1. 4.2. The sampling theorem related to the Dunkl transform. In our general setup developed in subsection 2.2, let us start by taking Ω = R, I = [−1, 1], dµ = dµα, with α > −1, and L2(I, dµ) = L2([−1, 1], dµα). On this space, we consider the kernel K(x, t) = Eα(ixt), so the corresponding operator is K = Fα, i.e., the abovementioned Dunkl transform. Now, as usual in sampling theory, we take the space of Paley-Wiener type that, in our setting, is defined as (18) P Wα =(cid:26)f ∈ L2(R, dµα) : f (x) =Z 1 −1 u(t)Eα(ixt) dµα(t), u ∈ L2([−1, 1], dµα)(cid:27) endowed with the norm of L2(R, dµα). (This space is characterized in [2, Theorem 5.1] as being the space of entire functions of exponential type 1 that belong to L2(R, dµα) when restricted to the real line.) Then, take, of course P = P Wα. It is well-known that the Bessel function Jα+1(x) has an increasing sequence of positive zeros {sn}n≥1. Consequently, the real function Im(Eα(ix)) = x 2(α+1) Iα+1(ix) is odd and it has an infinite sequence of zeros {sn}n∈Z (with s−n = −sn and s0 = 0). Then, following [6] (or [5]), let us define the functions (19) eα,n(t) = dnEα(isnt), n ∈ Z, t ∈ [−1, 1], L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA 8 where dn = 2α/2(Γ(α + 1))1/2 Iα(isn) , n 6= 0, d0 = 2(α+1)/2(Γ(α + 2))1/2. With this notation, the sequence of functions {eα,n}n∈Z is a complete orthonormal system in L2([−1, 1], dµα), for α > −1. Thus, let us take N = Z and Pn(t) = Qn(t) = eα,n(t). On the other hand, let us use that, for x, y ∈ R, x 6= y and α > −1, we have Z 1 −1 (20) Eα(ixt)Eα(iyt) dµα(t) = (the proof can be found in [3] or [6]). Then, 1 xIα+1(ix)Iα(iy) − yIα+1(iy)Iα(ix) 2α+1Γ(α + 2) x − y Sn(x) = eK(χ[−1,1]Qn)(x) =Z 1 dn −1 2α+1Γ(α + 2) = = dn xIα+1(ix)Iα(isn) x − sn Eα(ixt)eα,n(t) dµα(t) xIα+1(ix)Iα(isn) − snIα+1(isn)Iα(ix) 2α+1Γ(α + 2) because Iα+1(isn) = 0. Consequently, eα,n(t) Eα(ixt) =Xn∈Z x − sn dn 2α+1Γ(α + 2) xIα+1(ix)Iα(isn) x − sn = Iα+1(ix) + Xn∈Z\{0} Eα(isnt) 2(α + 1)Iα(isn) xIα+1(ix) x − sn , which corresponds to the formula (11) in Theorem 1. Finally, the formula (12) in Theorem 1 says that, if f ∈ P Wα, then f has the represen- tation (21) f (x) = f (s0)Iα+1(ix) + Xn∈Z\{0} f (sn) xIα+1(ix) 2(α + 1)Iα(isn)(x − sn) , that converges in the norm of L2(R, dµα). This is so because the coefficients cn(f ) in (12) are cn(f ) = dnf (sn), as we can see in what follows: where in the last step we have used that f ∈ P Wα. Moreover, by using L'Hopital rule in (20), it is not difficult to check that −1 cn(f ) =ZR =ZR =Z 1 = dnZ 1 (x·)α (cid:13)(cid:13)(cid:13)(cid:13) Eα(x·) 2 (cid:13)(cid:13)(cid:13)(cid:13) f (t)Fα(χ[−1,1]eα,n)(t) dµα(t) eα,n(x)Eα(−ixt) dµα(x) dµα(t) f (t)Eα(−ixt) dµα(t) dµα(x) −1 f (t)Z 1 eα,n(x)ZR Eα(isnx)ZR −1 f (t)Eα(−ixt) dµα(t) dµα(x) = dnf (sn), L2([−1,1],dµα) =Z 1 −1 Eα(ixr)2 dµα(r) 2α+1Γ(α + 2)(cid:18) x2 − (2α + 1)Iα+1(ix)Iα(ix) + 2(α + 1)I 2 2(α + 1)I 2 α+1(ix) = 1 α(ix)(cid:19), BILINEAR BIORTHOGONAL EXPANSIONS 9 and this norm is bounded on every compact set on the real line. So, by Remark 2, the series (21) converges uniformly in compact subsets of R. (21) is the sampling theorem related to the Dunkl transform that has been established in [6]. 4.3. Fourier-Neumann type expansion. Following [9, Definition 1.5.5, p. 27], let us introduce the generalized Gegenbauer polynomials C (λ,ν) (t) for λ > −1/2, ν ≥ 0 and n ≥ 0 (the case ν = 0 corresponding to the ordinary Gegenbauer polynomials); actually, for convenience with the notation of this paper, we are going to use C (β+1/2,α+1/2) (x). In this way, for β > −1 and α ≥ −1/2, the generalized Gegenbauer polynomials are defined by n n (22) (23) C (β+1/2,α+1/2) 2n C (β+1/2,α+1/2) 2n+1 (t) = (−1)n (α + β + 1)n (t) = (−1)n (α + β + 1)n+1 (α + 1)n+1 (α + 1)n P (α,β) n (1 − 2t2), tP (α+1,β) n (1 − 2t2), where in the coefficients we are using the Pochhammer symbol (a)n = a(a + 1)··· (a + n − 1) = Γ(a + n)/Γ(a). Note that there is no problem in extending the definition of the generalized Gegenbauer polynomials taking α > −1, so we will assume this situation. From the L2-norm of the Jacobi polynomials (see [12, Ch. 16.4, (5), p. 285]), it is easy to find (24) (25) h(β,α) 2n =Z 1 −1hC (β+1/2,α+1/2) 2n (t)i2 (1 − t2)β dµα(t) Γ(α + 1)Γ(β + n + 1)Γ(α + β + n + 1) 2α+1 (α + β + 2n + 1)Γ(α + β + 1)2Γ(α + n + 1) n! h(β,α) 2n+1 =Z 1 −1hC (β+1/2,α+1/2) 2n+1 (t)i2 (1 − t2)β dµα(t) Γ(α + 1)Γ(β + n + 1)Γ(α + β + n + 2) 2α+1 (α + β + 2n + 2)Γ(α + β + 1)2Γ(α + n + 2) n! 1 1 = = , . Finally, given α > −1, we define the functions Jα,n(x) by Jα,n(x) = Jα+n+1(x) xα+1 , x ∈ R, n = 0, 1, 2, . . . ; as these functions arise in Fourier-Neumann series, we will allude to Jα,n(x) using the name of Neumann functions.2 From the identities Z ∞ 0 Ja(x)Jb(x) x sin((b − a)π/2) 2 π Ja(x)2 x dx = Z ∞ 0 b2 − a2 1 dx = 2a , a > 0, , a > 0, b > 0, a 6= b, and taking into account that Jα,n(x) is even or odd according to n being even or odd, respectively, it follows that {Jα,n}n≥0 is an orthogonal system on L2(R, dµα), namely, ZR Jα,n(x)Jα,m(x) dµα(x) = δn,m 2α+1Γ(α + 1)(α + n + 1) . Generalized Gegenbauer polynomials and Neumann functions are the main ingredients for obtaining the Dunkl analogue of Gegenbauer's expansion of the plane wave. To establish this result we need a relation between them. By using the notation (26) (27) n P (α,β) Q(α,β) n (t) = C (β+1/2,α+1/2) (t), n (t) =(cid:0)h(β,α) n (cid:1)−1(1 − t2)βC (β+1/2,α+1/2) n (t) 2In the literature, the name "Neumann functions" is sometimes used for the Bessel functions of the second kind Yα(x), but these functions will not arise in this paper. 10 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA (where h(β,α) moreover, can have independent interest: n is given in (24) and (25)), this relation is given in the following lemma that, Lemma 1. Let α, β > −1, α + β > −1, and k = 0, 1, 2, . . . . The Dunkl transform of order α of Jα+β,k(x) is (28) Fα(Jα+β,k)(t) = 2α+β+1Γ(α + β + 1)(α + β + k + 1) Q(α,β) k (−i)k (t)χ[−1,1](t). Moreover, if β < 1, we have3 (29) Fα( · 2βJα+β,k)(t) = 2βΓ(α + β + 1) Γ(α + 1) (−i)kP (α,β) k (t), t ∈ [−1, 1]. Remark 3. There is a delicacy with the formulas in Lemma 1. Actually, Fα was defined, as a first step, as a Lebesgue integral for suitably integrable functions. Then, Fα is extended to Lp spaces where the integral representation is no longer valid for some functions. Now, the integrals Z ∞ 0 x−λJµ(ax)Jν (bx) dx (31) with (32) from [12, Ch. 8.11] or [26, Ch. XIII] that we will use in the proof are improper Riemann integrals. Hence, the proper understanding of those integrals should be as N→∞Z N lim 0 x−λJµ(ax)Jν (bx) dx. Since Jα+β,kχ[−N,N ] and · 2βJα+β,kχ[−N,N ] are integrable functions, the integral form of Fα is valid here and (28) and (29) can be understood as lim N→∞Fα(Jα+β,kχ[−N,N ])(t), lim N→∞Fα( · 2βJα+β,kχ[−N,N ])(t), and the identities in the lemma hold in the almost everywhere sense. Finally, the L2 boundedness of Fα allows us to understand these identities in the L2 sense. From now on, we will no longer mention these details. We postpone the proof of Lemma 1 to subsection 5.2. With Lemma 1, we already have all the tools for proving: Theorem 2. Let α, β > −1 and α + β > −1. Then for each x ∈ R the following expansion holds in L2([−1, 1], dµα): (30) Eα(ixt) = 2α+β+1Γ(α + β + 1) in(α + β + n + 1)Jα+β,n(x)C (β+1/2,α+1/2) (t). n ∞Xn=0 Moreover, for β < 1 and f ∈ P Wα, we have the orthogonal expansion f (x) = an(f )(α + β + n + 1)Jα+β,n(x) ∞Xn=0 an(f ) = 2α+β+1Γ(α + β + 1)ZR f (t)Jα+β,n(t) dµα+β(t). Furthermore, the series converges uniformly in compact subsets of R. 3Observe that nothing is said about outside the interval [−1, 1]; this does not mean that this expression vanishes for t > 1, that is not true when β 6= 0. BILINEAR BIORTHOGONAL EXPANSIONS 11 Proof. In the biorthogonal setup given in subsection 2.2, let Ω = R, I = [−1, 1], the space L2(I, dµ) = L2([−1, 1], dµα), and the kernel K(x, t) = Eα(ixt), from which the operator K becomes the Dunkl transform Fα (and eK = F−1 α ). Also, consider the Paley-Wiener space P = P Wα (see (18)). Finally, for N = N ∪ {0}, take the biorthonormal system given by Pn(t) = P (α,β) (t) and Qn(t) = Q(α,β) Sn(x) = 2α+β+1Γ(α + β + 1)in(α + β + n + 1)Jα+β,n(x). (t) as in (26) and (27). From (28), we have n n In this situation, the formula (11) in Theorem 1 gives (30). Now, let us consider Then, the identity given by (12) becomes Tn(x) = K(χ[−1,1]Pn)(x) = Fα(χ[−1,1]P (α,β) n )(x). f (x) = 2α+β+1Γ(α + β + 1) cn(f )in(α + β + n + 1)Jα+β,n(x), f ∈ P Wα, with ∞Xn=0 cn(f ) =ZR f (t)Fα(χ[−1,1]P (α,β) cn(f ) = (−i)nZR n f (t)Jα+β,n(t) dµα+β(t), )(t) dµα(t). cn(f ) =Z 1 −1 u(x)P (α,β) n (x) dµα(x). Let us see that, when β < 1, the coefficient cn(f ) can be written as which implies (32) with an(f ) = 2α+β+1Γ(α + β + 1)incn(f ). Indeed, if we consider u such that f = F−1 α (uχ[−1,1]) and use the multiplication for- mula (17), we can write Now, by (29) and the multiplication formula again, we have cn(f ) =Z 1 −1 inΓ(α + 1) u(x) 2βΓ(α + β + 1)Fα( · 2βJα+β,n)(x) dµα(x) It is clear that Fα(uχ[−1,1])(t) = f (−t), so the change of variable from t to −t gives = inZR Fα(uχ[−1,1])(t)Jα+β,n(t) dµα+β(t). cn(f ) = (−i)nZR because Jα+β,n(−t) = (−1)nJα+β,n(t). Remark 4. Actually, formulas (3) and (30) are equivalent for α ≥ −1/2. The proof in one direction is clear, just by specializing the parameters. To obtain (30) from (3), we can use the intertwining operator f (t)Jα+β,n(t) dµα+β(t) (cid:3) Vαg(t) = Γ(2α + 2) 22α+1Γ(α + 1/2)Γ(α + 3/2)Z 1 −1 g(st)(1 − s)α−1/2(1 + s)α+1/2 ds (see [9, Definition 1.5.1, p. 24], we change the parameter µ in the definition given in [9] by α + 1/2), defined for α ≥ −1/2. With this notation we have (t) = C (β+1/2,α+1/2) (t) (33) VαC α+β+1 n n and Vαei·(t) = Eα(it). In this way, applying Vα to (3) (with α+β +1 instead of β) we get (30). This idea has been used for the higher rank in [21], where the author assumes that (3) was already known, 12 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA and then (30) is established for α ≥ −1/2 by using the intertwining operator. This gives a considerably shorter proof. Instead, with the method followed in the proof of Theorem 2, the identity (30) not only can be found for α > −1, but also is proved directly and then, as a particular case, (3) holds. Remark 5. Another way of obtaining (30) is as follows. Some results from [13] were generalized in [25] to When z and w are replaced by zγ and w/γ, respectively, and we let γ → ∞, we get the companion formula ambm (zw)m m! = ∞Xm=0 (−z)n n! (γ + n)n ∞Xr=0 ∞Xn=0 (−n)s(n + γ)s s! asws! . bn+rzr r! (γ + 2n + 1)r! nXs=0 zj! nXk=0 n! ∞Xj=0 bn+j j! (−z)n (−n)kakwk! ambm(zw)m = ∞Xm=0 ∞Xn=0 (these formulas are also stated in [16, Ch. 9]). These expansions contain several expan- sions in terms of Jacobi polynomials of argument 1 − 2t2 (i.e., generalized Gegenbauer polynomials). In particular, (30) follows in this way. 4.4. Consequences for the Hankel transform. For α > −1, consider the so-called modified Hankel transform Hα, that is (34) Hαf (y) =Z ∞ 0 Jα(xy) (xy)α f (y)x2α+1 dx, x > 0. The kernel Eα(ixt) of the Dunkl transform (16) can be written in terms of the Bessel functions of order α and α + 1, and this clearly allows us to study the Hankel transform as a simple consequence of the Dunkl transform. In particular, if we have a function f ∈ L2((0,∞), x2α+1 dx), we can take the even extension f (·) ∈ L2(R, dµα). Then, using that Jα(x)/xα is even and Jα+1(x)/xα is odd, we write (34) as The Paley-Wiener space for the Hankel transform is given by Hαf (y) = Fα(f ( · ))(y). P W ′α =(f ∈ L2((0,∞), x2α+1 dx) : f (t) =Z 1 0 u(x) Jα(xt) (xt)α x2α+1 dx, u ∈ L2((0, 1), x2α+1 dx)); also, note that if f ∈ P W ′α, then both the even extension f ( · ) and the odd extension sgn(·)f ( · ) belong to P Wα. So, let us adapt the sampling formula of subsection 4.2 and the Theorem 2 of subsec- tion 4.3 to the context of the Hankel transform. 4.4.1. The sampling formula for the Hankel transform. For f ∈ P W ′α, taking its even extension f ( · ), using that s−n = −sn and grouping the summands corresponding to 1/(x − sn) and 1/(x + sn) in (21), we get f (x) = f (s0)Iα+1(ix) + Iα+1(ix) f (sn) x2 . (α + 1)Iα(isn) x2 − s2 n ∞Xn=1 Similarly, with the odd extension of f , (21) becomes f (x) = ∞Xn=1 f (sn) Iα+1(ix) (α + 1)Iα(isn) s2 n x2 − s2 n . BILINEAR BIORTHOGONAL EXPANSIONS 13 The latter identity corresponds to the well-known Higgins sampling theorem for the Hankel transform [15]. 4.4.2. A version of the Theorem 2 for the Hankel transform. Let us observe that Jα(xt) (xt)α = 1 2α+1Γ(α + 1)(cid:0)Eα(ixt) + Eα(ixt)(cid:1). From this, it is very easy to adapt (30) to the new context, and to write it in terms of Jacobi polynomials by using (22). Given f ∈ P W ′α, let us consider its even extension f ( · ) ∈ P Wα. Applying (31) to this even function, it becomes an expansion that only contains Jα+β,2n(x) = Jα+β+2n+1(x)/xα+β+1 (i.e., only with even indexes). Thus, the results corresponding to the Hankel transforms can be summarized in this way: Corollary 3. Let α, β > −1 and α + β > −1. Then for each x ∈ (0,∞) the following expansion holds in L2((0, 1), x2α+1 dx): (35) Jα(xt) (xt)α = 2β+1(α + β + 2n + 1)Γ(α + β + n + 1) Γ(α + n + 1) Jα+β,2n(x)P (α,β) n (1 − 2t2). ∞Xn=0 Moreover, for β < 1 and f ∈ P W ′α, we have the orthogonal expansion with f (x) = an(f )(α + β + 2n + 1)Jα+β,2n(x) ∞Xn=0 an(f ) = 2Z ∞ 0 f (t)Jα+β,2n(t) t2α+2β+1 dt. Furthermore, the series converges uniformly in compact subsets of (0,∞). Remark 6. In the particular case α = −1/2, on using J−1/2(z) = 21/2π−1/2z−1/2 cos(z) and (22), the formula (35) becomes (36) xβ+1/2 2β+1/2Γ(β + 1/2) cos(xt) = ∞Xn=0 (−1)n(2n + β + 1/2)Jβ+2n+1/2(x)C β+1/2 2n (t), which is already known (see [26, § 11.5, formula (5), p. 369]). On the other hand, following the procedure described in Remark 4, from this expression we can obtain another proof of (35), valid for α > −1/2. Let us assume that (36) is already known, and we write it with α + β + 1/2 instead of β; then, by applying the intertwining operator Vα and using (33), (22) and we get the identity (35), as desired. Vα cos(t) = 2αΓ(α + 1) Jα(t) tα , Let us conclude by observing that the Lp convergence of the orthogonal series that appear in the previous corollary has been studied in the papers [24, 4] for functions in an appropriate Lp extension of the Paley-Wiener space. 5. Technical lemmas The main goal of this section is to prove Lemma 1, which is key in our study of the Dunkl transform on the real line. The proof is contained in subsection 5.2. With this target, we need to previously establish some formulas. They are given in subsection 5.1. 14 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA 5.1. Some integrals involving Bessel functions. For the sake of completeness, let us start by proving two identities that express some integrals involving the product of two Bessel functions in terms of Jacobi polynomials. Such integrals are usually written in terms of hypergeometric functions; however their expressions as Jacobi polynomials are not easily found in the literature. For instance, they do not appear in the standard references [10, 26, 12]. In what follows, we can take into account the comments in Remark 3, but we will not repeat them. Lemma 2. For α, β > −1 with α + β > −1, and n = 0, 1, 2, . . . , let us define I−(α, β, n)(t) = t−αZ ∞ I+(α, β, n)(t) = t−αZ ∞ 0 0 x−βJα+β+2n+1(x)Jα(xt) dx, xβJα+β+2n+1(x)Jα(xt) dx. Then, we have (37) I−(α, β, n)(t) = 2−β Γ(n+1) Γ(β+n+1) (1 − t2)βP (α,β) n (1 − 2t2)χ[0,1](t), t ∈ (0,∞). Assume further that β < 1; then, (38) I+(α, β, n)(t) = 2β Γ(α+β+n+1) Γ(α+n+1) P (α,β) n (1 − 2t2), t ∈ (0, 1). Proof. We use the formula (39) Z ∞ 0 x−λJµ(ax)Jν (bx) dx = bν aλ−ν−1Γ( µ+ν−λ+1 ) 2λΓ(ν + 1)Γ( λ+µ−ν+1 2 ) 2 2F1(cid:18)µ + ν − λ + 1 2 , ν − λ − µ + 1 2 ; ν + 1; b2 a2(cid:19) , valid when 0 < b < a and −1 < λ < µ + ν + 1; here, 2F1 denotes the hypergeometric function (see [12, Ch. 8.11, (9), p. 48] or [26, Ch. XIII, 13.4 (2), p. 401]). Then, let us start with (37). Taking a = 1 and t = b in (39), and making the corre- sponding changes of variable and parameters (ν = α, µ = α + β + 2n + 1, λ = β) we get Γ(α + n + 1) I−(α, β, n)(t) = 2βΓ(α + 1)Γ(β + n + 1) 2F1(α + n + 1,−n − β; α + 1; t2), which is valid for α > −1 and β > −1 in the interval 0 < t < 1. Moreover, we have 2F1(−n, α + β + n + 1; α + 1; t), 2F1(α + n + 1,−n − β; α + 1; t) = (1 − t)β where α, β > −1, n = 0, 1, 2, . . . , and (y) = Γ(n+α+1) (40) whenever α, β > −1 and −1 < y < 1. Therefore, P (α,β) n Γ(α+1)Γ(n+1) 2F1(−n, α + β + n + 1; α + 1; 1−y 2 ), I−(α, β, n)(t) = 2−β Γ(n+1) Γ(β+n+1) (1 − t2)βP (α,β) n (1 − 2t2), t ∈ (0, 1). Now, we are going to evaluate I−(α, β, n)(t) for t > 1. To do that, let us take t = a, b = 1, ν = α + β + 2n + 1, µ = α, and λ = β in (39). In this way, 1 2 (λ + µ− ν + 1) = 0,−1,−2, . . . , so the coefficient 1/Γ( 1 Finally, let us prove the second part of the lemma. To this end, we take, in (39), a = 1 and t = b, with parameters λ = −β, µ = α + β + 2n + 1 and ν = α. Then, for β < 1, α + β > −1, and 0 < t < 1 we get 2 (λ + µ − ν + 1)) vanishes and we get I−(α, β, n)(t) = 0. I+(α, β, n)(t) = 2βΓ(α + β + n + 1) Γ(α + 1)Γ(n + 1) 2F1(α + β + n + 1,−n; α + 1; t2). Then, by using (40), (38) follows. (cid:3) BILINEAR BIORTHOGONAL EXPANSIONS 15 5.2. Proof of Lemma 1. We start by evaluating Fα(Jα+β,k)(t) for α > −1 and α + β > −1. By definition, Jα+β+k+1(x) 1 2ZR Fα(Jα+β,k)(t) = For the case k = 2n, by decomposing into even and odd functions, we can write (xt)α+1 xti(cid:19)x2α+1 dx. Jα(xt) (xt)α x2α+1 dx. Then, for t > 0, by using (37) in Lemma 2, (22) and (24), it follows that Fα(Jα+β,2n)(t) =Z ∞ (cid:18) Jα(xt) (xt)α − Jα+β+2n+1(x) Jα+1(xt) xα+β+1 (41) 0 xα+β+1 Fα(Jα+β,2n)(t) = t−αZ ∞ 0 x−βJα+β+2n+1(x)Jα(xt) dx = Γ(n + 1) 2βΓ(β + n + 1) = (−1)n Γ(α + β + 1)Γ(n + 1)Γ(α + n + 1) 2βΓ(α + 1)Γ(β + n + 1)Γ(α + β + n + 1) ik n (1 − t2)βP (α,β) (1 − 2t2)χ[0,1](t) (1 − t2)βC (β+1/2,α+1/2) 2α+β+1Γ(α + β + 1)(α + β + k + 1) Q(α,β) 2n k = (t)χ[0,1](t) (t)χ[0,1](t). For t < 0, let us make, in (41), the change t1 = −t, use the evenness of the function Jα(z)/zα, proceed as in the case t > 0, and undo the change. Then, we get Fα(Jα+β,2n)(t) = 2α+β+1Γ(α + β + 1)(α + β + k + 1) Q(α,β) k ik (t)χ[−1,0](t). Thus, (28) for even k is proved. The case k = 2n + 1 is completely similar. Proceeding in the same way, the formula (29) follows from (38). Acknowledgement. The authors whish to thank the referee for careful reading of the manuscript, comments and the suggestion of Remark 6. References [1] L. D. Abreu, The reproducing kernel structure arising from a combination of continuous and discrete orthogonal polynomials into Fourier systems, Constr. Approx. 28 (2008), no. 2, OF1 -- OF17. [2] N. B. Andersen and M. de Jeu, Elementary proofs of Paley-Wiener theorems for the Dunkl transform on the real line, Int. Math. Res. Not. 2005, no. 30, 1817 -- 1831. [3] J. Betancor, ´O. Ciaurri, and J. L. Varona, The multiplier of the interval [−1, 1] for the Dunkl transform on the real line, J. Funct. Anal. 242 (2007), 327 -- 336. [4] ´O. Ciaurri, J. J. Guadalupe, M. P´erez, and J. L. Varona, Mean and almost everywhere convergence of Fourier-Neumann series, J. Math. Anal. Appl. 236 (1999), 125 -- 147. [5] ´O. Ciaurri, M. P´erez, J. M. Reyes, and J. L. Varona, Mean convergence of Fourier-Dunkl series, J. Math. Anal. Appl. 372 (2010), 470 -- 485. [6] ´O. Ciaurri and J. L. Varona, A Whittaker-Shannon-Kotel'nikov sampling theorem related to the Dunkl transform, Proc. Amer. Math. Soc. 135 (2007), 2939 -- 2947. [7] C. F. Dunkl, Differential-difference operators associated with reflections groups, Trans. Amer. Math. Soc. 311 (1989), 167 -- 183. [8] C. F. Dunkl, Integral kernels with reflections group invariance, Canad. J. Math. 43 (1991), 1213 -- 1227. [9] C. F. Dunkl and Y. Xu, Orthogonal polynomials of several variables, Encyclopedia of Mathematics and its Applications No. 81, Cambridge University Press, 2001. [10] A. Erd´elyi, W. Magnus, F. Oberhettinger, and F. Tricomi, Higher Transcendental Functions, Vol. II, McGraw-Hill, New York, 1953. [11] A. Erd´elyi, W. Magnus, F. Oberhettinger, and F. Tricomi, Tables of Integral Transforms, Vol. I, McGraw-Hill, New York, 1954. [12] A. Erd´elyi, W. Magnus, F. Oberhettinger, and F. Tricomi, Tables of Integral Transforms, Vol. II, McGraw-Hill, New York, 1954. [13] J. L. Fields and J. Wimp, Expansions of hypergeometric functions in hypergeometric functions, Math. Comp. 15 (1961), 390 -- 395. 16 L. D. ABREU, ´O. CIAURRI, AND J. L. VARONA [14] A. G. Garc´ıa, Orthogonal sampling formulas: a unified approach, SIAM Rev. 42 (2000), no. 3, 499 -- 512. [15] J. R. Higgins, An interpolation series associated with the Bessel-Hankel transform, J. Lond. Math. Soc. 5 (1972), 707 -- 714 [16] M. E. H. Ismail, Classical and quantum orthogonal polynomials in one variable, Encyclopedia of Math- ematics and its Applications No. 98, Cambridge University Press, 2005. [17] M. E. H. Ismail and R. Zhang, Diagonalization of certain integral operators, Adv. Math. 109 (1994), no. 1, 1 -- 33. [18] M. F. E. de Jeu, The Dunkl transform, Invent. Math. 113 (1993), 147 -- 162. [19] H. P. Kramer, A generalized sampling theorem, J. Math. and Phys. 38 (1959/60), 68 -- 72. [20] M. Rosenblum, Generalized Hermite polynomials and the Bose-like oscillator calculus, Oper. Theory Adv. Appl. 73 (1994), 369 -- 396. [21] M. Rosler, A positive radial product formula for the Dunkl kernel, Trans. Amer. Math. Soc. 355 (2003), 2413 -- 2438. [22] S. Saitoh, One approach to some general integral transforms and its applications, Integral Transform. Spec. Funct. 3 (1995), no. 1, 49 -- 84. [23] S. Saitoh, Integral transforms, Reproducing Kernels and Their Applications, Pitman Research Notes in Mathematics Series 369, Addison Wesley Longman, Harlow, 1997. [24] J. L. Varona, Fourier series of functions whose Hankel transform is supported on [0, 1], Constr. Approx. 10 (1994), 65 -- 75. [25] A. Verma, Some transformations of series with arbitrary terms, Ist. Lombardo Accad. Sci. Lett. Rend. A 106 (1972), 342 -- 353. [26] G. N. Watson, A Treatise on the Theory of Bessel Functions (2nd edition), Cambridge Univ. Press, Cambridge, 1944. [27] A. I. Zayed, Advances in Shannon's sampling theory, CRC Press, Boca Raton, FL, 1993. [28] A. I. Zayed, A generalization of the prolate spheroidal wave functions, Proc. Amer. Math. Soc. 135 (2007), no. 7, 2193 -- 2203. CMUC and Departamento de Matem´atica, Universidade de Coimbra, Faculdade de Ciencias e Tecnologia (FCTUC), 3001-454 Coimbra, Portugal E-mail address: [email protected] CIME and Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, 26004 Logrono, Spain E-mail address: [email protected] CIME and Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, 26004 Logrono, Spain E-mail address: [email protected] URL: http://www.unirioja.es/cu/jvarona/
1105.3610
2
1105
2011-05-24T13:35:05
On the closed subideals of $L(\ell_p\oplus\ell_q)$
[ "math.FA" ]
In this paper we first review the known results about the closed subideals of the space of bounded operator on $\ell_p\oplus \ell_q$, $1<p<q<\infty$, and then construct several new ones.
math.FA
math
ON THE CLOSED SUBIDEALS OF L(ℓp ⊕ ℓq) TH. SCHLUMPRECHT Abstract. In this paper we first review the known results about the closed subideals of the space of bounded operator on ℓp ⊕ ℓq, 1 < p < q < ∞, and then construct several new ones. 1. Introduction For very few Banach spaces X all the closed subideals of L(X), the algebra of all bounded and linear operators on X, are determined. In 1941 Calkin [6] showed that the only proper, non-trivial and closed ideal of L(ℓ2) is the ideal of compact operators. The same was shown to be true for ℓp (1 ≤ p < ∞) and c0 in [13]. Until very recently it was open if there are any other infinite dimensional Banach spaces X, for which the compact operators are the only proper, non-trivial and closed subideal of L(X). We call such spaces simple. Then Argyros and Haydon [3] established the existence of Banach spaces with a basis on which all operators are a compact perturbation of a scalar multiple of the identity. It follows immediately that such spaces are simple. But it is not known whether or not there are any other simple spaces admitting an unconditional basis (and thus having a rich structure of operators on them). The structure of the closed ideals of operators on non separable Hilbert spaces was inde- pendently obtained by Gramsch [14] and Luft [21]. Recently Daws [7] extended their results to non separable ℓp-spaces, 1 ≤ p < ∞, and non separable c0-spaces. . for X = (cid:0)L∞ n=1 ℓ2(n)(cid:1)c0 n=1 ℓ2(n)(cid:1)ℓ1 and in [18] for X = (cid:0)L∞ Beyond these spaces the complete structure of closed ideals in L(X) was described in [16] In both cases, there are exactly two nested proper non-zero closed ideals, namely the compacts and the closure of all operators factoring through c0, or ℓ1, respectively. Apart from those mentioned above, there are no other separable Banach spaces X for which the structure of the closed ideals in L(X) is completely known. It is still open whether or not the closed subideals of the operators on the spaces (⊕∞ n=1ℓ∞(n))ℓ1 admit the same sublattice structure (for partial results see [17]). An interesting space for studying the closed subideals of its bounded linear operators is the space X introduced in [26]. This space is complementably minimal [1], which means that every infinite dimensional closed subspace of X contains a further subspace which is complemented in X and isomorphic to X. This implies that the strictly singular operators (see the definition at the end of this section) is the only maximal proper closed subideal n=1ℓ1(n))c0 and (⊕∞ 2000 Mathematics Subject Classification. Primary: 47L20. Secondary: 47B10, 47B37. Key words and phrases. Operator ideal, ℓp-space. Research partially supported by NSF grant DMS 0856148. 1 of L(X). As shown in [2], X admits strictly singular but not compact operators, and it is conjectured that L(X) contains infinitely many closed subideals, all of which have to lie between the ideal of compact operators, and the ideal of strictly singular operators. A space whose closed ideals of operators attracted the attention of several researches is the pth quasi reflexive James Jp, with 1 < p < ∞. Edelstein and Mityagin [10] observed that the ideal of weakly compact operators on Jp is the only maximal proper subideal of L(Jp). In [20], for p = 2, and in [15], for general p ∈ (1, ∞), it was shown that the closure of the operators on Jp factoring through ℓ2 contains strictly the ideal of compact operators and is strictly contained in the ideal of weakly compact operators. Very recently Bird, Jameson and Laustsen [5] found a new closed sub ideal of L(Jp) and proved that the closure of the ideal of operators factoring through the ℓp-sum of ℓ∞(n), n ∈ N, is strictly larger than the closure of the ideal of operators factoring through ℓp and strictly smaller then the ideal of weakly compact operators. Although studied in several papers (cf.[22],[24] and [25]) the structure of the closed subide- als of L(ℓp ⊕ ℓq), 1 < p < q < ∞ remains a mystery. It is not even known whether or not L(ℓp ⊕ℓq), contains infinitely many subideals. There were several results proved in the 1970's concerning various special ideals or special cases of p and q. We refer the reader to the book by Pietsch [24, Chapter 5] for details. In particular, [24, Theorem 5.3.2] asserts that L(ℓp ⊕ ℓq), with 1 ≤ p < q, has exactly two proper maximal ideals (namely, the ideal of operators which factor through ℓp and the ideals of operators which factor through ℓq), and establishes a one-to-one correspondence between the non-maximal proper subideals of L(ℓp ⊕ ℓq) and the closed ideals in L(ℓp, ℓq). By proving that the formal identity I(p, q) : ℓp → ℓq is finitely strictly singular (see the definition at the end of this section) and establishing the existence of an operator T : ℓp → ℓq which is not finitely strictly singular Milman [22] concluded that L(ℓp, ℓq) contains at least two non trivial, proper and closed subideals. In [25] the study of the structure of the closed subideals of L(ℓp, ℓq) was continued, and, among other results, it was discovered that the lattice of subideals of L(ℓp, ℓq) is not linearly ordered, and contains at least 4 nontrivial, proper and closed subideals if 1 < p < 2 < q < ∞. In this paper we increase this number to 7. In Section 2 we will recall the known results on the closed subideals of L(ℓp ⊕ ℓq) and L(ℓp, ℓq), and sketch the proof of several of them. In Section 3 we will formulate and prove our main result (see Theorem 3.1). Let us first recall some necessary notation. If X and Y are Banach spaces, L(X, Y ) denotes the space of bounded linear operators T : X → Y , and if X = Y we write L(X) instead of L(X, X). A linear subspace J ⊂ L(X, Y ), is called a subideal of L(X, Y ), if for all A ∈ L(Y ), B ∈ L(X), and T ∈ J also A◦ T ◦ B ∈ J . A closed subideal of L(X, Y ) is a subideal which is closed in the operator norm. We say that a subideal J ⊂ L(X, Y ) is non trivial if it is not the zero ideal {0} and proper if it is not all of L(X, Y ). The following is a list of some important closed subideals of L(X, Y ). 2 F D(X, Y ) is the closure of the ideal of operators with finite dimensional rank. Note that any nontrivial closed subideal J in L(X, Y ) contains all of F D(X, Y ). This follows from the fact that J is closed under taking sums, under multiplication by elements of L(X) from the right, under multiplication from the left by elements of L(Y ), and that it must contain a non zero operator (and thus a rank 1 operator). Thus, for all infinite dimensional Banach spaces X and Y the ideal F D(X, Y ) is the minimal nontrivial closed subideal of L(X, Y ). K(X, Y ) denotes the ideal of compact operators . All the spaces we consider are spaces with a basis. Thus, these spaces have the approximation property, which means that F D(X, Y ) = K(X, Y ). StSi(X, Y ) is the closed ideal of operators T : X → Y which are strictly singular, i.e. on no infinite dimensional subspace Z of X is the restriction of T onto Z an isomorphism. F SS is the closed ideal of finitely strictly singular operators. A linear bounded operator T : X → Y , is called finitely strictly singular if for all ε > 0 there is an n = nε ∈ N so that for any n-dimensional subspace E of X, there is an x ∈ E, with kxk = 1, so that kT (x)k ≤ ε. If W and Z are Banach spaces and S : W → Z a bounded linear operator, we denote by J S(X, Y ) the closure of the ideal generated by all operators T ∈ L(X, Y ), which factor through S, thus T = A ◦ S ◦ B, with A ∈ L(Z, Y ) and B ∈ L(X, W ). In general the set {A ◦ S ◦ B, A ∈ L(Z, Y ) and B ∈ L(X, W )} is not closed under addition and therefore not an ideal. But if the operator S ⊕ S : W ⊕ W → Z ⊕ Z, (w1, w2) 7→ (S(w1), S(w2)), factors through S, then {A ◦ S ◦ B, A ∈ L(Z, Y ) and B ∈ L(X, W )} is an ideal and we conclude in that case that (1) J S(X, Y ) = {A◦S ◦B : A ∈ L(Z, Y ) and B ∈ L(X, W )}. Let I(p, q) : ℓp → ℓq be the formal inclusion (using that ℓp is a subset of ℓq), for 1 ≤ p < q ≤ ∞. It is easily seen that I(p, q) ⊕ I(p, q) factors through I(p, q) and we conclude that J I(p,q)(X, Y ) = {A◦I(p, q)◦B : A ∈ L(ℓq, Y ) and B ∈ L(X, ℓp)}. If IZ is the identity on some Banach space Z we write J Z instead of J IZ , and we note that if Z is isomorphic to Z ⊕ Z it follows that (2) J Z(X, Y ) = {A◦S ◦B : A ∈ L(Z, Y ) and B ∈ L(X, Z)}. If X = Y we will write K(X), F SS(X) etc. instead of K(X, X), F SS(X, X) etc. For 1 ≤ p < ∞, we denote the unit vector basis of ℓp = ℓp(N) by (e(p,j) : j ∈ N) (if p = ∞ we consider c0 instead of ℓ∞). The conjugate of p is denoted by p′, i.e. 1 p′ = 1. For n ∈ N we denote the n-dimensional ℓp space by ℓp(n) and its unit vector basis by (e(p,n,j) : j = 1, 2, . . . , n). The usual norm on ℓp or ℓp(n), n ∈ N is denoted by k · kp. If Xn is a Banach space for n ∈ N, the ℓp-sum of Xn, n ∈ N, is the space of all sequences (xn : n ∈ N), with xn ∈ Xn, for n ∈ N, and p + 1 k(xn)n∈Nkp =(cid:16)Xn∈N kxnkp(cid:17)1/p 3 < ∞, if p < ∞, and We denote the ℓp-sum of (Xn) by (⊕∞ n=1Xn)∞ the c0-sum, the space of all sequences (xn), with xn ∈ Xn, for n ∈ N, for which limn→∞ kxnk = 0. The sphere and the unit ball of a Banach space are denoted by SX and BX , respectively. For simplicity all our Banach spaces are defined over the real field R. It is easy to see how our results can be extended to Banach spaces over the complex field C. n=1Xn)p. If p = ∞ we denote by (⊕∞ 2. Review of the known results on the closed subideals of L(ℓp ⊕ ℓq) and L(ℓp, ℓq) We will now review the known results on the lattice structure of subideals of L(ℓp ⊕ ℓq). We will assume from now on that 1 < p < q < ∞ and later that 1 < p < 2 < q < ∞. Every operator T = ℓp ⊕ ℓq → ℓp ⊕ ℓq, consists of four operators T(1,1) ∈ L(ℓp), T(1,2) ∈ L(ℓq, ℓp) and T(2,1) ∈ L(ℓp, ℓq), and T(2,2) ∈ L(ℓp, ℓp), and acts as a 2 by 2 matrix on the elements of ℓp ⊕ ℓq T =(cid:18)T(1,1) T(1,2) T(2,1) T(2,2)(cid:19) : ℓp⊕ℓq → ℓp⊕ℓq, (x, y) 7→(cid:0)T(1,1)(x)+T(1,2)(y), T(2,1)(x)+T(2,2)(y)(cid:1). By the above cited result from [13], the operators T(1,1) and T(2,2) are either compact or the identity on ℓp, respectively ℓq, factors through them. By Pitt's Theorem (c.f. [11, Proposition 6.25]), T(1,2) is compact, and since every infinite dimensional subspace of ℓp contains a subspace isomorphic to ℓp, and since ℓp and ℓq are incomparable, we conclude that T(2,1) must be strictly singular. So, if J is a closed subideal of L(ℓp ⊕ ℓq) which contains an operator T for which T(1,1) and T(2,2) are not compact, we conclude that the identity on ℓp ⊕ ℓq factors through T and thus J = L(ℓp ⊕ ℓq). If J contains an operator for which T(1,1) is not compact, but for all elements U ∈ J , U(2,2) is compact, then the identity on ℓp factors through T , but not the identity on ℓq, and we therefore deduce that J must be the closure of the operators factoring through ℓp, which must therefore be a maximal proper subideal of L(ℓp ⊕ ℓq) (for more details see [24, Theorem 5.3.2]). Similarly we conclude that the closure of all operators factoring through ℓq is a maximal proper subideal of L(ℓp ⊕ ℓq). For all other closed proper subideals J ⊂ L(ℓp ⊕ ℓq), and all T ∈ J it therefore follows that T(1,1), T(1,2) and T(2,2) are compact, and can therefore be approximated by finite rank operators which factor through ℓp as well as ℓq. Of course T(2,1) also factors through ℓp as well as ℓq, and we deduce that all other closed ideals are subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq (ℓp ⊕ ℓq), and thus not maximal proper closed ideals. Assume now that J ⊂ J ℓp(ℓp ⊕ ℓq) ∩ J ℓp(ℓp ⊕ ℓq) is a closed ideal in L(ℓp ⊕ ℓq) An easy computation yields that J := {T(2,1) : T ∈ J } is a closed subideal of L(ℓp, ℓq), and that for two different ideals J1, J2 ⊂ J ℓp(ℓp ⊕ ℓq) ∩ J ℓp(ℓp ⊕ ℓq) the ideals J1 and J2 are different. Conversely if J is a closed subideal of L(ℓp, ℓq) then J ′ =(cid:26)(cid:18)T(1,1) T(1,2) T(2,1) T(2,2)(cid:19) : T(2,10 ∈ J and T(1,1) ∈ K(ℓp), T(1,2) ∈ K(ℓq, ℓp), and T(2,2) ∈ K(ℓq)(cid:27) 4 1 and J ′ is a closed subideal of L(ℓp ⊕ ℓq) and for two different closed subideals J1, J2 ⊂ L(ℓp, ℓq), J ′ 2 are different. Thus there is a bijection between the set of all closed subideals of L(ℓp, ℓq) and the non maximal closed subideals of L(ℓp ⊕ ℓq), which preserves the lattice structure with respect to inclusions. Let us summarize the observations we just made in the following proposition. Proposition 2.1. For 1 < p < q < ∞, the space L(ℓp ⊕ ℓq) has exactly two maximal proper closed subideals, namely J ℓp(ℓp ⊕ ℓq) and J ℓq (ℓp ⊕ ℓq). All other closed subideals of L(ℓp ⊕ ℓq), are subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq(ℓp ⊕ ℓq), and there is a bijection between the closed subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq(ℓp ⊕ ℓq) and closed subideals of L(ℓp, ℓq) which preserves the lattice structure. We are therefore interested in the closed subideals of L(ℓp, ℓq). Instead of writing K(ℓp, ℓq), F SS(ℓp, ℓq), or J S(X, Y ) etc. we will from now on simply write K, F SS or J S etc. The following diagram summarizes the results established in [22] and [25], under the assumption that 1 < p < 2 < q < ∞. K 3 J I(p,q) / F SS ∩ J ℓ2 F SS UUUU UUUU 4i i i TTTTT −k J ℓ2 4j j j F SS ∨ J ℓ2 / L(ℓp, ℓq) Here arrows stand for inclusions. A solid arrow (⇒ or →) between two ideals means that there are no other ideals sitting properly between the two, while a double arrow coming out of an ideal indicates the only immediate successor. A hyphenated arrow (−−>) indicates a proper inclusion, while a dotted one indicates that we do not know whether or not the inclusion is proper. In particular, the closed ideals in L(ℓp, ℓq) are not totally ordered. Let us explain the diagram "from the left to the right" (for a more detailed explanation we refer the reader to [25]): If T : ℓp → ℓq is not compact, then there is a normalized block sequence (xn) in ℓp whose image (yn) = (T (xn) is equivalent to (e(q,j) : j ∈ N) (the unit vector basis in ℓq) and so that span(yn : n ∈ N) is complemented in ℓp. It follows that I(p, q) factors through T , and that therefore J I(p,q) is the only successor of K. It is clear that J I(p,q) ⊂ J ℓ2 (recall that we assume that p < 2 < q). The fact that J I(p,q) ⊂ F SS follows from the following result in [22] (see also [25, Proposition 3.3]). Proposition 2.2. For any choices of 1 ≤ p < q ≤ ∞ is the formal identity I(p, q) is a finitely strictly singular operator. The way to verify Proposition 2.2 is to show first (see [22] or [25, Lemma 3.4]) by induction on n ∈ N, that in every n-dimensional subspace E of c0 there is x ∈ E which attains its sup- norm on at least n coordinates. In order to see then, that I(p, q) is finitely strictly singular, let ε > 0 and pick n ∈ N with n−(q−p)/q < ε. If E is any subspace of ℓp of dimension n we can find x ∈ E, kxkp = 1, so that kxk∞ ≤ n−1/p (since the maximum is attained on at 5 & . + / 4 * * / 4 p ≤ np−q and thus least n coordinates), and thus kxkq kxkq ≤ n−(q−p)/q ≤ ε. We therefore established that J I(p,q) ⊂ F SS ∩ J ℓ2. In Section 2 we will show that this inclusion is strict. More precisely, we will show that the ideals J I(p,2) and J I(2,q) are two distinct closed ideals which lie between J I(p,q) and F SS ∩ J ℓ2. i=1 x(i)q−px(i)p ≤ kxkq−p ∞ kxkp q =P∞ In order to show that F SS ∩ J ℓ2 is not all of L(ℓp, ℓq) Milman [22] used the fact that ℓp (and ℓq)is isomorphic the ℓp-sum (respectively the ℓq sum) of ℓ2(n), n ∈ N (see [19, page 73]). Letting U : ℓp → (⊕n∈Nℓ2(n))p and V : ℓq → (⊕n∈Nℓ2(n))q be isomorphisms and letting I ′(p, q) be the formal identity I ′(p, q) : (⊕n∈Nℓ2(n))p → (⊕n∈Nℓ2(n))q, (xn) 7→ (xn), we define T (p, q) = V ◦ I ′(p, q) ◦ U. T (p, q) depends on the choice of the isomorphisms U and V , nevertheless it is easy to see that for any other isomorphisms U : ℓp → (⊕n∈Nℓ2(n))p and V : ℓq → (⊕n∈Nℓ2(n))q, the operator T (p, q) = V ◦ I ′(p, q) ◦ U , factors through T (p, q) and vice versa, and thus J T (p,q) = J T (p,q). Clearly T (p, q) 6∈ F SS, and thus F SS is a proper closed subideal of L(ℓp, ℓq). It is clear that J T (p,q) ⊂ J ℓ2. Conversely, Theorem 4.7 in [25] shows that every operator S : ℓp → ℓq, which factors through ℓ2, belongs to J T (p,q), thus we deduce that J T(p,q) = J ℓ2. Moreover, if S ∈ L(ℓp, ℓq) is not in F SS, it follows from Khintchine's theorem (for more detail see Theorem 3.2 in Section 3 and the remarks thereafter) that for some c > 0 there are c-complemented subspaces Fn ⊂ ℓp, which are c-isomorphic to ℓ2(n), for n ∈ N, on which S is a c-isomorphism. After perturbing S we can find a sequence (kn) ⊂ N, so that if we write ℓp as an ℓp-sum of ℓp(kn) and ℓq as the ℓq-sum of ℓq(kn), we can assume that Fn ⊂ ℓp(kn) ⊂ ℓp and S(Fn) ⊂ ℓq(kn) ⊂ ℓq. From this (see [25, Theorem 4.13]) it follows that T (p, q) factors through S. We deduce therefore that the ideal J ℓ2 ∨ F SS = J T (p,q) ∨ F SS (the closed ideal generated by the elements of F SS and J ℓ2) is the only successor of F SS. Finally we need to construct an operator U : ℓp → ℓq which is in F SS but cannot be approximated by operators which factor through ℓ2. This will show that F SS and J ℓ2 are incomparable, they both strictly contain F SS ∩J ℓ2 and are properly contained in J ℓ2 ∨F SS. To do that we write ℓp as ℓp sum of ℓp(2n), n ∈ N, and ℓq as ℓq-sum of ℓq(2n), n ∈ N. For n ∈ N ∪ {0} let Hn be the n-th Hadamard matrix. This is an 2n by 2n matrix with entries which are either 1 or −1, and can be defined by induction as follows; H0 = (1), and assuming that Hn has been defined one puts Hn+1 =(cid:18)Hn Hn Hn −Hn(cid:19). It is easy to see that Hn as operator from ℓ1(2n) → ℓ∞(2n) is of norm 1, and that 2−n/2Hn is a unitary matrix (i.e., an isometry on ℓ2(2n)). It follows therefore from the Riesz Thorin Interpolation Theorem (c.f. [4]) that Un = 2−n min(p′,q) Hn is of norm at most 1 as an operator in L(ℓp(2n), ℓq(2n)). 1 We define U : ℓp =(cid:0) ⊕∞ n=1 ℓp(2n)(cid:1)p →(cid:0) ⊕∞ n=1 ℓp(2n)(cid:1)q, 6 (xn) 7→ (Un(xn)). The fact that U can not be approximated by operators which factor though ℓ2 can be obtained from the following Corollary of Theorem 9.13 in [9] (see also [25, Theorem]). (3) Proposition 2.3. cf. [25, Corollary] Let m ∈ N, C > 1, and r > 1, and assume that V is r,ℓr′ ). Then kBkL(ℓp,ℓr) · kAkL(ℓr,ℓq) ≥ δ−1 for an invertible m by m matrix. Let δ = kV −1kL(ℓ′ keV − V kL(ℓp,ℓq) ≤(cid:0)2 max any factorization V = AB. Moreover, if eV is another m by m matrix with then it follows that for any factorization eV = AB we have kBkL(ℓp,ℓr) · kAkL(ℓr,ℓq) ≥ (2δ)−1. Indeed if p′ < q, it follows that Un = 2−n/p′Hn, and we deduce again form the Riesz Thorin Interpolation Theorem that Un is as operator between ℓp(2n) and ℓp′(2n) of norm not larger than 1, and thus U ∈ L(ℓp, ℓp′). But this implies that U (as element in L(ℓp, ℓq)) factors through I(p′, q), which is finitely strictly singular by Proposition 2.3. A similar argument shows that if p′ > q, and thus p < q′, then U factors through I(p, q′). If q 6= p′ then it is easy to see that U is finitely strictly singular. kV −1eikp(cid:1)−1, 1≤i≤m The hard case is the case q = p′ 6= 2, in which the previous factorization argument does not work. In this case it is better to see ℓp(n) as the space Lp(n), the space of all p- integrable functions on {1, 2 . . . n} with the normalized counting measure (i.e. kxkLp = 1 n/1p kxkp)). Using interpolation between Schatten p-classes one can prove the following result Theorem 2.4. [25, Theorem 6.5] Suppose that T : Lp(N) → ℓp′(N). Let E be a k-dimensional subspace of Lp(N), and C1, C2, and C3 be positive constants such that (1) kT kL(L2(N ),ℓ2(N )) ≤ 1 and kT kL(L1(N ),ℓ∞(N ))) ≤ 1; (2) E is C1-isomorphic to ℓk 2; (3) F = T (E) is C2-complemented in ℓN p′ ; and (4) TE is invertible and(cid:13)(cid:13)(TE)−1(cid:13)(cid:13) ≤ C3. Then k ≤(cid:0)C 3 G(cid:1)p′ 1 C2C 2 3 K 2 . Here KG denotes the Grothendieck constant. Now, if q = p′, then we apply for n ∈ N Theorem 2.4 to N = 2n and Tn = 1 n Hn (note Tn satisfies (1) of Theorem 2.4 ). If U where not finitely strictly singular, we could find constants C1, C2 and C3 and for any k ∈ N we would could find n = nk ∈ N large enough so that (2) and (3) of Theorem 2.4 are satisfied (using again Theorem 3.2 in Section 2) for T = Tn. But this contradicts the conclusion of Theorem 2.4. n1/p Un = 1 3. Two new closed ideals of L(ℓp, ℓq) We now state our main result, which exhibits two new closed subideals of L(ℓp, ℓq), and shows that J I(p,q) ( F SS ∩ J ℓ2 and increases the count of the known closed proper and non trivial subideals of L(ℓp, ℓq) to 7. Theorem 3.1. Assume that 1 < p < 2 < q < ∞. Then the two ideals J I(p,2) and J I(2,q) are two incomparable closed subideals of F SS ∩ J ℓ2. 7 We assume from now on that 1 < p < 2 < q < ∞. It is clear that J I(p,q) ⊂ J I(p,2) and 2 and similarly J I(p,q) ⊂ J I(2,q) ⊂ F SS ∩ J ℓ2. that by Proposition 2.2 J I(p,2) ⊂ F SS ∩ J ℓ We can therefore extend the diagram of Section 2 to the following diagram. K 3 J I(p,q) 5k k k S S S J I(p,2) −k J I(2,q) U U U 4i i i F SS ∩ J ℓ2 F SS TTTT TTTT 4j j j TTTTT −k J ℓ2 4j j j F SS ∨ J ℓ2 / L(ℓp, ℓq) This solves Question (i) in [25] and shows that J I(p,q) is different from F SS ∩ J ℓ2, and that the two (different) closed subideals J I(p,2) and J I(2,q)) lie between them. In order to show Theorem 3.1 we need to find two operators T and S in F SS ∩ J ℓ2 , so that T ∈ J I(p,2) \ J I(2,q) and S ∈ J I(2,q) \ J I(p,2). We will first need the following result. Theorem 3.2. For every 1 < r < ∞ there exists a constant K = K(r) > 0 and for all n ∈ N a number N = N(n, r) ∈ N, such that every N -- dimensional subspace F ⊂ ℓr contains an n -- dimensional subspace E which is K -- complemented in ℓr and K -- isomorphic to ℓ2(n). Remark 3.3. Theorem 3.2 follows from the finite dimensional version of Khintchin's The- orem (see [11, Theorem 6.28]). Better estimates on N(n, r) and K(r) can be obtained by applying simultaneously Dvoretzky's theorem both to a subspace F ⊂ ℓr and to its dual F ∗ (see e.g., [23]). This gives the result with N = Cnr/2 and K = C ′pmax{r, r′}, where C, C ′ > 0 are absolute constants. This theorem can also be viewed, for example, as a special case of results in [12]. Proof of Theorem 3.1. We will now construct the operators T ∈ J I(p,2) \ J I(2,q) and S ∈ J I(2,q) \ J I(p,2). Put C = max(K(p), K(q)) and for n ∈ N let kn = max(N(p, n), N(q, n)), where K(p), K(q), N(p, n) and N(q, n) are chosen as in Theorem 3.2. Using that result we can find for every n ∈ N a sequence (x(n,i))n i=1 in CBℓp(kn) so that (4) (5) i=1 is C-equivalent to the unit vector basis of ℓ2(n) and (x(n,i))n there is a projection Pn from ℓp(kn) onto span(x(n,i) : i = 1, 2, . . . n) with kPnk ≤ C. For n ∈ N we define In : span(x(n) : i = 1, 2 . . . , n) → ℓ2(n), by In(x(n,i)) = e(2,n,i), i = 1, . . . n. In is thus a C-isomorphism. Writing ℓp as ℓp-sum of ℓp(kn)and ℓ2 as ℓ2-sum of ℓ2(n), n ∈ N, we define S as follows i S :(cid:0) ⊕∞ n=1 ℓp(kn)(cid:1)p →(cid:0) ⊕∞ n=1 ℓ2(n)(cid:1)2, (xn) 7→(cid:0)In ◦ Pn(xn) : n ∈ N(cid:1). It follows that k Sk ≤ C 2. Finally we let S := I(2, q) ◦ S ∈ J I(2,q). 8 * * & . + 5 ) ) 4 * * / 4 4 The construction of T : ℓp → ℓq is similar. Using again Theorem 3.2 we find for each n ∈ N vectors (y(n,i) : i = 1, 2 . . . n) in CBℓq(kn) so that (6) (7) i=1 is C-equivalent to the unit vector basis of ℓ2(n), and (y(n,i))n there is a projection Qn from ℓq(kn) onto span(y(n,i) : i = 1, 2, . . . n) with kQnk ≤ C. Let Jn : ℓ2(n) → ℓq(kn), be the linear map which assigns to e(2,n,i) the vector y(n,i), i = 1, 2 . . . n, then Jn is a C-isomorphism onto its image, and by writing again ℓ2 as ℓ2-sum of ℓ2(n) and ℓq as ℓq-sum of ℓq(kn), n ∈ N, we define T as T :(cid:0) ⊕∞ n=1 ℓ2(n)(cid:1)2 →(cid:0) ⊕∞ n=1 ℓq(kn)(cid:1)q, (xn) 7→(cid:0)Jn(xn) : n ∈ N(cid:1). Thus T is a bounded operator with k T k ≤ C and T := T ◦ I(p, 2) ∈ J I(p,2). In order to show that S 6∈ J I(p,2) and T 6∈ J I(2,q) we will find two functionals Φ and Ψ in L∗(ℓp, ℓq) so that Φ(S) = 1 and ΦJ I(p,2) ≡ 0, and, conversely Ψ(T ) = 1 and ΨJ I(2,q) ≡ 0 . Let q′ be the conjugate of q (i.e. 1 q + 1 q′ = 1). For n ∈ N we define Φn : L(ℓp(kn), ℓq(n)) → R, with Φn(V ) = 1 n he(q′,n,i), V (x(ni))i. nXi=1 Since by choice kx(n,i)k ≤ C, for i = 1, . . . , n, it follows that k Φnk ≤ C. We extend Φn in the canonical way to a functional in L∗(ℓp, ℓq), i.e let En : ℓp(kn) → ℓp = (⊕∞ n=1ℓp(kn)) be the canonical embedding to the n-component and let Fn : ℓq = (⊕∞ j=1ℓq(j)) → ℓq(n) be the projection onto the n-th component, for n ∈ N and put Φn(U) = Φn(Fn ◦ U ◦ En) for U ∈ L(ℓp, ℓq). Then also kΦnk ≤ C and we let Φ ∈ L∗(ℓp, ℓq) be a w∗ accumulation point of the sequence (Φn) in L∗(ℓp, ℓq). Since Fn ◦ S ◦ En(x(n,i)) is the i-th unit vector in ℓq(n) it follows that Φ(S) = limn→∞ Φn(S) = 1. The definition of Ψ ∈ L∗(ℓp, ℓq) is as follows. Since (y(n,i) : i = 1, 2, . . . , n) is C-isomorphic to (e(2,n,i) : i = 1, 2 . . . , n) and its linear span is C-complemented in ℓq(kn), we can find a sequence (y∗ (n,i) : i = 1, 2 . . . , n) ⊂ ℓq′(kn), which is C-isomorphic to (e(2,n,i) : i = 1, 2 . . . , n), and satisfies hy∗ (n,i), y(n,j)i = δ(i,j) for 1 ≤ i, j ≤ n. For n ∈ N we can then write the projection Qn : ℓq(kn) → span(y(n,i) : i = 1, 2, . . . , n) (which was introduced in (7)) as Qn = y(n,i) ⊗ y∗ (n,i) : ℓq(kn) → span(y(n,i) : i = 1, 2, . . . , n), z 7→ nXi=1 Then we define for n ∈ N y(n,i)hy∗ (n,i), zi. nXi=1 Ψn : L(ℓp(n), ℓq(kn)) → R by Ψ(U) = 1 n hy∗ (n,i), U(e(p,n.i))i. nXi=1 We let Ψn be the canonical extension of Ψ to a functional in L∗(ℓp, ℓq), i.e. for U ∈ L(ℓp, ℓq) we let Ψn(U) = Ψ(F ′ n : ℓp(n) → ℓp =(cid:0) ⊕j∈N ℓp(j)(cid:1)p, is the canonical n), where E′ n ◦ U ◦ E′ 9 embedding into the n-th component, and F ′ onto the n-th component. Since ky∗ (n,i)kq′ ≤ C, for i = 1, 2 . . . n, it follows that kΨnk ≤ C and we let Ψ ∈ L∗(ℓp, ℓq) be a w∗-accumulation point of (Ψn). Since T (e(p,n,i)) = y(n,i) for i = 1, 2 . . . , n, it follows that Ψ(T ) = limn→∞hΨn, T i = 1. n : (cid:0) ⊕j∈N ℓq(kj)(cid:1)q → ℓq(kn) is the projection It is left to show that J I(p,2) ⊂ ker(Φ) and that J I(2,q) ⊂ ker(Ψ). To do so, we need a result which is of independent interest and will therefore be stated separately and more generally than needed. (cid:3) Definition 3.4. Let X be a finite or infinite dimensional Banach space with a normalized basis (ei). If X is infinite dimensional put for j ∈ N, and if j ≤ dim(X) < ∞, then put nX (j) = minn(cid:13)(cid:13)(cid:13)Xi∈I nX (j) = minn(cid:13)(cid:13)(cid:13)Xi∈I NX (j) = maxn(cid:13)(cid:13)(cid:13)Xi∈I ei(cid:13)(cid:13)(cid:13) : I ⊂ N, #I = jo, and NX(j) = maxn(cid:13)(cid:13)(cid:13)Xi∈I ei(cid:13)(cid:13)(cid:13) : I ⊂ {1, 2, . . . dim(X)}, #I = jo and ei(cid:13)(cid:13)(cid:13) : I ⊂ {1, 2, . . . dim(X)}, #I = jo. ei(cid:13)(cid:13)(cid:13) : I ⊂ N, #I = jo, Lemma 3.5. Assume that E and F are two finite dimensional spaces, both having Cu- unconditional and normalized bases (ei : i = 1, 2 . . . m) and (fj : j = 1, . . . n), respectively. Assume further that there are 1 < t < s < ∞ and positive constants c1, and c2, so that for all ℓ ∈ N (8) NE(ℓ) ≤ c1ℓ1/s and nF (ℓ) ≥ c2ℓ1/t. Then there exists a number c > 0, depending only on s, t, cu, c1, and c2, so that for every linear operator T : E → F and any ρ > 0 (cid:12)(cid:12)(cid:8)i ≤ m : T (ei)∞ = max j≤n f ∗ j (T (ei)) ≥ kT kρ(cid:9)(cid:12)(cid:12) ≤ cρ −s2 (s−1)(s−t) , j ) are the coordinate functionals to (fj). Moreover, if cu = c1 = c2 = 1, then we (9) where (f ∗ can choose c = 1. Corollary 3.6. Under the assumptions of Lemma 3.5, it follows that (10) 1 m mXi=1 kT (ei)k∞ ≤ kT k(1 + c)m−r(s,t), where r(s, t) = (s − 1)(s − t) (s − 1)(s − t) + s2 , for s > t ≥ 1. 10 Proof. First note that for any ρ > 0 Lemma 3.5 yields 1 m mXi=1 kT (ei)k∞ = 1 m mXi=1,kT (ei)k∞≤ρkT k kT (ei)k∞ + ≤ kT kρ + ckT k ρ −s2 (s−1)(s−t) . m 1 m mXi=1,kT (ei)k∞>ρkT k kT (ei)k∞ Then we let which implies that ρ = m− (s−1)(s−t) (s−1)(s−t)+s2 , 1 m mXi=1 kT (ei)k∞ ≤ kT km− (s−1)(s−t) (s−1)(s−t)+s2 + ckT km−1m (s−1)(s−t) (s−1)(s−t)+s2 s2 (s−1)(s−t) = kT km− (s−1)(s−t) (s−1)(s−t)+s2 + ckT km− (s−1)(s−t) (s−1)(s−t)+s2 = (1 + c)kT km−r(s,t). (cid:3) Proof of Lemma 3.5. For the sake of a better readability we will assume that c1 = c2 = cu = 1. The general case follows in the same way. We can also assume that kT k = 1. j=1 β(i, j)fj. Let ρ > 0 and put Let T : E → F and write yi = T (ei) as yi =Pn A = Aρ =(cid:8)i ∈ {1, 2, . . . , m} : max β(i, j) ≥ ρ(cid:9). For i ∈ A choose ji ∈ {1, 2 . . . , n} so that β(i, ji) ≥ ρ. Let A = {ji : i ∈ A} and for j ∈ A let Aj = {i ∈ A : ji = j}. In order to estimate Aj and then A we compute Aj1/s ≥ NE(Aj) (By (8)) sign(β(i, j))ej(cid:13)(cid:13)(cid:13)E ≥(cid:13)(cid:13)(cid:13)Xi∈Aj ≥(cid:13)(cid:13)(cid:13)T(cid:0)Xi∈Aj sign(β(i, j))ej(cid:17)(cid:13)(cid:13)(cid:13)F T(cid:16)Xi∈Aj ≥Df ∗ j ,Xi∈Aj =Xi∈Aj β(i, j) ≥ Ajρ sign(β(i, j))ej(cid:17)E (Since kT k = 1) which yields Aj1− 1 s ≤ ρ−1, and thus Since A =Pj∈ A Aj ≤ A · ρ− s (11) Aj ≤ ρ−1/(1− 1 s−1 , we obtain s ) = ρ− s s−1 . A ≥ Aρ s s−1 . 11 Let (rj)m j=1 be a Rademacher sequence on some probability space (Ω, Σ, P), this means that r1, r2, . . . rm are independent and {±1}-valued, with P({rj = 1}) = P({rj = −1}) = 1/2 for j = 1, 2 . . . n. We compute (Since kT k ≤ 1) (By 1-unconditionality of (fj)). Applying the multidimensional version of Jensen's inequality (c.f [8, 10.2.6, page 348]) j=1 zjfjkF and the Rn valued random vector Z = A1/s ≥ NE(A) (By (8)) ≥ E(cid:16)(cid:13)(cid:13)(cid:13)Xi∈A ≥ E(cid:16)(cid:13)(cid:13)(cid:13)Xi∈A = E(cid:16)(cid:13)(cid:13)(cid:13) nXj=1 riei(cid:13)(cid:13)(cid:13)E(cid:17) riβ(i, j)fj(cid:13)(cid:13)(cid:13)F(cid:17) nXj=1 fj(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)F(cid:17) to the convex function Rn ∋ z → kPn (cid:0)Pi∈A riβ(i, j) : j ≤ n(cid:1)) we obtain riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17) = E(cid:16)(cid:12)(cid:12)(cid:12)rij β(ij, j) + r Xi∈A\{ij } 2(cid:12)(cid:12)(cid:12)rij β(ij, j) + Xi∈A\{ij } riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A A1/s ≥(cid:13)(cid:13)(cid:13) nXj=1 ≥(cid:13)(cid:13)(cid:13)Xj∈ A A1/s ≥(cid:13)(cid:13)(cid:13)Xj∈ A E(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A = E 1 For each j ∈ A there is an ij ∈ A so that β(i, j) ≥ ρ. Let r be anther ±1 random variable with P(r = 1) = P(r = −1) = 1/2, which is independent to (rj : j = 1, . . . m) then (By 1-uncondtionality of (fj)). 1 riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17) riβ(i, j)(cid:12)(cid:12)(cid:12) + 2(cid:12)(cid:12)(cid:12)rij β(ij, j) − Xi∈A\{ij } fj(cid:13)(cid:13)(cid:13)F ≥ ρ(cid:13)(cid:13)(cid:13)Xj∈ A ≥ nF ( A), riβ(i, j)(cid:12)(cid:12)(cid:12)! ≥ E(rij β(ij, j)) ≥ ρ quad (Since a + b + a − b ≥ 2a). Using again the 1-unconditionality of (fj : j = 1, 2 . . . n) we deduce therefore that and thus by our assumption (8) and by (11) we obtain A1/s ≥ nF ( A) ≥ A1/t ≥ A1/tρ s ts−t . 12 Solving for A yields which proves our claim. A ≤ ρ− s ts−t st s−t = ρ −s2 (s−1)(s−t) , (cid:3) Continuation of Proof of Theorem 3.1. In order to show that J I(p,2) ⊂ ker(Φ) we let A ∈ L2(ℓ2, ℓq) and B ∈ L(ℓp). We need to show that Φ(A ◦ I(p, 2) ◦ B) = 0. W.lo.g. we assume that kAk, kBk ≤ 1. Consider B′ canonically embedded into ℓp = (cid:0) ⊕∞ Corollary 3.6 to B′, s = 2 and t = p, we obtain n : ℓ2(n) → ℓp(kn) with B′(e(2,n,i)) = B(x(n,i)), where we consider ℓp(kn) nk ≤ C and applying therefore kB(x(n,i))k∞ = n(e(2,n,i))k∞ ≤ 2Cn−r(2,p). 1 n nXi=1 which by the concavity of [0, ∞) ∋ ξ 7→ ξ(2−p)/2 implies that ≤ (2C)(2−p)/2n−r(2,p)(2−p)/2. 1 n kB′ j=1 ℓ(kj)(cid:1). Then kB′ nXi=1 kB(x(n,i))k∞(cid:17)(2−p)/2 nXi=1 B(x(n,i))(j)2(cid:17)1/2 B(x(n,i))(j)pB(x(n,i))(j)2−p(cid:17)1/2 Secondly we observe that for any i = 1, 2 . . . n (12) (13) 1 n n ∞ kB(x(n,i))k(2−p)/2 ≤(cid:16) 1 nXi=1 kI(p, 2)(B(x(n,i)))k2 =(cid:16) knXj=1 =(cid:16) knXj=1 ≤ kB(x(n,i))k(2−p)/2 ∞ · kB(x(n,i))kp/2 p ≤ C p/2kB(x(n,i))k(2−p)/2 ∞ . It follows therefore that (cid:12)(cid:12)Φn(A ◦ I(p, 2) ◦ B)(cid:12)(cid:12)(cid:12) = = ≤ 1 1 n(cid:12)(cid:12)(cid:12) nXi=1 n(cid:12)(cid:12)(cid:12) nXi=1 nXi=1 1 n he(q′,n,i), A ◦ I(p,2) ◦ B(x(n,i))i(cid:12)(cid:12)(cid:12) hA∗(e(q′,n,i)), I(p,2) ◦ B(x(n,i))i(cid:12)(cid:12)(cid:12) nXi=1 kx(n,i)k(2−p)/2 kA∗(e(q′,n,i))k2kI(p,2) ◦ B(x(n,i)k2 ∞ ≤ kA∗kC p/2 1 n (By (13)) ≤ C p/2(2C)(2−p)/2n−r(2,p)(2−p)/2 →n→∞ 0 (By (12)). This implies that J I(p,2) ⊂ ker(Φ). 13 In order to show that J I(2,q) ⊂ ker(Ψ), let B ∈ L(ℓp, ℓ2) and A ∈ L(ℓq) with kBk, kAk ≤ 1. We need to show that Ψ(A ◦ I(2,q) ◦ B) = 0. n : ℓ2(n) → ℓq′(kn), defined by A′ Let A′ ℓq′(kn) in the canonical way as subspace of ℓq′ = (⊕∞ of (y∗ deduce that (n,i) : i = 1, 2 . . . n) that kA′ n(e(2,n,i)) = A∗(y∗ (n,i)), i = 1, 2 . . . n (we consider j=1ℓq′(kn))q). It follows from the choice nk ≤ C and from Corollary 3.6 (with s = 2 and t = q′) we 1 n nXi=1 kA∗(y∗ (n,i))k∞ = 1 n kA′(e(2,n,i))k∞ ≤ 2Cn−r(2,q′). nXi=1 Using the concavity of the function [0, ∞) ∋ ξ → ξ(2−q′)/2 we deduce (14) 1 n nXi=1 kA∗(y∗ (n,i))k(2−q′)/2 ∞ =(cid:16) 1 n nXi=1 kA∗(y∗ (n,i))k∞(cid:17)(2−q′)/2 ≤ (2C)(2−q′)/2n−r(2,q′)(2−q′)/2. It is easy to see that I(q′,2) is the adjoint of I(2,q) and we compute for i = 1, 2 . . . n (15) kI(q′, 2) ◦ A∗(y∗(n, i))k2 =(cid:16) knXj=1(cid:0)A∗(y∗(n, i))(j)(cid:1)2(cid:17)1/2 =(cid:16) knXj=1(cid:12)(cid:12)A∗(y∗(n, i))(j)(cid:12)(cid:12)q′(cid:12)(cid:12)A∗(y∗(n, i))(j)(cid:12)(cid:12)2−q′(cid:17)1/2 ≤ kA∗(y∗(n, i))k(2−q′)/2 ≤ C q′/2kA∗(y∗(n, i))k(2−q′)/2 ∞ . ky(n,i)kq′/2 q′ ∞ Therefore it follows hψn, Ui = = ≤ 1 1 n(cid:12)(cid:12)(cid:12) nXi=1 n(cid:12)(cid:12)(cid:12) nXi=1 nXi=1 1 n hA ◦ I(2,q) ◦ B(e(p,n,i)), y∗(n, i)i(cid:12)(cid:12)(cid:12) hB(e(p,n,i)), I(q′,2) ◦ A∗(y∗(n, i))i(cid:12)(cid:12)(cid:12) nXi=1 kB(e(p,n,i))k2 · kI(q′,2) ◦ A∗(y∗ kA∗(y∗(n, i))k(2−q′)/2 ∞ (n,i))k2 ≤ kBkC q′/2 1 n (By (15)) ≤ C q′/2(C + 1)(2−q′)/2n−r(2,q′)(2−q′)/2 →n→∞ 0 (By (14)). which implies our claim, and finishes the proof or Theorem 3.1. (cid:3) 14 References [1] G. Androulakis and Th. Schlumprecht, Strictly singular, non-compact operators exist on the space of Gowers and Maurey, J. London Math. Soc. 64 (2001), no. 3, 65 -- 674. [2] G. Androulakis and Th. Schlumprecht, The Banach space S is complementably minimal and subsequen- tially prime, Studia Math. 156 (2003), no. 3, 22 -- 242. [3] S.A. Argyros and R. Haydon, A hereditarily indecomposable L∞-space that solves the scalar-plus- compact, preprint, arXiv 0903.3921. [4] J. Bergh and J. Lofstrom, Interpolation spaces. An introduction. Grundlehren der Mathematischen Wissenschaften, No. 223. Springer-Verlag, Berlin-New York, 1976. [5] A. Bird, G. J. O. Jameson and N.-J. Laustsen, The GiesyJames theorem for general index p, with an application to closed ideals of operators on the p th quasi-reflexive James space, in preparation. [6] J.W. Calkin, Two-sided ideals and congruences in the ring of bounded operators in Hilbert space, Ann. of Math. 42 (1941) no.2, 839 -- 873 . [7] M. Daws, Closed ideals in the Banach algebra of operators on classical non-separable spaces, Math.Proc.Camb.Phil. Soc. 140 (2006) 317 -- 332. [8] R.M. Dudley, Real analysis and probability. Revised reprint of the 1989 original. Cambridge Studies in Advanced Mathematics, 74. Cambridge University Press, Cambridge, 2002. x+555 pp. ISBN: 0-521- 00754-2, [9] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. [10] I. S. Edelstein and B. S. Mityagin, Homotopy type of linear groups of two classes of Banach spaces, Functional Anal. Appl. 4 (1970), 221 -- 231. [11] M. Fabian, P. Habala, P. Hajek, V. Montesinos Santaluc´ıa, J. Pelant, and V. Zizler, Functional analysis and infinite-dimensional geometry. CMS Books in Mathematics/Ouvrages de Math´ematiques de la SMC, 8. Springer-Verlag, New York (2001) x+451 pp. [12] T. Figiel and N. Tomczak-Jaegermann. Projections onto Hilbertian subspaces of Banach spaces, Israel J. Math., 33 (1979) no. 2, 155 -- 171. [13] I.C. Gohberg, A.S. Markus and I.A. Feldman, Normally solvable operators and ideals associated with them, (Russian) Bul. Akad. Stiince RSS Moldoven, 76 (10) (1960) no.10 51 -- 70. English translation: American. Math. Soc. Translat. 61 (1967) 63 -- 84. [14] B. Gramsch, Eine Idealstruktur Banachscher Operatoralgebren, J. Reine Angew. Math 225 (1967) 97 -- 115. [15] N. J. Laustsen, Maximal ideals in the algebra of operators on certain Banach spaces, Proc. Edin. Math. Soc. 45 (2002), 523 -- 546. [16] N.J. Laustsen, R.J. Loy, and C.J. Read, The lattice of closed ideals in the Banach algebra of operators on certain Banach spaces, J. Funct. Anal. 214 (2004) no.1, 106 -- 131. [17] N.J. Laustsen, E. Odell, Th.Schlumprecht, and A. Zsak, Dichotomy theorems for random matrices and closed ideals of operators on(cid:0)L∞ 1(cid:1)c0 n=1 ℓn , preprint. [18] N.J. Laustsen, Th. Schlumprecht, and A. Zsak, The lattice of closed ideals in the Banach algebra of operators on a certain dual Banach space, J. of Operator Theory, 56(2006) no 2, 391-402. [19] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces. II. Springer-Verlag, Berlin, 1979. [20] R. J. Loy and G. A. Willis, Continuity of derivations on B(E) for certain Banach spaces E, J. London Math. Soc. 40 (1989), 327346. [21] E. Luft, The two-sided closed ideals of the algebra of bounded linear operators of a Hilbertspace, Czechoslovak Math. J. 18 (1968) 595 -- 605. [22] V.D. Milman, Operators of class C0 and C ∗ 0 . (Russian), Teor. Funkciı Funkcional. Anal. i Prilozen., 10 (1970) 15 -- 26. 15 [23] V.D. Milman and G. Schechtman. Asymptotic Theory of Finite Dimensional Normed Spaces, LNM 1200, Springer-Verlag, New York, 1986. [24] A. Pietsch. Operator ideals, volume 16 of Mathematische Monographien [Mathematical Monographs]. VEB Deutscher Verlag der Wissenschaften, Berlin, 1978. [25] B. Sari, Th. Schlumprecht, N. Tomczak-Jaegermann, and V. G. Troitsky, On norm closed ideals in L(ℓp, ℓq). Studia Math. 179 (2007), no. 3, 239 -- 262. [26] Th. Schlumprecht, An arbitrarily distortable Banach space, Israel J. Math, 76 (1991), no. 1-2, 8195. Department of Mathematics, Texas A&M University, College Station, Texas 77843, USA E-mail address: [email protected] 16 1 1 0 2 y a M 4 2 ] . A F h t a m [ 2 v 0 1 6 3 . 5 0 1 1 : v i X r a ON THE CLOSED SUBIDEALS OF L(ℓp ⊕ ℓq) TH. SCHLUMPRECHT Abstract. In this paper we first review the known results about the closed subideals of the space of bounded operator on ℓp ⊕ ℓq, 1 < p < q < ∞, and then construct several new ones. 1. Introduction For very few Banach spaces X all the closed subideals of L(X), the algebra of all bounded and linear operators on X, are determined. In 1941 Calkin [6] showed that the only proper, non-trivial and closed ideal of L(ℓ2) is the ideal of compact operators. The same was shown to be true for ℓp (1 ≤ p < ∞) and c0 in [13]. Until very recently it was open if there are any other infinite dimensional Banach spaces X, for which the compact operators are the only proper, non-trivial and closed subideal of L(X). We call such spaces simple. Then Argyros and Haydon [3] established the existence of Banach spaces with a basis on which all operators are a compact perturbation of a scalar multiple of the identity. It follows immediately that such spaces are simple. But it is not known whether or not there are any other simple spaces admitting an unconditional basis (and thus having a rich structure of operators on them). The structure of the closed ideals of operators on non separable Hilbert spaces was inde- pendently obtained by Gramsch [14] and Luft [21]. Recently Daws [7] extended their results to non separable ℓp-spaces, 1 ≤ p < ∞, and non separable c0-spaces. . for X = (cid:0)L∞ n=1 ℓ2(n)(cid:1)c0 n=1 ℓ2(n)(cid:1)ℓ1 and in [18] for X = (cid:0)L∞ Beyond these spaces the complete structure of closed ideals in L(X) was described in [16] In both cases, there are exactly two nested proper non-zero closed ideals, namely the compacts and the closure of all operators factoring through c0, or ℓ1, respectively. Apart from those mentioned above, there are no other separable Banach spaces X for which the structure of the closed ideals in L(X) is completely known. It is still open whether or not the closed subideals of the operators on the spaces (⊕∞ n=1ℓ∞(n))ℓ1 admit the same sublattice structure (for partial results see [17]). An interesting space for studying the closed subideals of its bounded linear operators is the space X introduced in [26]. This space is complementably minimal [1], which means that every infinite dimensional closed subspace of X contains a further subspace which is complemented in X and isomorphic to X. This implies that the strictly singular operators (see the definition at the end of this section) is the only maximal proper closed subideal n=1ℓ1(n))c0 and (⊕∞ 2000 Mathematics Subject Classification. Primary: 47L20. Secondary: 47B10, 47B37. Key words and phrases. Operator ideal, ℓp-space. Research partially supported by NSF grant DMS 0856148. 1 of L(X). As shown in [2], X admits strictly singular but not compact operators, and it is conjectured that L(X) contains infinitely many closed subideals, all of which have to lie between the ideal of compact operators, and the ideal of strictly singular operators. A space whose closed ideals of operators attracted the attention of several researches is the pth quasi reflexive James Jp, with 1 < p < ∞. Edelstein and Mityagin [10] observed that the ideal of weakly compact operators on Jp is a maximal proper subideal of L(Jp) and Laustsen proved in [15] that it is the only one. In [20], for p = 2, and in [15], for general p ∈ (1, ∞), it was shown that the closure of the operators on Jp factoring through ℓ2 contains strictly the ideal of compact operators and is strictly contained in the ideal of weakly compact operators. Very recently Bird, Jameson and Laustsen [5] found a new closed sub ideal of L(Jp) and proved that the closure of the ideal of operators factoring through the ℓp-sum of ℓ∞(n), n ∈ N, is strictly larger than the closure of the ideal of operators factoring through ℓp and strictly smaller then the ideal of weakly compact operators. Although studied in several papers (cf.[22],[24] and [25]) the structure of the closed subide- als of L(ℓp ⊕ ℓq), 1 < p < q < ∞ remains a mystery. It is not even known whether or not L(ℓp ⊕ℓq), contains infinitely many subideals. There were several results proved in the 1970's concerning various special ideals or special cases of p and q. We refer the reader to the book by Pietsch [24, Chapter 5] for details. In particular, [24, Theorem 5.3.2] asserts that L(ℓp ⊕ ℓq), with 1 ≤ p < q, has exactly two proper maximal ideals (namely, the ideal of operators which factor through ℓp and the ideals of operators which factor through ℓq), and establishes a one-to-one correspondence between the non-maximal proper subideals of L(ℓp ⊕ ℓq) and the closed ideals in L(ℓp, ℓq). By proving that the formal identity I(p, q) : ℓp → ℓq is finitely strictly singular (see the definition at the end of this section) and establishing the existence of an operator T : ℓp → ℓq which is not finitely strictly singular Milman [22] concluded that L(ℓp, ℓq) contains at least two non trivial, proper and closed subideals. In [25] the study of the structure of the closed subideals of L(ℓp, ℓq) was continued, and, among other results, it was discovered that the lattice of subideals of L(ℓp, ℓq) is not linearly ordered, and contains at least 4 nontrivial, proper and closed subideals if 1 < p < 2 < q < ∞. In this paper we increase this number to 7. In Section 2 we will recall the known results on the closed subideals of L(ℓp ⊕ ℓq) and L(ℓp, ℓq), and sketch the proof of several of them. In Section 3 we will formulate and prove our main result (see Theorem 3.1). Let us first recall some necessary notation. If X and Y are Banach spaces, L(X, Y ) denotes the space of bounded linear operators T : X → Y , and if X = Y we write L(X) instead of L(X, X). A linear subspace J ⊂ L(X, Y ), is called a subideal of L(X, Y ), if for all A ∈ L(Y ), B ∈ L(X), and T ∈ J also A◦ T ◦ B ∈ J . A closed subideal of L(X, Y ) is a subideal which is closed in the operator norm. We say that a subideal J ⊂ L(X, Y ) is non trivial if it is not the zero ideal {0} and proper if it is not all of L(X, Y ). The following is a list of some important closed subideals of L(X, Y ). 2 F D(X, Y ) is the closure of the ideal of operators with finite dimensional rank. Note that any nontrivial closed subideal J in L(X, Y ) contains all of F D(X, Y ). This follows from the fact that J is closed under taking sums, under multiplication by elements of L(X) from the right, under multiplication from the left by elements of L(Y ), and that it must contain a non zero operator (and thus a rank 1 operator). Thus, for all infinite dimensional Banach spaces X and Y the ideal F D(X, Y ) is the minimal nontrivial closed subideal of L(X, Y ). K(X, Y ) denotes the ideal of compact operators . All the spaces we consider are spaces with a basis. Thus, these spaces have the approximation property, which means that F D(X, Y ) = K(X, Y ). StSi(X, Y ) is the closed ideal of operators T : X → Y which are strictly singular, i.e. on no infinite dimensional subspace Z of X is the restriction of T onto Z an isomorphism. F SS is the closed ideal of finitely strictly singular operators. A linear bounded operator T : X → Y , is called finitely strictly singular if for all ε > 0 there is an n = nε ∈ N so that for any n-dimensional subspace E of X, there is an x ∈ E, with kxk = 1, so that kT (x)k ≤ ε. If W and Z are Banach spaces and S : W → Z a bounded linear operator, we denote by J S(X, Y ) the closure of the ideal generated by all operators T ∈ L(X, Y ), which factor through S, thus T = A ◦ S ◦ B, with A ∈ L(Z, Y ) and B ∈ L(X, W ). In general the set {A ◦ S ◦ B, A ∈ L(Z, Y ) and B ∈ L(X, W )} is not closed under addition and therefore not an ideal. But if the operator S ⊕ S : W ⊕ W → Z ⊕ Z, (w1, w2) 7→ (S(w1), S(w2)), factors through S, then {A ◦ S ◦ B, A ∈ L(Z, Y ) and B ∈ L(X, W )} is an ideal and we conclude in that case that (1) J S(X, Y ) = {A◦S ◦B : A ∈ L(Z, Y ) and B ∈ L(X, W )}. Let I(p, q) : ℓp → ℓq be the formal inclusion (using that ℓp is a subset of ℓq), for 1 ≤ p < q ≤ ∞. It is easily seen that I(p, q) ⊕ I(p, q) factors through I(p, q) and we conclude that J I(p,q)(X, Y ) = {A◦I(p, q)◦B : A ∈ L(ℓq, Y ) and B ∈ L(X, ℓp)}. If IZ is the identity on some Banach space Z we write J Z instead of J IZ , and we note that if Z is isomorphic to Z ⊕ Z it follows that (2) J Z(X, Y ) = {A◦S ◦B : A ∈ L(Z, Y ) and B ∈ L(X, Z)}. If X = Y we will write K(X), F SS(X) etc. instead of K(X, X), F SS(X, X) etc. For 1 ≤ p < ∞, we denote the unit vector basis of ℓp = ℓp(N) by (e(p,j) : j ∈ N) (if p = ∞ we consider c0 instead of ℓ∞). The conjugate of p is denoted by p′, i.e. 1 p′ = 1. For n ∈ N we denote the n-dimensional ℓp space by ℓp(n) and its unit vector basis by (e(p,n,j) : j = 1, 2, . . . , n). The usual norm on ℓp or ℓp(n), n ∈ N is denoted by k · kp. If Xn is a Banach space for n ∈ N, the ℓp-sum of Xn, n ∈ N, is the space of all sequences (xn : n ∈ N), with xn ∈ Xn, for n ∈ N, and p + 1 k(xn)n∈Nkp =(cid:16)Xn∈N kxnkp(cid:17)1/p 3 < ∞, if p < ∞, and We denote the ℓp-sum of (Xn) by (⊕∞ n=1Xn)∞ the c0-sum, the space of all sequences (xn), with xn ∈ Xn, for n ∈ N, for which limn→∞ kxnk = 0. The sphere and the unit ball of a Banach space are denoted by SX and BX , respectively. For simplicity all our Banach spaces are defined over the real field R. It is easy to see how our results can be extended to Banach spaces over the complex field C. n=1Xn)p. If p = ∞ we denote by (⊕∞ 2. Review of the known results on the closed subideals of L(ℓp ⊕ ℓq) and L(ℓp, ℓq) We will now review the known results on the lattice structure of subideals of L(ℓp ⊕ ℓq). We will assume from now on that 1 < p < q < ∞ and later that 1 < p < 2 < q < ∞. Every operator T = ℓp ⊕ ℓq → ℓp ⊕ ℓq, consists of four operators T(1,1) ∈ L(ℓp), T(1,2) ∈ L(ℓq, ℓp) and T(2,1) ∈ L(ℓp, ℓq), and T(2,2) ∈ L(ℓp, ℓp), and acts as a 2 by 2 matrix on the elements of ℓp ⊕ ℓq T =(cid:18)T(1,1) T(1,2) T(2,1) T(2,2)(cid:19) : ℓp⊕ℓq → ℓp⊕ℓq, (x, y) 7→(cid:0)T(1,1)(x)+T(1,2)(y), T(2,1)(x)+T(2,2)(y)(cid:1). By the above cited result from [13], the operators T(1,1) and T(2,2) are either compact or the identity on ℓp, respectively ℓq, factors through them. By Pitt's Theorem (c.f. [11, Proposition 6.25]), T(1,2) is compact, and since every infinite dimensional subspace of ℓp contains a subspace isomorphic to ℓp, and since ℓp and ℓq are incomparable, we conclude that T(2,1) must be strictly singular. So, if J is a closed subideal of L(ℓp ⊕ ℓq) which contains an operator T for which T(1,1) and T(2,2) are not compact, we conclude that the identity on ℓp ⊕ ℓq factors through T and thus J = L(ℓp ⊕ ℓq). If J contains an operator for which T(1,1) is not compact, but for all elements U ∈ J , U(2,2) is compact, then the identity on ℓp factors through T , but not the identity on ℓq, and we therefore deduce that J must be the closure of the operators factoring through ℓp, which must therefore be a maximal proper subideal of L(ℓp ⊕ ℓq) (for more details see [24, Theorem 5.3.2]). Similarly we conclude that the closure of all operators factoring through ℓq is a maximal proper subideal of L(ℓp ⊕ ℓq). For all other closed proper subideals J ⊂ L(ℓp ⊕ ℓq), and all T ∈ J it therefore follows that T(1,1), T(1,2) and T(2,2) are compact, and can therefore be approximated by finite rank operators which factor through ℓp as well as ℓq. Of course T(2,1) also factors through ℓp as well as ℓq, and we deduce that all other closed ideals are subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq (ℓp ⊕ ℓq), and thus not maximal proper closed ideals. Assume now that J ⊂ J ℓp(ℓp ⊕ ℓq) ∩ J ℓp(ℓp ⊕ ℓq) is a closed ideal in L(ℓp ⊕ ℓq) An easy computation yields that J := {T(2,1) : T ∈ J } is a closed subideal of L(ℓp, ℓq), and that for two different ideals J1, J2 ⊂ J ℓp(ℓp ⊕ ℓq) ∩ J ℓp(ℓp ⊕ ℓq) the ideals J1 and J2 are different. Conversely if J is a closed subideal of L(ℓp, ℓq) then J ′ =(cid:26)(cid:18)T(1,1) T(1,2) T(2,1) T(2,2)(cid:19) : T(2,10 ∈ J and T(1,1) ∈ K(ℓp), T(1,2) ∈ K(ℓq, ℓp), and T(2,2) ∈ K(ℓq)(cid:27) 4 1 and J ′ is a closed subideal of L(ℓp ⊕ ℓq) and for two different closed subideals J1, J2 ⊂ L(ℓp, ℓq), J ′ 2 are different. Thus there is a bijection between the set of all closed subideals of L(ℓp, ℓq) and the non maximal closed subideals of L(ℓp ⊕ ℓq), which preserves the lattice structure with respect to inclusions. Let us summarize the observations we just made in the following proposition. Proposition 2.1. For 1 < p < q < ∞, the space L(ℓp ⊕ ℓq) has exactly two maximal proper closed subideals, namely J ℓp(ℓp ⊕ ℓq) and J ℓq (ℓp ⊕ ℓq). All other closed subideals of L(ℓp ⊕ ℓq), are subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq(ℓp ⊕ ℓq), and there is a bijection between the closed subideals of J ℓp(ℓp ⊕ ℓq) ∩ J ℓq(ℓp ⊕ ℓq) and closed subideals of L(ℓp, ℓq) which preserves the lattice structure. We are therefore interested in the closed subideals of L(ℓp, ℓq). Instead of writing K(ℓp, ℓq), F SS(ℓp, ℓq), or J S(X, Y ) etc. we will from now on simply write K, F SS or J S etc. The following diagram summarizes the results established in [22] and [25], under the assumption that 1 < p < 2 < q < ∞. K 3 J I(p,q) / F SS ∩ J ℓ2 F SS UUUU UUUU 4i i i TTTTT −k J ℓ2 4j j j F SS ∨ J ℓ2 / L(ℓp, ℓq) Here arrows stand for inclusions. A solid arrow (⇒ or →) between two ideals means that there are no other ideals sitting properly between the two, while a double arrow coming out of an ideal indicates the only immediate successor. A hyphenated arrow (−−>) indicates a proper inclusion, while a dotted one indicates that we do not know whether or not the inclusion is proper. In particular, the closed ideals in L(ℓp, ℓq) are not totally ordered. Let us explain the diagram "from the left to the right" (for a more detailed explanation we refer the reader to [25]): If T : ℓp → ℓq is not compact, then there is a normalized block sequence (xn) in ℓp whose image (yn) = (T (xn) is equivalent to (e(q,j) : j ∈ N) (the unit vector basis in ℓq) and so that span(yn : n ∈ N) is complemented in ℓp. It follows that I(p, q) factors through T , and that therefore J I(p,q) is the only successor of K. It is clear that J I(p,q) ⊂ J ℓ2 (recall that we assume that p < 2 < q). The fact that J I(p,q) ⊂ F SS follows from the following result in [22] (see also [25, Proposition 3.3]). Proposition 2.2. For any choices of 1 ≤ p < q ≤ ∞ is the formal identity I(p, q) is a finitely strictly singular operator. The way to verify Proposition 2.2 is to show first (see [22] or [25, Lemma 3.4]) by induction on n ∈ N, that in every n-dimensional subspace E of c0 there is x ∈ E which attains its sup- norm on at least n coordinates. In order to see then, that I(p, q) is finitely strictly singular, let ε > 0 and pick n ∈ N with n−(q−p)/q < ε. If E is any subspace of ℓp of dimension n we can find x ∈ E, kxkp = 1, so that kxk∞ ≤ n−1/p (since the maximum is attained on at 5 & . + / 4 * * / 4 p ≤ np−q and thus least n coordinates), and thus kxkq kxkq ≤ n−(q−p)/q ≤ ε. We therefore established that J I(p,q) ⊂ F SS ∩ J ℓ2. In Section 2 we will show that this inclusion is strict. More precisely, we will show that the ideals J I(p,2) and J I(2,q) are two distinct closed ideals which lie between J I(p,q) and F SS ∩ J ℓ2. i=1 x(i)q−px(i)p ≤ kxkq−p ∞ kxkp q =P∞ In order to show that F SS ∩ J ℓ2 is not all of L(ℓp, ℓq) Milman [22] used the fact that ℓp (and ℓq)is isomorphic the ℓp-sum (respectively the ℓq sum) of ℓ2(n), n ∈ N (see [19, page 73]). Letting U : ℓp → (⊕n∈Nℓ2(n))p and V : ℓq → (⊕n∈Nℓ2(n))q be isomorphisms and letting I ′(p, q) be the formal identity I ′(p, q) : (⊕n∈Nℓ2(n))p → (⊕n∈Nℓ2(n))q, (xn) 7→ (xn), we define T (p, q) = V ◦ I ′(p, q) ◦ U. T (p, q) depends on the choice of the isomorphisms U and V , nevertheless it is easy to see that for any other isomorphisms U : ℓp → (⊕n∈Nℓ2(n))p and V : ℓq → (⊕n∈Nℓ2(n))q, the operator T (p, q) = V ◦ I ′(p, q) ◦ U , factors through T (p, q) and vice versa, and thus J T (p,q) = J T (p,q). Clearly T (p, q) 6∈ F SS, and thus F SS is a proper closed subideal of L(ℓp, ℓq). It is clear that J T (p,q) ⊂ J ℓ2. Conversely, Theorem 4.7 in [25] shows that every operator S : ℓp → ℓq, which factors through ℓ2, belongs to J T (p,q), thus we deduce that J T(p,q) = J ℓ2. Moreover, if S ∈ L(ℓp, ℓq) is not in F SS, it follows from Khintchine's theorem (for more detail see Theorem 3.2 in Section 3 and the remarks thereafter) that for some c > 0 there are c-complemented subspaces Fn ⊂ ℓp, which are c-isomorphic to ℓ2(n), for n ∈ N, on which S is a c-isomorphism. After perturbing S we can find a sequence (kn) ⊂ N, so that if we write ℓp as an ℓp-sum of ℓp(kn) and ℓq as the ℓq-sum of ℓq(kn), we can assume that Fn ⊂ ℓp(kn) ⊂ ℓp and S(Fn) ⊂ ℓq(kn) ⊂ ℓq. From this (see [25, Theorem 4.13]) it follows that T (p, q) factors through S. We deduce therefore that the ideal J ℓ2 ∨ F SS = J T (p,q) ∨ F SS (the closed ideal generated by the elements of F SS and J ℓ2) is the only successor of F SS. Finally we need to construct an operator U : ℓp → ℓq which is in F SS but cannot be approximated by operators which factor through ℓ2. This will show that F SS and J ℓ2 are incomparable, they both strictly contain F SS ∩J ℓ2 and are properly contained in J ℓ2 ∨F SS. To do that we write ℓp as ℓp sum of ℓp(2n), n ∈ N, and ℓq as ℓq-sum of ℓq(2n), n ∈ N. For n ∈ N ∪ {0} let Hn be the n-th Hadamard matrix. This is an 2n by 2n matrix with entries which are either 1 or −1, and can be defined by induction as follows; H0 = (1), and assuming that Hn has been defined one puts Hn+1 =(cid:18)Hn Hn Hn −Hn(cid:19). It is easy to see that Hn as operator from ℓ1(2n) → ℓ∞(2n) is of norm 1, and that 2−n/2Hn is a unitary matrix (i.e., an isometry on ℓ2(2n)). It follows therefore from the Riesz Thorin Interpolation Theorem (c.f. [4]) that Un = 2−n min(p′,q) Hn is of norm at most 1 as an operator in L(ℓp(2n), ℓq(2n)). 1 We define U : ℓp =(cid:0) ⊕∞ n=1 ℓp(2n)(cid:1)p →(cid:0) ⊕∞ n=1 ℓp(2n)(cid:1)q, 6 (xn) 7→ (Un(xn)). The fact that U can not be approximated by operators which factor though ℓ2 can be obtained from the following Corollary of Theorem 9.13 in [9] (see also [25, Theorem]). (3) Proposition 2.3. cf. [25, Corollary] Let m ∈ N, C > 1, and r > 1, and assume that V is r,ℓr′ ). Then kBkL(ℓp,ℓr) · kAkL(ℓr,ℓq) ≥ δ−1 for an invertible m by m matrix. Let δ = kV −1kL(ℓ′ keV − V kL(ℓp,ℓq) ≤(cid:0)2 max any factorization V = AB. Moreover, if eV is another m by m matrix with then it follows that for any factorization eV = AB we have kBkL(ℓp,ℓr) · kAkL(ℓr,ℓq) ≥ (2δ)−1. Indeed if p′ < q, it follows that Un = 2−n/p′Hn, and we deduce again form the Riesz Thorin Interpolation Theorem that Un is as operator between ℓp(2n) and ℓp′(2n) of norm not larger than 1, and thus U ∈ L(ℓp, ℓp′). But this implies that U (as element in L(ℓp, ℓq)) factors through I(p′, q), which is finitely strictly singular by Proposition 2.3. A similar argument shows that if p′ > q, and thus p < q′, then U factors through I(p, q′). If q 6= p′ then it is easy to see that U is finitely strictly singular. kV −1eikp(cid:1)−1, 1≤i≤m The hard case is the case q = p′ 6= 2, in which the previous factorization argument does not work. In this case it is better to see ℓp(n) as the space Lp(n), the space of all p- integrable functions on {1, 2 . . . n} with the normalized counting measure (i.e. kxkLp = 1 n/1p kxkp)). Using interpolation between Schatten p-classes one can prove the following result Theorem 2.4. [25, Theorem 6.5] Suppose that T : Lp(N) → ℓp′(N). Let E be a k-dimensional subspace of Lp(N), and C1, C2, and C3 be positive constants such that (1) kT kL(L2(N ),ℓ2(N )) ≤ 1 and kT kL(L1(N ),ℓ∞(N ))) ≤ 1; (2) E is C1-isomorphic to ℓk 2; (3) F = T (E) is C2-complemented in ℓN p′ ; and (4) TE is invertible and(cid:13)(cid:13)(TE)−1(cid:13)(cid:13) ≤ C3. Then k ≤(cid:0)C 3 G(cid:1)p′ 1 C2C 2 3 K 2 . Here KG denotes the Grothendieck constant. Now, if q = p′, then we apply for n ∈ N Theorem 2.4 to N = 2n and Tn = 1 n Hn (note Tn satisfies (1) of Theorem 2.4 ). If U where not finitely strictly singular, we could find constants C1, C2 and C3 and for any k ∈ N we would could find n = nk ∈ N large enough so that (2) and (3) of Theorem 2.4 are satisfied (using again Theorem 3.2 in Section 2) for T = Tn. But this contradicts the conclusion of Theorem 2.4. n1/p Un = 1 3. Two new closed ideals of L(ℓp, ℓq) We now state our main result, which exhibits two new closed subideals of L(ℓp, ℓq), and shows that J I(p,q) ( F SS ∩ J ℓ2 and increases the count of the known closed proper and non trivial subideals of L(ℓp, ℓq) to 7. Theorem 3.1. Assume that 1 < p < 2 < q < ∞. Then the two ideals J I(p,2) and J I(2,q) are two incomparable closed subideals of F SS ∩ J ℓ2. 7 We assume from now on that 1 < p < 2 < q < ∞. It is clear that J I(p,q) ⊂ J I(p,2) and 2 and similarly J I(p,q) ⊂ J I(2,q) ⊂ F SS ∩ J ℓ2. that by Proposition 2.2 J I(p,2) ⊂ F SS ∩ J ℓ We can therefore extend the diagram of Section 2 to the following diagram. K 3 J I(p,q) 5k k k S S S J I(p,2) −k J I(2,q) U U U 4i i i F SS ∩ J ℓ2 F SS TTTT TTTT 4j j j TTTTT −k J ℓ2 4j j j F SS ∨ J ℓ2 / L(ℓp, ℓq) This solves Question (i) in [25] and shows that J I(p,q) is different from F SS ∩ J ℓ2, and that the two (different) closed subideals J I(p,2) and J I(2,q)) lie between them. In order to show Theorem 3.1 we need to find two operators T and S in F SS ∩ J ℓ2 , so that T ∈ J I(p,2) \ J I(2,q) and S ∈ J I(2,q) \ J I(p,2). We will first need the following result. Theorem 3.2. For every 1 < r < ∞ there exists a constant K = K(r) > 0 and for all n ∈ N a number N = N(n, r) ∈ N, such that every N -- dimensional subspace F ⊂ ℓr contains an n -- dimensional subspace E which is K -- complemented in ℓr and K -- isomorphic to ℓ2(n). Remark 3.3. Theorem 3.2 follows from the finite dimensional version of Khintchin's The- orem (see [11, Theorem 6.28]). Better estimates on N(n, r) and K(r) can be obtained by applying simultaneously Dvoretzky's theorem both to a subspace F ⊂ ℓr and to its dual F ∗ (see e.g., [23]). This gives the result with N = Cnr/2 and K = C ′pmax{r, r′}, where C, C ′ > 0 are absolute constants. This theorem can also be viewed, for example, as a special case of results in [12]. Proof of Theorem 3.1. We will now construct the operators T ∈ J I(p,2) \ J I(2,q) and S ∈ J I(2,q) \ J I(p,2). Put C = max(K(p), K(q)) and for n ∈ N let kn = max(N(p, n), N(q, n)), where K(p), K(q), N(p, n) and N(q, n) are chosen as in Theorem 3.2. Using that result we can find for every n ∈ N a sequence (x(n,i))n i=1 in CBℓp(kn) so that (4) (5) i=1 is C-equivalent to the unit vector basis of ℓ2(n) and (x(n,i))n there is a projection Pn from ℓp(kn) onto span(x(n,i) : i = 1, 2, . . . n) with kPnk ≤ C. For n ∈ N we define In : span(x(n) : i = 1, 2 . . . , n) → ℓ2(n), by In(x(n,i)) = e(2,n,i), i = 1, . . . n. In is thus a C-isomorphism. Writing ℓp as ℓp-sum of ℓp(kn)and ℓ2 as ℓ2-sum of ℓ2(n), n ∈ N, we define S as follows i S :(cid:0) ⊕∞ n=1 ℓp(kn)(cid:1)p →(cid:0) ⊕∞ n=1 ℓ2(n)(cid:1)2, (xn) 7→(cid:0)In ◦ Pn(xn) : n ∈ N(cid:1). It follows that k Sk ≤ C 2. Finally we let S := I(2, q) ◦ S ∈ J I(2,q). 8 * * & . + 5 ) ) 4 * * / 4 4 The construction of T : ℓp → ℓq is similar. Using again Theorem 3.2 we find for each n ∈ N vectors (y(n,i) : i = 1, 2 . . . n) in CBℓq(kn) so that (6) (7) i=1 is C-equivalent to the unit vector basis of ℓ2(n), and (y(n,i))n there is a projection Qn from ℓq(kn) onto span(y(n,i) : i = 1, 2, . . . n) with kQnk ≤ C. Let Jn : ℓ2(n) → ℓq(kn), be the linear map which assigns to e(2,n,i) the vector y(n,i), i = 1, 2 . . . n, then Jn is a C-isomorphism onto its image, and by writing again ℓ2 as ℓ2-sum of ℓ2(n) and ℓq as ℓq-sum of ℓq(kn), n ∈ N, we define T as T :(cid:0) ⊕∞ n=1 ℓ2(n)(cid:1)2 →(cid:0) ⊕∞ n=1 ℓq(kn)(cid:1)q, (xn) 7→(cid:0)Jn(xn) : n ∈ N(cid:1). Thus T is a bounded operator with k T k ≤ C and T := T ◦ I(p, 2) ∈ J I(p,2). In order to show that S 6∈ J I(p,2) and T 6∈ J I(2,q) we will find two functionals Φ and Ψ in L∗(ℓp, ℓq) so that Φ(S) = 1 and ΦJ I(p,2) ≡ 0, and, conversely Ψ(T ) = 1 and ΨJ I(2,q) ≡ 0 . Let q′ be the conjugate of q (i.e. 1 q + 1 q′ = 1). For n ∈ N we define Φn : L(ℓp(kn), ℓq(n)) → R, with Φn(V ) = 1 n he(q′,n,i), V (x(ni))i. nXi=1 Since by choice kx(n,i)k ≤ C, for i = 1, . . . , n, it follows that k Φnk ≤ C. We extend Φn in the canonical way to a functional in L∗(ℓp, ℓq), i.e let En : ℓp(kn) → ℓp = (⊕∞ n=1ℓp(kn)) be the canonical embedding to the n-component and let Fn : ℓq = (⊕∞ j=1ℓq(j)) → ℓq(n) be the projection onto the n-th component, for n ∈ N and put Φn(U) = Φn(Fn ◦ U ◦ En) for U ∈ L(ℓp, ℓq). Then also kΦnk ≤ C and we let Φ ∈ L∗(ℓp, ℓq) be a w∗ accumulation point of the sequence (Φn) in L∗(ℓp, ℓq). Since Fn ◦ S ◦ En(x(n,i)) is the i-th unit vector in ℓq(n) it follows that Φ(S) = limn→∞ Φn(S) = 1. The definition of Ψ ∈ L∗(ℓp, ℓq) is as follows. Since (y(n,i) : i = 1, 2, . . . , n) is C-isomorphic to (e(2,n,i) : i = 1, 2 . . . , n) and its linear span is C-complemented in ℓq(kn), we can find a sequence (y∗ (n,i) : i = 1, 2 . . . , n) ⊂ ℓq′(kn), which is C-isomorphic to (e(2,n,i) : i = 1, 2 . . . , n), and satisfies hy∗ (n,i), y(n,j)i = δ(i,j) for 1 ≤ i, j ≤ n. For n ∈ N we can then write the projection Qn : ℓq(kn) → span(y(n,i) : i = 1, 2, . . . , n) (which was introduced in (7)) as Qn = y(n,i) ⊗ y∗ (n,i) : ℓq(kn) → span(y(n,i) : i = 1, 2, . . . , n), z 7→ nXi=1 Then we define for n ∈ N y(n,i)hy∗ (n,i), zi. nXi=1 Ψn : L(ℓp(n), ℓq(kn)) → R by Ψ(U) = 1 n hy∗ (n,i), U(e(p,n.i))i. nXi=1 We let Ψn be the canonical extension of Ψ to a functional in L∗(ℓp, ℓq), i.e. for U ∈ L(ℓp, ℓq) we let Ψn(U) = Ψ(F ′ n : ℓp(n) → ℓp =(cid:0) ⊕j∈N ℓp(j)(cid:1)p, is the canonical n), where E′ n ◦ U ◦ E′ 9 embedding into the n-th component, and F ′ onto the n-th component. Since ky∗ (n,i)kq′ ≤ C, for i = 1, 2 . . . n, it follows that kΨnk ≤ C and we let Ψ ∈ L∗(ℓp, ℓq) be a w∗-accumulation point of (Ψn). Since T (e(p,n,i)) = y(n,i) for i = 1, 2 . . . , n, it follows that Ψ(T ) = limn→∞hΨn, T i = 1. n : (cid:0) ⊕j∈N ℓq(kj)(cid:1)q → ℓq(kn) is the projection It is left to show that J I(p,2) ⊂ ker(Φ) and that J I(2,q) ⊂ ker(Ψ). To do so, we need a result which is of independent interest and will therefore be stated separately and more generally than needed. (cid:3) Definition 3.4. Let X be a finite or infinite dimensional Banach space with a normalized basis (ei). If X is infinite dimensional put for j ∈ N, and if j ≤ dim(X) < ∞, then put nX (j) = minn(cid:13)(cid:13)(cid:13)Xi∈I nX (j) = minn(cid:13)(cid:13)(cid:13)Xi∈I NX (j) = maxn(cid:13)(cid:13)(cid:13)Xi∈I ei(cid:13)(cid:13)(cid:13) : I ⊂ N, #I = jo, and NX(j) = maxn(cid:13)(cid:13)(cid:13)Xi∈I ei(cid:13)(cid:13)(cid:13) : I ⊂ {1, 2, . . . dim(X)}, #I = jo and ei(cid:13)(cid:13)(cid:13) : I ⊂ {1, 2, . . . dim(X)}, #I = jo. ei(cid:13)(cid:13)(cid:13) : I ⊂ N, #I = jo, Lemma 3.5. Assume that E and F are two finite dimensional spaces, both having Cu- unconditional and normalized bases (ei : i = 1, 2 . . . m) and (fj : j = 1, . . . n), respectively. Assume further that there are 1 < t < s < ∞ and positive constants c1, and c2, so that for all ℓ ∈ N (8) NE(ℓ) ≤ c1ℓ1/s and nF (ℓ) ≥ c2ℓ1/t. Then there exists a number c > 0, depending only on s, t, cu, c1, and c2, so that for every linear operator T : E → F and any ρ > 0 (cid:12)(cid:12)(cid:8)i ≤ m : T (ei)∞ = max j≤n f ∗ j (T (ei)) ≥ kT kρ(cid:9)(cid:12)(cid:12) ≤ cρ −s2 (s−1)(s−t) , j ) are the coordinate functionals to (fj). Moreover, if cu = c1 = c2 = 1, then we (9) where (f ∗ can choose c = 1. Corollary 3.6. Under the assumptions of Lemma 3.5, it follows that (10) 1 m mXi=1 kT (ei)k∞ ≤ kT k(1 + c)m−r(s,t), where r(s, t) = (s − 1)(s − t) (s − 1)(s − t) + s2 , for s > t ≥ 1. 10 Proof. First note that for any ρ > 0 Lemma 3.5 yields 1 m mXi=1 kT (ei)k∞ = 1 m mXi=1,kT (ei)k∞≤ρkT k kT (ei)k∞ + ≤ kT kρ + ckT k ρ −s2 (s−1)(s−t) . m 1 m mXi=1,kT (ei)k∞>ρkT k kT (ei)k∞ Then we let which implies that ρ = m− (s−1)(s−t) (s−1)(s−t)+s2 , 1 m mXi=1 kT (ei)k∞ ≤ kT km− (s−1)(s−t) (s−1)(s−t)+s2 + ckT km−1m (s−1)(s−t) (s−1)(s−t)+s2 s2 (s−1)(s−t) = kT km− (s−1)(s−t) (s−1)(s−t)+s2 + ckT km− (s−1)(s−t) (s−1)(s−t)+s2 = (1 + c)kT km−r(s,t). (cid:3) Proof of Lemma 3.5. For the sake of a better readability we will assume that c1 = c2 = cu = 1. The general case follows in the same way. We can also assume that kT k = 1. j=1 β(i, j)fj. Let ρ > 0 and put Let T : E → F and write yi = T (ei) as yi =Pn A = Aρ =(cid:8)i ∈ {1, 2, . . . , m} : max β(i, j) ≥ ρ(cid:9). For i ∈ A choose ji ∈ {1, 2 . . . , n} so that β(i, ji) ≥ ρ. Let A = {ji : i ∈ A} and for j ∈ A let Aj = {i ∈ A : ji = j}. In order to estimate Aj and then A we compute Aj1/s ≥ NE(Aj) (By (8)) sign(β(i, j))ej(cid:13)(cid:13)(cid:13)E ≥(cid:13)(cid:13)(cid:13)Xi∈Aj ≥(cid:13)(cid:13)(cid:13)T(cid:0)Xi∈Aj sign(β(i, j))ej(cid:17)(cid:13)(cid:13)(cid:13)F T(cid:16)Xi∈Aj ≥Df ∗ j ,Xi∈Aj =Xi∈Aj β(i, j) ≥ Ajρ sign(β(i, j))ej(cid:17)E (Since kT k = 1) which yields Aj1− 1 s ≤ ρ−1, and thus Since A =Pj∈ A Aj ≤ A · ρ− s (11) Aj ≤ ρ−1/(1− 1 s−1 , we obtain s ) = ρ− s s−1 . A ≥ Aρ s s−1 . 11 Let (rj)m j=1 be a Rademacher sequence on some probability space (Ω, Σ, P), this means that r1, r2, . . . rm are independent and {±1}-valued, with P({rj = 1}) = P({rj = −1}) = 1/2 for j = 1, 2 . . . n. We compute (Since kT k ≤ 1) (By 1-unconditionality of (fj)). Applying the multidimensional version of Jensen's inequality (c.f [8, 10.2.6, page 348]) j=1 zjfjkF and the Rn valued random vector Z = A1/s ≥ NE(A) (By (8)) ≥ E(cid:16)(cid:13)(cid:13)(cid:13)Xi∈A ≥ E(cid:16)(cid:13)(cid:13)(cid:13)Xi∈A = E(cid:16)(cid:13)(cid:13)(cid:13) nXj=1 riei(cid:13)(cid:13)(cid:13)E(cid:17) riβ(i, j)fj(cid:13)(cid:13)(cid:13)F(cid:17) nXj=1 fj(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)F(cid:17) to the convex function Rn ∋ z → kPn (cid:0)Pi∈A riβ(i, j) : j ≤ n(cid:1)) we obtain riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17) = E(cid:16)(cid:12)(cid:12)(cid:12)rij β(ij, j) + r Xi∈A\{ij } 2(cid:12)(cid:12)(cid:12)rij β(ij, j) + Xi∈A\{ij } riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17)(cid:13)(cid:13)(cid:13)F fjE(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A A1/s ≥(cid:13)(cid:13)(cid:13) nXj=1 ≥(cid:13)(cid:13)(cid:13)Xj∈ A A1/s ≥(cid:13)(cid:13)(cid:13)Xj∈ A E(cid:16)(cid:12)(cid:12)(cid:12)Xi∈A = E 1 For each j ∈ A there is an ij ∈ A so that β(i, j) ≥ ρ. Let r be anther ±1 random variable with P(r = 1) = P(r = −1) = 1/2, which is independent to (rj : j = 1, . . . m) then (By 1-uncondtionality of (fj)). 1 riβ(i, j)(cid:12)(cid:12)(cid:12)(cid:17) riβ(i, j)(cid:12)(cid:12)(cid:12) + 2(cid:12)(cid:12)(cid:12)rij β(ij, j) − Xi∈A\{ij } fj(cid:13)(cid:13)(cid:13)F ≥ ρ(cid:13)(cid:13)(cid:13)Xj∈ A ≥ nF ( A), riβ(i, j)(cid:12)(cid:12)(cid:12)! ≥ E(rij β(ij, j)) ≥ ρ quad (Since a + b + a − b ≥ 2a). Using again the 1-unconditionality of (fj : j = 1, 2 . . . n) we deduce therefore that and thus by our assumption (8) and by (11) we obtain A1/s ≥ nF ( A) ≥ A1/t ≥ A1/tρ s ts−t . 12 Solving for A yields which proves our claim. A ≤ ρ− s ts−t st s−t = ρ −s2 (s−1)(s−t) , (cid:3) Continuation of Proof of Theorem 3.1. In order to show that J I(p,2) ⊂ ker(Φ) we let A ∈ L2(ℓ2, ℓq) and B ∈ L(ℓp). We need to show that Φ(A ◦ I(p, 2) ◦ B) = 0. W.lo.g. we assume that kAk, kBk ≤ 1. Consider B′ canonically embedded into ℓp = (cid:0) ⊕∞ Corollary 3.6 to B′, s = 2 and t = p, we obtain n : ℓ2(n) → ℓp(kn) with B′(e(2,n,i)) = B(x(n,i)), where we consider ℓp(kn) nk ≤ C and applying therefore kB(x(n,i))k∞ = n(e(2,n,i))k∞ ≤ 2Cn−r(2,p). 1 n nXi=1 which by the concavity of [0, ∞) ∋ ξ 7→ ξ(2−p)/2 implies that ≤ (2C)(2−p)/2n−r(2,p)(2−p)/2. 1 n kB′ j=1 ℓ(kj)(cid:1). Then kB′ nXi=1 kB(x(n,i))k∞(cid:17)(2−p)/2 nXi=1 B(x(n,i))(j)2(cid:17)1/2 B(x(n,i))(j)pB(x(n,i))(j)2−p(cid:17)1/2 Secondly we observe that for any i = 1, 2 . . . n (12) (13) 1 n n ∞ kB(x(n,i))k(2−p)/2 ≤(cid:16) 1 nXi=1 kI(p, 2)(B(x(n,i)))k2 =(cid:16) knXj=1 =(cid:16) knXj=1 ≤ kB(x(n,i))k(2−p)/2 ∞ · kB(x(n,i))kp/2 p ≤ C p/2kB(x(n,i))k(2−p)/2 ∞ . It follows therefore that (cid:12)(cid:12)Φn(A ◦ I(p, 2) ◦ B)(cid:12)(cid:12)(cid:12) = = ≤ 1 1 n(cid:12)(cid:12)(cid:12) nXi=1 n(cid:12)(cid:12)(cid:12) nXi=1 nXi=1 1 n he(q′,n,i), A ◦ I(p,2) ◦ B(x(n,i))i(cid:12)(cid:12)(cid:12) hA∗(e(q′,n,i)), I(p,2) ◦ B(x(n,i))i(cid:12)(cid:12)(cid:12) nXi=1 kx(n,i)k(2−p)/2 kA∗(e(q′,n,i))k2kI(p,2) ◦ B(x(n,i)k2 ∞ ≤ kA∗kC p/2 1 n (By (13)) ≤ C p/2(2C)(2−p)/2n−r(2,p)(2−p)/2 →n→∞ 0 (By (12)). This implies that J I(p,2) ⊂ ker(Φ). 13 In order to show that J I(2,q) ⊂ ker(Ψ), let B ∈ L(ℓp, ℓ2) and A ∈ L(ℓq) with kBk, kAk ≤ 1. We need to show that Ψ(A ◦ I(2,q) ◦ B) = 0. n : ℓ2(n) → ℓq′(kn), defined by A′ Let A′ ℓq′(kn) in the canonical way as subspace of ℓq′ = (⊕∞ of (y∗ deduce that (n,i) : i = 1, 2 . . . n) that kA′ n(e(2,n,i)) = A∗(y∗ (n,i)), i = 1, 2 . . . n (we consider j=1ℓq′(kn))q). It follows from the choice nk ≤ C and from Corollary 3.6 (with s = 2 and t = q′) we 1 n nXi=1 kA∗(y∗ (n,i))k∞ = 1 n kA′(e(2,n,i))k∞ ≤ 2Cn−r(2,q′). nXi=1 Using the concavity of the function [0, ∞) ∋ ξ → ξ(2−q′)/2 we deduce (14) 1 n nXi=1 kA∗(y∗ (n,i))k(2−q′)/2 ∞ =(cid:16) 1 n nXi=1 kA∗(y∗ (n,i))k∞(cid:17)(2−q′)/2 ≤ (2C)(2−q′)/2n−r(2,q′)(2−q′)/2. It is easy to see that I(q′,2) is the adjoint of I(2,q) and we compute for i = 1, 2 . . . n (15) kI(q′, 2) ◦ A∗(y∗(n, i))k2 =(cid:16) knXj=1(cid:0)A∗(y∗(n, i))(j)(cid:1)2(cid:17)1/2 =(cid:16) knXj=1(cid:12)(cid:12)A∗(y∗(n, i))(j)(cid:12)(cid:12)q′(cid:12)(cid:12)A∗(y∗(n, i))(j)(cid:12)(cid:12)2−q′(cid:17)1/2 ≤ kA∗(y∗(n, i))k(2−q′)/2 ≤ C q′/2kA∗(y∗(n, i))k(2−q′)/2 ∞ . ky(n,i)kq′/2 q′ ∞ Therefore it follows hψn, Ui = = ≤ 1 1 n(cid:12)(cid:12)(cid:12) nXi=1 n(cid:12)(cid:12)(cid:12) nXi=1 nXi=1 1 n hA ◦ I(2,q) ◦ B(e(p,n,i)), y∗(n, i)i(cid:12)(cid:12)(cid:12) hB(e(p,n,i)), I(q′,2) ◦ A∗(y∗(n, i))i(cid:12)(cid:12)(cid:12) nXi=1 kB(e(p,n,i))k2 · kI(q′,2) ◦ A∗(y∗ kA∗(y∗(n, i))k(2−q′)/2 ∞ (n,i))k2 ≤ kBkC q′/2 1 n (By (15)) ≤ C q′/2(C + 1)(2−q′)/2n−r(2,q′)(2−q′)/2 →n→∞ 0 (By (14)). which implies our claim, and finishes the proof or Theorem 3.1. (cid:3) 14 References [1] G. Androulakis and Th. Schlumprecht, Strictly singular, non-compact operators exist on the space of Gowers and Maurey, J. London Math. Soc. 64 (2001), no. 3, 65 -- 674. [2] G. Androulakis and Th. Schlumprecht, The Banach space S is complementably minimal and subsequen- tially prime, Studia Math. 156 (2003), no. 3, 22 -- 242. [3] S.A. Argyros and R. Haydon, A hereditarily indecomposable L∞-space that solves the scalar-plus- compact, preprint, arXiv 0903.3921. [4] J. Bergh and J. Lofstrom, Interpolation spaces. An introduction. Grundlehren der Mathematischen Wissenschaften, No. 223. Springer-Verlag, Berlin-New York, 1976. [5] A. Bird, G. J. O. Jameson and N.-J. Laustsen, The GiesyJames theorem for general index p, with an application to closed ideals of operators on the p th quasi-reflexive James space, in preparation. [6] J.W. Calkin, Two-sided ideals and congruences in the ring of bounded operators in Hilbert space, Ann. of Math. 42 (1941) no.2, 839 -- 873 . [7] M. Daws, Closed ideals in the Banach algebra of operators on classical non-separable spaces, Math.Proc.Camb.Phil. Soc. 140 (2006) 317 -- 332. [8] R.M. Dudley, Real analysis and probability. Revised reprint of the 1989 original. Cambridge Studies in Advanced Mathematics, 74. Cambridge University Press, Cambridge, 2002. x+555 pp. ISBN: 0-521- 00754-2, [9] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. [10] I. S. Edelstein and B. S. Mityagin, Homotopy type of linear groups of two classes of Banach spaces, Functional Anal. Appl. 4 (1970), 221 -- 231. [11] M. Fabian, P. Habala, P. Hajek, V. Montesinos Santaluc´ıa, J. Pelant, and V. Zizler, Functional analysis and infinite-dimensional geometry. CMS Books in Mathematics/Ouvrages de Math´ematiques de la SMC, 8. Springer-Verlag, New York (2001) x+451 pp. [12] T. Figiel and N. Tomczak-Jaegermann. Projections onto Hilbertian subspaces of Banach spaces, Israel J. Math., 33 (1979) no. 2, 155 -- 171. [13] I.C. Gohberg, A.S. Markus and I.A. Feldman, Normally solvable operators and ideals associated with them, (Russian) Bul. Akad. Stiince RSS Moldoven, 76 (10) (1960) no.10 51 -- 70. English translation: American. Math. Soc. Translat. 61 (1967) 63 -- 84. [14] B. Gramsch, Eine Idealstruktur Banachscher Operatoralgebren, J. Reine Angew. Math 225 (1967) 97 -- 115. [15] N. J. Laustsen, Maximal ideals in the algebra of operators on certain Banach spaces, Proc. Edin. Math. Soc. 45 (2002), 523 -- 546. [16] N.J. Laustsen, R.J. Loy, and C.J. Read, The lattice of closed ideals in the Banach algebra of operators on certain Banach spaces, J. Funct. Anal. 214 (2004) no.1, 106 -- 131. [17] N.J. Laustsen, E. Odell, Th.Schlumprecht, and A. Zsak, Dichotomy theorems for random matrices and closed ideals of operators on(cid:0)L∞ 1(cid:1)c0 n=1 ℓn , preprint. [18] N.J. Laustsen, Th. Schlumprecht, and A. Zsak, The lattice of closed ideals in the Banach algebra of operators on a certain dual Banach space, J. of Operator Theory, 56(2006) no 2, 391-402. [19] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces. II. Springer-Verlag, Berlin, 1979. [20] R. J. Loy and G. A. Willis, Continuity of derivations on B(E) for certain Banach spaces E, J. London Math. Soc. 40 (1989), 327346. [21] E. Luft, The two-sided closed ideals of the algebra of bounded linear operators of a Hilbertspace, Czechoslovak Math. J. 18 (1968) 595 -- 605. [22] V.D. Milman, Operators of class C0 and C ∗ 0 . (Russian), Teor. Funkciı Funkcional. Anal. i Prilozen., 10 (1970) 15 -- 26. 15 [23] V.D. Milman and G. Schechtman. Asymptotic Theory of Finite Dimensional Normed Spaces, LNM 1200, Springer-Verlag, New York, 1986. [24] A. Pietsch. Operator ideals, volume 16 of Mathematische Monographien [Mathematical Monographs]. VEB Deutscher Verlag der Wissenschaften, Berlin, 1978. [25] B. Sari, Th. Schlumprecht, N. Tomczak-Jaegermann, and V. G. Troitsky, On norm closed ideals in L(ℓp, ℓq). Studia Math. 179 (2007), no. 3, 239 -- 262. [26] Th. Schlumprecht, An arbitrarily distortable Banach space, Israel J. Math, 76 (1991), no. 1-2, 8195. Department of Mathematics, Texas A&M University, College Station, Texas 77843, USA E-mail address: [email protected] 16
1212.2373
1
1212
2012-12-11T10:25:52
Muckenhoupt inequality with three measures and applications to Sobolev orthogonal polynomials
[ "math.FA" ]
We generalize the classical Muckenhoupt inequality with two measures to three under appropriate conditions. As a consequence, we prove a simple characterization of the undedness of the multiplication operator and thus of the boundedness of the zeros and the asymptotic behavior of the Sobolev orthogonal polynomials, for a large class of measures which includes the most usual examples in the literature.
math.FA
math
MUCKENHOUPT INEQUALITY WITH THREE MEASURES AND APPLICATIONS TO SOBOLEV ORTHOGONAL POLYNOMIALS E. COLORADO(1), D. PESTANA(2), J. M. RODR´IGUEZ(2),(3), AND E. ROMERA(4) Abstract. We generalize the classical Muckenhoupt inequality with two measures to three under appropri- ate conditions. As a consequence, we prove a simple characterization of the boundedness of the multiplication operator and thus of the boundedness of the zeros and the asymptotic behavior of the Sobolev orthogonal polynomials, for a large class of measures which includes the most usual examples in the literature. 1. Introduction The starting point of our work is the classical Muckenhoupt inequality, i.e., there exists a constant c > 0 such that (1.1) for any regular enough function f , where 1 < p ≤ q < ∞, and µ, ν are nonnegative σ-finite Borel measures on (0, ∞). We are interested in considering three measures instead of two in this inequality. Precisely, we look for conditions on the measures ν1, ν2, ν3 for which it is true kf kLq(µ) ≤ c kf ′kLp(ν) kf kLp(ν1) ≤ c(cid:0)kf kLp(ν2) + kf ′kLp(ν3)(cid:1) for all regular enough functions f (see the precise statement in (3.2)). The inequality (1.2) is obviously true if ν1 ≤ kν2 for some constant k. We remember the precise result about the Muckenhoupt inequality with the necessary and sufficient condition in order to (1.1) be satisfied. Theorem 1.1 (Muckenhoupt [17]). Assume 1 < p ≤ q < ∞, let µ, ν be nonnegative σ-finite Borel measures on (0, ∞). Then there exists a constant C such that (1.2) (1.3) (1.4) holds for all measurable functions f in (0, ∞) iff 0 Z x (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lq((0,∞),µ) µ ([r, ∞))1/q"Z r 0 (cid:18) dν B := sup r>0 ≤ Ckf kLp((0,∞),ν) dt(cid:19)−1/(p−1) dt#(p−1)/p < ∞. Remark 1.2. In (1.4) we assume the usual convention 0·∞ = 0. Along the paper, every density and singular part of any measure is considered with respect to the Lebesgue measure. Note that, in fact, Muckenhoupt inequality (1.3) must be satisfied for all measurable functions f such that (cid:13)(cid:13)R x (although it can be infinite); we will follow this Muckenhoupt convention. 0 f (t) dt(cid:13)(cid:13)Lq((0,∞),µ) makes sense The classical Muckenhoupt inequality (1.3) (see [17]) appears in many contexts of mathematics, see for example [15, p. 40], where we find an equivalent condition for the estimate that some measures must hold, that is in connection with the condition for the classical Ap weights, see for instance [6]. Note that (1.3) is also related with the classical Hardy inequality, which is also known as an expression of the Heisenberg Date: June 3, 2018. (1) Partially supported by Research Projects of MICINN-Spain (Refs. MTM2009-10878, MTM2010-18128). (2) Partially supported by Research Project of MICINN-Spain (Ref. MTM2009-07800). (3) Partially supported by Research Project of CONACYT-Mexico (Ref. CONACYT-UAG I0110/62/10). (4) Partially supported by Research Project of MICINN-Spain (Ref. MTM2010/00005/001). 1 2 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA uncertainty principle, first formulated as a principle of quantum mechanics in 1927, see [7]. Later on it was studied by other authors with different perspectives, see for example the classical paper by Fefferman, [4]. In harmonic analysis, estimates of different operators with respect to weights have been largely studied; in the classical book [5] we find a general presentation of the theory. The estimates on Lp with one weight are known for operators like the Hardy-Littlewood maximal operator or the Hilbert transform, for which we need the Ap weights. One can also find strong estimates with two weights where one is obtained from the other. But although the Ap condition is generalized for pairs of weights, even for the Hardy-Littlewood maximal operator it is not enough to obtain the strong estimate on Lp with two weights; this is a very active problem now in harmonic analysis. The field of application of our new Muckenhoupt inequality will be weighted Sobolev spaces, and, in particular, the multiplication operator (MO) defined by M f (z) = z f (z). In [10] these spaces are studied in the context of partial differential equations. Also in approximation theory they are of great interest. We will focus on this last topic and its relationship with Sobolev orthogonal polynomials (SOP). SOP have been widely investigated in the last years. In particular, in [8, 9], the authors showed that the expansions with SOP can avoid the Gibbs phenomenon which appears with classical orthogonal series in L2 (see also [14]). In [20, 21, 22] it was developed a theory of general Sobolev spaces with respect to measures on the real line, in order to apply it to the study of SOP. See [2] for the generalization of this theory to curves in the complex plane. Our interest in the MO arises from its relationship with SOP controlling their zeros. In the theory of SOP we don't have the usual three term recurrence relation for orthogonal polynomials in L2 so, it is really difficult to find an explicit expression for the SOP of degree n. Hence, one of the central problems in the study of these polynomials is to determine its asymptotic behavior. In [13] it was shown how to obtain the nth root asymptotic of SOP if the zeros of these polynomials are contained in a compact set of the complex plane. Although the uniform bound of the zeros of orthogonal polynomials holds for every measure with compact support in the case without derivatives, it is an open problem to bound the zeros of SOP with respect to the norm kf kW N,p(µ0,µ1,...,µN ) := N Xk=0(cid:13)(cid:13)(cid:13) f (k)(cid:13)(cid:13)(cid:13) 1/p p Lp(µk)! , (1.5) where µ0, µ1, . . . , µN are Borel measures and p = 2. The boundedness of the zeros is a consequence of the boundedness of the MO in PN,p(µ0, µ1, . . . , µN ) (the completion of the linear space of polynomials P with respect to the norm (1.5)); in fact, the zeros of the SOP are contained in the disk {z : z ≤ 2kM k} (see [13, Theorem 2.1]). If p 6= 2, then we have an analogue of SOP, precisely, we say that qn(z) = zn + an−1zn−1 + · · · + a1z + a0 is an nth monic extremal polynomial with respect to the norm in (1.5) if kqnkW N,p(µ0,µ1,...,µN ) = inf(cid:8)kqkW N,p(µ0,µ1,...,µN ) : q(z) = zn + bn−1zn−1 + · · · + b1z + b0, It is clear that there exists at least an nth monic extremal polynomial. Furthermore, it is unique if 1 < p < ∞. If p = 2, then the nth monic extremal polynomial is precisely the nth monic SOP with respect to the inner product corresponding to W N,2(µ0, µ1, . . . , µN ). In [12] the authors prove also for 1 < p < ∞ that the boundedness of the MO allows us to obtain the boundedness of the zeros and the asymptotic behavior of the extremal polynomials. It is possible to generalize these results also in the context of "nondiagonal" Sobolev norms (see [12, 18, 19]). bj ∈ R(cid:9) . In [2, 21, 22, 23, 24, 25], there are some answers to the question stated in [13] about appropriate conditions for M to be bounded: the most general results on this topic appear in [23, Theorem 4.1] and [2, Theorem 8.1]; they characterize in a simple way (in terms of equivalent norms in Sobolev spaces) the boundedness of M in PN,p(µ0, µ1, . . . , µN ). The rest of the papers mention several conditions which guarantee the equivalence of norms in Sobolev spaces, and consequently, the boundedness of M . However, these works have two objections: on the one hand, they require that the measures lead us to obtain a well defined Sobolev space (note that W 1,p(µ0, µ1) is not well defined if (µ1)s 6= 0, since when the distributional derivative is a locally integrable function, it is defined up to sets with zero Lebesgue measure); on the other hand, they obtain MUCKENHOUPT INEQUALITY WITH THREE MEASURES 3 conditions which guarantee the boundedness of M if it is defined in the Sobolev space W 1,p(µ0, µ1) instead of P1,p(µ0, µ1). In this paper we avoid these two objections. We recall now the two classical definitions of Sobolev space on a compact interval I ⊂ R (with respect to the Lebesgue measure): (1) The Sobolev space W 1,p(I) is the set of functions f ∈ Lp(I) whose distributional derivative is also a function in Lp(I). (2) The Sobolev space P1,p(I) := H 1,p(I) is the completion with respect to the Sobolev norm of W 1,p(I) of the linear space of polynomials P (or Ck(I) with k ∈ N, C∞(I), Holder spaces, etc.). Note that by construction in (2) the spaces P, Ck(I), etc, are dense in P1,p(I). In 1964 it was shown by Meyers and Serrin, see [16], that H = W , i.e., the previous definitions of Sobolev space (with respect to the Lebesgue measure) are equivalent (see also [1] and the references therein). In 1984, Kufner and Opic showed in [11] that the situation is not so simple when one considers weights instead of the Lebesgue measure; however, if the weights wj verify w−1/(p−1) ∈ L1, then they give the right definition following the philosophy of definition (1). j Following the work [11], in [20, 21] it appears a definition of Sobolev space for a large class of measures instead of weights. For general measures, it is not possible to define W 1,p(I), but it is possible to define the Sobolev space as the completion P1,p(I) of the linear space of polynomials P. (Note that it is always possible to define the completion of a normed space as the set of equivalence classes of Cauchy sequences, which generate a Banach space). Although the following is a very simple definition of Sobolev space, we show with this example the difficulties about it: Let us consider kf k2 W 1,2([0,1],µ0,µ1) := R 1 0 f 2 + f (0)2 + f ′(0)2. If we only work with polynomials or for example C1-functions, this is a well-defined norm; however, it has no meaning for other general sets In order to determine the completion P1,2([0, 1], µ0, µ1) of the polynomials with the norm of functions. k · kW 1,2([0,1],µ0,µ1) note that any function f ∈ L2((0, 1)) may be approximated in this norm by functions g ∈ C1([0, 1]) with the values of g(0) and g′(0) fixed beforehand. Therefore, the completion of the polynomials with respect to this norm is isomorphic to L2([0, 1]) × R2. Observe that given a function g in C1([0, 1]), there are infinitely many equivalence classes in L2([0, 1]) × R2 whose restrictions to L2([0, 1]) coincide almost everywhere with g. This Sobolev space is a very strange object and it shows some difficulties in our study, because we do not require a "good behavior" of µ0 and µ1. However, this kind of Sobolev norms appears in the study of SOP, and the results in this paper allow us to prove that the MO is bounded with respect to this norm. The case of one derivative (N = 1), is the most usual in applications and in the theory of Sobolev spaces and SOP. In that case, the operator M is bounded in P1,p(µ0, µ1) if and only if kf kW 1,p(µ0,µ1) ≍ kf kW 1,p(µ0+µ1,µ1) for all f ∈ P, where the simbol A ≍ B means that there exist two positive constants, k1 and k2, such that k1A ≤ B ≤ k2A. This is equivalent to for every f ∈ P and some constant c. kf kLp(µ1) ≤ c(cid:16)kf kLp(µ0) + kf ′kLp(µ1)(cid:17) (1.6) That is the main reason why we deal with three measures instead of two in inequality (1.2). The paper contains four more sections. In section 2 we establish some notation and preliminaries. Section 3 deals with the generalized Muckenhoupt inequality (3.2); Proposition 3.1 provides a very simple sufficient condition. Theorems 3.4, 3.7, 3.8 and 4.1 give several different hypotheses for which this condition is also necessary. We also prove in Theorem 3.9 that for finite measures (3.2) holds for every measurable function iff it holds for every polynomial. A counter-example in which the Muckenhoupt inequality (3.2) is not satisfied is shwon in section 4. Finally, applying the theorems in Section 3 we obtain several results in Section 5 about the MO. In particular, the sum of Theorem 5.2 and Corollary 5.10 characterizes the boundedness of the MO for a large class of measures which includes the most usual examples in the literature of orthogonal 4 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA polynomials (see Example 5.11). Furthermore, Theorems 5.9 and 5.13 and Corollary 5.15 give sufficient conditions in order to obtain the boundedness of the MO for a wider class of measures (see Example 5.14). 2. Notation and preliminaries Notation. Along the paper we just consider nonnegative σ-finite Borel measures on R. Besides: • We assume that 1 < p < ∞ in the whole work, so we omit it to simplify. • Measures are denoted by νj or µj, and densities (with respect to the Lebesgue measure) by wj . • (ν)s or (µ)s denote singular parts, and (ν)ac or (µ)ac absolutely continuous parts, with respect to the Lebesgue measure. • (ν1 − ν2)+ denotes the positive part of ν1 − ν2. • Given a measurable set A ⊂ R, we define the space of measurable functions M(A) = {f : A → R f is measurable on A}. • For a measurable set A ⊂ R, we denote by IA the characteristic function of A. • If b ∈ R, δb denotes the Dirac delta measure concentrated at {b}. • For two finite measures µ0, µ1 on [a, b], we denote by P1,p([a, b], (µ0, µ1)) or simply P1,p(µ0, µ1) the completion of the linear space of polynomials P with respect to the Sobolev norm k·kW 1,p([a,b],(µ0,µ1)). Remark 2.1. In general, (ν1 − ν2)+ makes sense if ν1 and ν2 are finite measures; however, it is possible to define (ν1 −ν2)+ for σ-finite measures as follows. Let us consider two increasing sequences of measurable sets {X j n} is an increasing n∩X 2 define the total variation ν1 − ν2 of ν1 − ν2, and its positive and negative parts (ν1 − ν2)+, (ν1 − ν2)− as nonnegative σ-finite measures on X 1 ∩ X 2. Also, it is possible to define ν1 − ν2 := (ν1 − ν2)+ := ν1 on n(cid:1) < ∞, X j := ∪nX j n} with νj(cid:0)X j n(cid:1) < ∞ and X 1 ∩ X 2 = ∪n(cid:0)X 1 sequence of measurable sets with νj(cid:0)X 1 X 1 \ X 2, ν1 − ν2 := (ν1 − ν2)− := ν2 on X 2 \ X 1, and ν1 − ν2 := 0 on R \(cid:0)X 1 n and νj(cid:0)R \ X j(cid:1) = 0 (j = 1, 2). Therefore, {X 1 (ν1 − ν2)+ and (ν1 − ν2)− are nonnegative σ-finite measures on R, although (ν1 − ν2)(E) is not defined when ν1(E) = ν2(E) = ∞. n(cid:1), and it is possible to n(cid:1). Hence, ν1 − ν2, n ∪ X 2 n ∩ X 2 n ∩ X 2 We want to generalize the Muckenhoupt condition (1.4) to the case of three measures (with exponents p = q) and to fix our interest in an interval [a, b] instead of (0, ∞). In order to do this, we rewrite B as follows. Definition 2.2. Let ν1, ν2 be measures and w2 = dν2 dx . We define: Λp,a(ν1, ν2) := Λp,[a,b],a(ν1, ν2) := sup r∈(a,b) Λp,b(ν1, ν2) := Λp,[a,b],b(ν1, ν2) := sup r∈(a,b) Λ′ p,b(ν1, ν2) := Λ′ p,[a,b],b(ν1, ν2) := sup r∈(a,b) , r ν1 ([a, r]) Z b ν1 ([r, b])(cid:18)Z r ν1 ([r, b))(cid:18)Z r a a w2(t)−1/(p−1)dt!p−1 w2(t)−1/(p−1)dt(cid:19)p−1 w2(t)−1/(p−1)dt(cid:19)p−1 , . Note that Theorem 1.1 in our setting, i.e., with (0, ∞) replaced by (a, b) and 1 < q = p < ∞, can be read as follows. Theorem 2.3. Let ν1, ν2 be measures on (a, b). There exists a constant C > 0 such that holds for all f ∈ M([a, b]) iff Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) Λ′ p,b(ν1, ν2) < ∞ . ≤ Ckf kLp([a,b],ν2) (2.1) In order to apply our results to SOP, we need to deal with measures on the compact interval [a, b]. Hence, we need the following version of Theorem 2.3 for compact intervals. MUCKENHOUPT INEQUALITY WITH THREE MEASURES 5 Theorem 2.4. Let ν1, ν2 be measures on [a, b]. There exists a constant C > 0 such that holds for all f ∈ M([a, b]) iff Z x a (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) Λp,b(ν1, ν2) < ∞ . ≤ C kf kLp([a,b],ν2) Remark 2.5. A similar result holds replacing b by a. Along the paper, most of the results will be stated just for one endpoint of the interval, but they also hold for the other one by symmetry. Proof of Theorem 2.4. Fix any measurable subset S of [a, b] with zero Lebesgue measure and such that (ν2)sS = (ν2)s. First of all, note that the singular part of ν2 does not play any role in Λp,b(ν1, ν2). For f ∈ M([a, b]) we define the function f0 := f I[a,b]\S. Then we have Z x a (cid:13)(cid:13)(cid:13)(cid:13) f0(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) Z x =(cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) and we conclude that the singular part of ν2 does not play any role in (2.1). Furthermore, if F (x) := a f (t) dt, it verifies F (a) = 0. Hence, if ν1 is a measure on [a, b) and ν2 is a measure on [a, b], then Theorem 2.3 proves that (2.1) holds iff Λ′ p,b(ν1, ν2) < ∞. R x , kf0kLp([a,b],ν2) = kf kLp([a,b],(ν2)ac) , Now, let's observe that maxnΛ′ p,b(ν1, ν2), ν1({b})(cid:16)Z b a w2(t)−1/(p−1)dt(cid:17)p−1o Therefore, if ν1({b}) = 0 then Λp,b(ν1, ν2) = Λ′ If Λp,b(ν1, ν2) < ∞ then also Λp,b(ν1 − ν1({b})δb, ν2) = Λ′ ≤ Λp,b(ν1, ν2) ≤ Λ′ p,b(ν1, ν2) + ν1({b})(cid:16)Z b a w2(t)−1/(p−1)dt(cid:17)p−1 p,b(ν1, ν2) and we are done. So, let us suppose that ν1({b}) > 0. p,b(ν1, ν2) < ∞ and by Theorem 2.3 there exists (2.2) . Hence, in order to obtain (2.1) it suffices to prove that there exists a constant C1 such that a f kLp({b}, δb) ≤ C1 kf kLp([a,b],ν2) for all f ∈ M([a, b]) or, equivalently, that ≤ C kf kLp([a,b],ν2) , ∀f ∈ M([a, b]) . a constant C such that a Z b a Z b a kR x Z x (cid:13)(cid:13)(cid:13)(cid:13) a By Holder inequality, f (t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1−ν1({b})δb) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z b (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (t)w2(t)1/pw2(t)−1/p dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Since ν1({b}) > 0 and Λp,b(ν1, ν2) < ∞ imply that (cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1−ν1({b})δb) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z b =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:12)(cid:12)(cid:12) ν1({b})1/p(cid:12)(cid:12)(cid:12)R b Using again Theorem 2.3, we deduce that Λ′ Assume now that (2.1) holds then, in particular, f (t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) a a ≤ C kf kLp([a,b],ν2) and therefore ≤ C2 kf kLp([a,b],ν2) , ∀f ∈ M([a, b]) . w−1/(p−1) (p−1)/p L1([a,b]) , ∀f ∈ M([a, b]) . < ∞ (see (2.2)), we deduce (2.1). (cid:13)(cid:13)(cid:13) 2 ≤ kf kLp([a,b],ν2)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)L1([a,b]) w−1/(p−1) 2 ≤ C kf kLp([a,b],ν2) , ∀f ∈ M([a, b]) . p,b(ν1, ν2) < ∞. Also, from (2.1), we have that ≤ C′ kf kLp([a,b],ν2) , ∀f ∈ M([a, b]) . 6 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA In particular, if we define fε = max{w2, ε}−1/(p−1)I[a,b]\S, for ε > 0, we obtain Z b a fε(t) dt ≤ C′ Z b a Therefore, max{w2(t), ε}−p/(p−1) w2(t) dt!1/p ≤ C′ Z b max{w2(t), ε}−1/(p−1)dt!(p−1)/p a Z b a max{w2(t), ε}−1/(p−1)dt!1/p < ∞ . ≤ C′. By the monotone convergence Theorem, R b a w2(t)−1/(p−1)dt < ∞ and, by (2.2), Λp,b(ν1, ν2) < ∞. (cid:3) 3. Muckenhoupt inequality with three measures Let us start with a first approach to our problem, which gives a sufficient condition for (1.2). We will prove in Sections 3 and 4 that, in many situations, this condition is also necessary. Proposition 3.1. Let ν1, ν2, ν3 be measures on [a, b]. Assume that for some constant k ≥ 0. Then there exists a constant C such that Λp,b ((ν1 − kν2)+, ν3) < ∞ , (3.1) Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν3)! , ∀ f ∈ M([a, b]). (3.2) The hypothesis (3.1) includes the two known cases: when ν1 ≤ kν2, and when the Muckenhoupt condition is fulfilled for ν1 and ν3, i.e., Λp,b(ν1, ν3) < ∞. Proof. First of all, we have ν1(E) − ν2(E) = (ν1 − ν2)(E) = (ν1 − ν2)+(E) − (ν1 − ν2)−(E) ≤ (ν1 − ν2)+(E) for every measurable set E with ν2(E) < ∞. Hence, Z b a g(x) dν1 − kZ b a g(x) dν2 ≤Z b a g(x) d(ν1 − kν2)+ (3.3) for every simple ν2-integrable function g. If g is now a nonnegative ν2-integrable function and 0 < α < β, then to g. 2.4 that there exists a constant c0 such that p ν2(cid:0)g−1([α, β))(cid:1) < ∞, and we deduce (3.3) by approximating g by simple ν2-integrable functions increasing Without loss of generality we can assume that R x holds. Therefore, we deduce from (3.3) for g(x) =(cid:12)(cid:12)R x a f (t) dt(cid:12)(cid:12) Z x Z x Z x (cid:13)(cid:13)(cid:13)(cid:13) − k(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13) for all f ∈ M([a, b]) with R x a f (t) dt ∈ Lp([a, b], ν2), since otherwise the inequality , Λp,b ((ν1 − kν2)+, ν3) < ∞ and by Theorem a f (t) dt ∈ Lp([a, b], ν2). Then, taking C := max{k, c0}1/p we conclude the f (t) dt(cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13) Lp([a,b],ν3) , ≤ c0 kf kp Lp([a,b],(ν1−kν2)+) Lp([a,b],ν1) Lp([a,b],ν2) proof. (cid:3) a a a p p p In terms of Sobolev spaces this estimate can be read as: where g ∈ W 1,p([a, b], (ν1 + ν2, ν3)), if this Sobolev space is well defined. kg − g(a)kLp(ν1) ≤ C(cid:0)kg − g(a)kLp(ν2) + kg′kLp(ν3)(cid:1) , In order to deal with the weights that one usually finds in applications, and to obtain a characterization of (3.2) in terms of them, we will use the following definition. MUCKENHOUPT INEQUALITY WITH THREE MEASURES 7 Definition 3.2. We say that an ordered pair of weights (w1, w2) on [a, b] is in the class Ca([a, b]) if we have either lim x→a+ w1(x) w2(x) = ∞ or lim sup x→a+ w1(x) w2(x) < ∞ . Similarly, (w1, w2) ∈ Cb([a, b]) if we have either lim x→b− w1(x) w2(x) = ∞ or lim sup x→b− w1(x) w2(x) < ∞ . Note that if, for example, there exists the limit lim x→a+ w1(x) w2(x) = L ∈ [0, ∞] , then (w1, w2) ∈ Ca([a, b]). Remark 3.3. The class Ca([a, b]) contains every pair of weights obtained by the products of: x − aα1 , exp(cid:0)−βx − a−α2(cid:1) , with αj ∈ R and β ≥ 0. α3 , log (cid:12)(cid:12)(cid:12)(cid:12) 1 x − a(cid:12)(cid:12)(cid:12)(cid:12) log (cid:12)(cid:12)(cid:12)(cid:12) log(cid:12)(cid:12)(cid:12)(cid:12) log(cid:12)(cid:12)(cid:12)(cid:12) · · ·(cid:12)(cid:12)(cid:12)(cid:12) 1 x − a(cid:12)(cid:12)(cid:12)(cid:12) · · ·(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) α4 , Theorem 3.4. Let ν1, ν2, ν3 be measures on [a, b]. Assume that ν1 is finite, w−1/(p−1) ∈ L1([a, r]) for every r ∈ (a, b), and (w1, w2) ∈ Cb([a, b]). Suppose also that there exist constants ε0, c0 > 0 verifying the following: (ν2)s([b− ε0, b]) = 0 if w1(x)/w2(x) < ∞. w1(x)/w2(x) = ∞, and (ν1)s ≤ c0(ν2)s on [b− ε0, b] if lim sup x→b− lim x→b− 3 Then, there exists a constant c such that Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν3)! , ∀ f ∈ M([a, b]) (3.4) iff there exists a constant k ≥ 0 such that Λp,b ((ν1 − kν2)+, ν3) < ∞ . (3.5) The following result will be useful in the proof of Theorem 3.4. Lemma 3.5. Let ν1, ν2 be measures on [a, b]. Assume that ν1 is finite and w−1/(p−1) r0 ∈ (a, b). Then, 2 ∈ L1([a, r0]) for some Λp,[a,b],b(ν1, ν2) < ∞ ⇐⇒ Λp,[r0,b],b(ν1, ν2) < ∞ . Proof of Lemma 3.5. Let us define Λp,[a,b],b,r0(ν1, ν2) := sup r∈(r0,b) ν1 ([r, b])(cid:18)Z r a w−1/(p−1) 2 (cid:19)p−1 . We are going to prove the lemma by showing the following equivalences: Λp,[a,b],b(ν1, ν2) < ∞ ⇐⇒ Λp,[a,b],b,r0(ν1, ν2) < ∞ ⇐⇒ Λp,[r0,b],b(ν1, ν2) < ∞ . (3.6) Note that, since ν1 is finite and w−1/(p−1) 2 ∈ L1 ([a, r0]), sup r∈(a,r0] ν1([r, b])(cid:18)Z r a w−1/(p−1) 2 (cid:19)p−1 ≤ ν1([a, b])(cid:18)Z r0 a w−1/(p−1) 2 (cid:19)p−1 =: J < ∞ . Then, we deduce and the first equivalence in (3.6) holds. Λp,[a,b],b,r0(ν1, ν2) ≤ Λp,[a,b],b(ν1, ν2) ≤ max(cid:8)Λp,[a,b],b,r0(ν1, ν2), J(cid:9) , 8 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA In order to prove the second one, let us define K := R r0 p > 2, and L := cp K p−1 ν1([r0, b]). We have a w−1/(p−1) 2 , cp := 1 if 1 < p ≤ 2, cp := 2p−2 if sup r∈(r0,b) ν1([r, b])(cid:18)Z r a w−1/(p−1) 2 (cid:19)p−1 = sup r∈(r0,b) ≤ sup r∈(r0,b) w−1/(p−1) 2 ν1([r, b])(cid:18)K +Z r ν1([r, b]) cp K p−1 +(cid:18)Z r r0 r0 (cid:19)p−1 w−1/(p−1) 2 (cid:19)p−1! ≤ cp K p−1 ν1([r0, b]) + cp sup r∈(r0,b) = L + cp Λp,[r0,b],b(ν1, ν2) . ν1([r, b])(cid:18)Z r r0 w−1/(p−1) 2 (cid:19)p−1 Then, we deduce which proves the second equivalence in (3.6). (cid:3) Λp,[r0,b],b(ν1, ν2) ≤ Λp,[a,b],b,r0(ν1, ν2) ≤ L + cp Λp,[r0,b],b(ν1, ν2) , Proof of Theorem 3.4. By Proposition 3.1 it suffices to prove that (3.4) implies (3.5). Therefore, let's assume that (3.4) holds. • Suppose first that w1(x)/w2(x) = ∞. Hence, we have dν2 = w2 dx on [b − ε0, b]. Let us choose 0 < ε < ε0 with w1(x)/w2(x) ≥ (2c)p for every x ∈ [b − ε, b). For any f ∈ M([a, b]) with suppf ⊆ [b − ε, b], the following holds lim x→b− Z x c (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) w2(x) dx!1/p (2c)−p w1(x) dx!1/p p p b−ε b−ε = c Z b Z x b−ε(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) ≤ c Z b Z x b−ε(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) Z x 2 (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([b−ε,b],ν1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([b−ε,b],ν1) b−ε ≤ 1 . ≤ 2 c kf kLp([b−ε,b],ν3) . Z x (cid:13)(cid:13)(cid:13)(cid:13) b−ε Therefore, by (3.4), we have for every f ∈ M([a, b]) with suppf ⊆ [b − ε, b] that: Then, by Theorem 2.4 we obtain Λp,[b−ε,b],b(ν1, ν3) < ∞. Hence, by Lemma 3.5: Λp,[a,b],b(ν1, ν3) < ∞, which is (3.5) with k = 0. • Assume now that lim sup x→b− w1(x)/w2(x) < ∞. Therefore, (ν1)s ≤ c0 (ν2)s on [b − ε0, b] and there exist constants k0 > 0 and 0 < ε < ε0 with w1(x) ≤ k0w2(x) for every x ∈ [b − ε, b). Hence, if we define k := max{c0, k0}, then ν1 ≤ kν2 on [b − ε, b] and (ν1 − kν2)+ = 0 on [b − ε, b]. Thus, Λp,b ((ν1 − kν2)+, ν3) = sup (ν1 − kν2)+ ([r, b])(cid:18)Z r (ν1 − kν2)+ ([r, b])(cid:18)Z r w3(t)−1/(p−1)dt(cid:19)p−1 w3(t)−1/(p−1)dt(cid:19)p−1 a a r∈(a,b) = sup r∈(a,b−ε) w3(t)−1/(p−1)dt!p−1 ≤ (ν1 − kν2)+ ([a, b]) Z b−ε ≤ ν1 ([a, b]) Z b−ε a a w3(t)−1/(p−1)dt!p−1 < ∞ . (cid:3) MUCKENHOUPT INEQUALITY WITH THREE MEASURES 9 As a consequence of Theorem 3.4 we have the following result. Corollary 3.6. Let ν1, ν2, ν3 be measures on [a, b]. Assume that ν1 is finite, w−1/(p−1) ∈ L1([a, r]) for every r ∈ (a, b), and (w1, w2) ∈ Cb([a, b]). Suppose also that there exists ε0 > 0 verifying (ν1)s([b − ε0, b]) = (ν2)s([b − ε0, b]) = 0. Then, there exists a constant c such that 3 Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν3)! , ∀ f ∈ M([a, b]) iff there exists a constant k ≥ 0 such that We also have a result similar to Theorem 3.4 if w−1/(p−1) 3 6∈ L1. Λp,b ((ν1 − kν2)+, ν3) < ∞ . Theorem 3.7. Let ν1, ν2, ν3 be measures on [a, b]. Assume that w−1/(p−1) /∈ L1(I) for every interval I ⊆ [a, b], and (w1, w2) is a pair in the class Cb([a, b]). Suppose also that there exist constants ε0, c0 > 0 verifying the following: 3 • (ν2)s([b − ε0, b]) = 0 if lim x→b− w1(x) w2(x) = ∞ and (ν1)s ≤ c0(ν2)s on [b − ε0, b] if lim sup x→b− w1(x) w2(x) < ∞, • for each ε > 0 there exists a constant cε > 0 with ν1 ≤ cεν2 on [a, b − ε]. Then, there exists a constant c such that Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν3)! , ∀ f ∈ M([a, b]) , (3.7) iff there exists a constant k ≥ 0 such that Proof. By Proposition 3.1 it suffices to prove (3.8) assuming that (3.7) holds. Λp,b ((ν1 − kν2)+, ν3) < ∞. (3.8) If we have lim x→b− w1(x)/w2(x) = ∞, then, as in the proof of Theorem 3.4, we can choose 0 < ε < ε0 such that for any f ∈ M([a, b]) with suppf ⊆ [b − ε, b], Z x (cid:13)(cid:13)(cid:13)(cid:13) b−ε f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([b−ε,b],ν1) ≤ 2 c kf kLp([b−ε,b],ν3) . Then, by Theorem 2.4 we obtain Λp,[b−ε,b],b(ν1, ν3) < ∞. Since w−1/(p−1) /∈ L1(I) for every interval I ⊆ [a, b], ν1 = 0 on (b − ε, b]. By hypothesis, there exists a constant k ≥ 0 with ν1 ≤ kν2 on [a, b − ε]. Hence, we conclude ν1 ≤ kν2 on [a, b], (ν1 − kν2)+ = 0 on [a, b], and Λp,b ((ν1 − kν2)+, ν3) = 0. 3 If we suppose now that lim sup x→b− w1(x)/w2(x) < ∞, then we have (ν1)s ≤ c0(ν2)s on [b − ε0, b] and there exist constants k0 > 0 and 0 < ε < ε0 with w1(x) ≤ k0w2(x) for every x ∈ [b − ε, b). Thus, taking k1 := max{c0, k0}, we have ν1 ≤ k1ν2 on [b − ε, b]. Since ν1 ≤ cεν2 on [a, b − ε], if we define k := max{cε, k1}, then ν1 ≤ kν2 on [a, b] and (ν1 − kν2)+ = 0 on [a, b], and finally, Λp,b ((ν1 − kν2)+, ν3) = 0. (cid:3) For the case of weights comparable to monotone functions we show in the next theorem sufficient conditions to obtain our estimate. Theorem 3.8. Let ν1, ν2 be measures on [a, b]. Assume that we have either (1) (ν1)s ≤ k0(ν2)s for some constant k0 and w1 is comparable to a non-increasing function on [a, b], or (2) ν1 is a finite measure, w−1/(p−1) 1 ∈ L1([a, a + ε]) for some ε > 0, and w1 is comparable to a non- decreasing function on [a, b]. Then, there exists a constant c such that Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν1)! , ∀ f ∈ M([a, b]) , 10 and E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA for some constant k ≥ 0. Λp,b ((ν1 − kν2)+, ν1) < ∞ (3.9) Proof. By Proposition 3.1, it suffices to prove (3.9). Without loss of generality we can assume that w1 is a monotone function on [a, b]. • Assume first that (ν1)s ≤ k0(ν2)s for some constant k0 and that w1 is a non-increasing function on [a, b]. Then (ν1 − k0ν2)+ ≤ (ν1)ac, and it suffices to prove that Λp,b ((ν1)ac, ν1) < ∞, since then (3.9) holds. We have Λp,b ((ν1)ac, ν1) = sup r∈(a,b) (ν1)ac ([r, b])(cid:18)Z r w1(t) dt!(cid:18)Z r w1(t)−1/(p−1)dt(cid:19)p−1 w1(t)−1/(p−1)dt(cid:19)p−1 a a = sup r∈(a,b) Z b r w1(r) (b − r) w1(r)−1(r − a)p−1 ≤ sup r∈(a,b) ≤ (b − a)p < ∞ . • Assume now that ν1 is a finite measure, w−1/(p−1) ∈ L1([a, a + ε]) for some ε > 0, and w1 is a non- decreasing function on [a, b]. In this case we have w1(x) ≥ w1(a + ε) > 0 for every x ∈ [a + ε, b] and we conclude that w−1/(p−1) ∈ L1([a, b]). Therefore, for any k ≥ 0, 1 1 Λp,b ((ν1 − kν2)+, ν1) ≤ Λp,b (ν1, ν1) = sup r∈(a,b) ν1 ([r, b])(cid:18)Z r w1(t)−1/(p−1)dt!p−1 a w1(t)−1/(p−1)dt(cid:19)p−1 < ∞ . (cid:3) ≤ ν1 ([a, b]) Z b a We finish this section with a result on polynomial approximation for the Muckenhoupt inequality with three measures. Theorem 3.9. Let ν1, ν2, ν3 be finite measures on [a, b] with w−1/(p−1) Then there exists a constant c > 0 such that the following inequality 3 ∈ L1([a, r]) for every r ∈ (a, b). Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kf kLp([a,b],ν3)! (3.10) holds ∀ f ∈ M([a, b]) iff it holds for any polynomial. belongs to Lp([a, b], ν3). Proof. Let us assume that (3.10) holds for every polynomial and define c1 := max(cid:8)ν1([a, b])1/p, ν2([a, b])1/p, ν3([a, b])1/p(cid:9). Fix a function f ∈ M([a, b]) and ε > 0; without loss of generality we can assume that f Let's assume first that f ∈ L1([a, b]). The classical proof of the density of the continuous functions in Lp (using the density of the simple functions and Lusin's Theorem) gives, in fact, that there exists a function g ∈ C([a, b]) with Weierstrass' Theorem provides a polynomial q with kg − qkL∞([a,b]) < ε. Hence, kf − gkLp([a,b],ν3) + kf − gkL1([a,b]) < ε . kg − qkLp([a,b],ν3) + kg − qkL1([a,b]) < ε ν3([a, b])1/p + ε (b − a) ≤ (c1 + b − a) ε , kf − qkLp([a,b],ν3) + kf − qkL1([a,b]) < (c1 + b − a + 1) ε =: c2 ε , f (t) dt −Z x a Z x (cid:13)(cid:13)(cid:13)(cid:13) a q(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],νj) ≤ kf − qkL1([a,b])νj([a, b])1/p < c1 c2 ε . MUCKENHOUPT INEQUALITY WITH THREE MEASURES 11 Since ε > 0 is arbitrary, (3.10) holds if f ∈ L1([a, b]). Now, let's suppose that f /∈ L1([a, b]). Since w−1/(p−1) 3 ∈ L1([a, x]) for every x ∈ (a, b), we have by Holder inequality, Z x a f (t) dt =Z x a f (t) w3(t)1/pw3(t)−1/p dt ≤ kf kLp([a,b],ν3)(cid:13)(cid:13)(cid:13) sequence with lim n→∞ and then f ∈ L1([a, x]) for every x ∈ (a, b). If the functionR x of b, let {bn} be an increasing sequence withR bn ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x fn(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) Z x (cid:13)(cid:13)(cid:13)(cid:13) a a a f = 0 and lim n→∞ fn(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) + kfnkLp([a,b],ν3)! w−1/(p−1) 3 (p−1)/p L1([a,x]) < ∞ , (cid:13)(cid:13)(cid:13) a f has infinitely many zeros in any neighborhood bn = b; otherwise, let {bn} be any increasing bn = b. Consider the sequence of functions fn := f I[a,bn] ∈ L1([a, b]); we have proved for every n. Since fn and (cid:12)(cid:12)R x Theorem. a fn(t) dt(cid:12)(cid:12) increase with n, (3.10) holds for f by the monotone convergence (cid:3) 4. A negative condition We show in this section a class of measures which do not satisfy the generalization of Muckenhoupt inequality (3.2). Theorem 4.1. Let ν1, ν2, ν3 be measures on [a, b]. Assume that there exists b0 ∈ [a, b) such that ν2([b0, b]) < ∞, and w−1/(p−1) /∈ L1([r, b]) for every r ∈ (b0, b). If ν1({b}) > 0 and ν2({b}) = 0, then there is no constant 3 c for which Z x (cid:13)(cid:13)(cid:13)(cid:13) a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν1) with ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) Λp,b ((ν1 − kν2)+, ν3) = ∞, + kf kLp([a,b],ν3)! , ∀ k ≥ 0. ∀ f ∈ M([a, b]) , (4.1) Proof. By Proposition 3.1, it suffices to prove the first statement. Assume first that w−1/(p−1) 3 ∈ L1([b0, r]) for every r ∈ (b0, b). Since ν1({b}) > 0, it suffices to show that there does not exist any constant c verifying ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a a (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z b f (t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) there exists bn ∈ (an, b) with R bn fn an 3 f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],ν2) Arguing by contradiction, let us suppose that there exists c > 0 satisfying (4.2). define an := max{b0, b − 1/n}. Since w−1/(p−1) Let S be a set with zero Lebesgue measure and such that (ν3)sS = (ν3)s. For every natural number n, we /∈ L1([r, b]) for every r ∈ (b0, b), I[an,bn]\S. By (4.2) applied to = n; let us define fn := w−1/(p−1) ∈ L1([b0, r]) and w−1/(p−1) w−1/(p−1) 3 3 3 + kf kLp([a,b],ν3)! , ∀ f ∈ M([a, b]). (4.2) n ≤ c n ν2([an, b])1/p + c Z bn an w−p/(p−1) 3 w3!1/p ≤ c n ν2([an, b])1/p + c n1/p . Since ν2([b0, b]) < ∞, we deduce that ν2([an, b]) = ν2({b}) = 0. Hence, there exists some n0 ∈ N such that c ν2([an, b])1/p ≤ 1/2, ∀ n ≥ n0. Therefore, n ≤ 2 c n1/p, ∀ n ≥ n0, which is a contradiction since 1 < p < ∞. Then the conclusion holds if w−1/(p−1) ∈ L1([b0, r]) for every r ∈ (b0, b). lim n→∞ 3 We deal now with the general case. Since w−1/(p−1) increasing sequence {cn} ⊂ (b0, b) with lim n→∞ /∈ L1([r, b]) for every r ∈ (b0, b), we can choose an ∈ [1, ∞] for every n. Let {nj} be w−1/(p−1) 3 3 cn = b and R cn cn−1 12 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA the ordered set of indices n with R cn cn−1 w−1/(p−1) 3 = ∞, if any. For each j, let us choose εj ≥ 0 verifying 1 ≤Z cnj cnj −1 max {w3, εj}−1/(p−1) < ∞ . If {nj} = ∅, then we define w3 := w3. Otherwise, we define w3 :=( w3 , max {w3, εj} , on [cn−1, cn] if n /∈ {nj} , on (cnj −1, cnj ) if n = nj , and ν3 by dν 3 := w3 dx + d(ν3)s. Note that Z cn cn−1 w−1/(p−1) 3 < ∞ ∀n, and moreover w−1/(p−1) 3 Z b b0 ≥ Xn /∈{nj }Z cn cn−1 w−1/(p−1) 3 + card {nj} . As a consequence, w−1/(p−1) proved that (4.1) is not satisfied with ν3 instead of ν3. Since ν3 ≤ ν3, the conclusion holds for ν3. /∈ L1([r, b]) for any r ∈ (b0, b). Therefore, we have (cid:3) ∈ L1([b0, r]) and w−1/(p−1) 3 3 5. Application to Sobolev orthogonal polynomials We start with the introduction of the concept of regular points, which will be the basis of the results of this section. Definition 5.1. Let µ1 be a measure on [a, b]. If w−1/(p−1) then we define the interval of regular points Reg([a, b]) as follows: ∈ L1([a, b]), then Reg([a, b]) = [a, b]. (1) In the case w−1/(p−1) 1 1 ∈ L1([a + ε, b − ε]) for every 0 < ε < (b − a)/2, Moreover, if there exists ε > 0 such that: (2) w−1/(p−1) (3) w−1/(p−1) (4) w−1/(p−1) ∈ L1([a, a + ε]) and w−1/(p−1) /∈ L1([a, a + ε]) and w−1/(p−1) /∈ L1([a, a + ε]) and w−1/(p−1) 1 1 1 1 1 1 /∈ L1([b − ε, b]), then Reg([a, b]) = [a, b). ∈ L1([b − ε, b]), then Reg([a, b]) = (a, b]. /∈ L1([b − ε, b]), then Reg([a, b]) = (a, b). The concept of Reg([a, b]) is natural, as the following results show. Theorem 5.2. Let µ0, µ1 be finite measures on [a, b]. Assume that w−1/(p−1) 0 < ε < (b − a)/2. 1 ∈ L1([a + ε, b − ε]) for every (1) In the case Reg([a, b]) = [a, b], then the MO is bounded in P1,p(µ0, µ1) iff µ0 ([a, b]) > 0. (2) If Reg([a, b]) = [a, b), we assume also that (w1, w0) ∈ Cb([a, b]) and (µ0)s([b−ε, b]) = (µ1)s([b−ε, b]) = 0 for some ε > 0. Then the MO is bounded in P1,p(µ0, µ1) iff µ0 ([a, b]) > 0 and Λp,[a,b],b ((µ1 − kµ0)+, µ1) < ∞ , (5.1) for some constant k. (3) For Reg([a, b]) = (a, b], we assume also that (w1, w0) ∈ Ca([a, b]) with (µ0)s([a, a + ε]) = (µ1)s([a, a + ε]) = 0 for some ε > 0. Then the MO is bounded in P1,p(µ0, µ1) iff µ0 ([a, b]) > 0 and Λp,[a,b],a ((µ1 − kµ0)+, µ1) < ∞ , (5.2) for some constant k. (4) When Reg([a, b]) = (a, b), we assume also that (w1, w0) ∈ Ca([a, b]) ∩ Cb([a, b]), and (µ0)s([a, a + ε]) = (µ1)s([a, a + ε]) = (µ0)s([b − ε, b]) = (µ1)s([b − ε, b]) = 0 for some ε > 0. Let us fix x0 ∈ (a, b). Then the MO is bounded in P1,p(µ0, µ1) iff µ0 ([a, b]) > 0 and Λp,[a,x0],a ((µ1 − kµ0)+, µ1) < ∞ , Λp,[x0,b],b ((µ1 − kµ0)+, µ1) < ∞ , (5.3) for some constant k. MUCKENHOUPT INEQUALITY WITH THREE MEASURES 13 Remark 5.3. Note that the hypotheses in Theorem 5.2 imply µ0 ([a, b] \ Reg([a, b])) = µ1 ([a, b] \ Reg([a, b])) = 0, i.e., µ0Reg([a,b]) = µ0 and µ1Reg([a,b]) = µ1. We need finite measures since it is necessary to have kgkW 1,p([a,b],(µ0,µ1)) < ∞, ∀g ∈ P. We also want to point out that in [3] the compactness of the supports of the measures is a necessary condition in order to have the boundedness of the MO. Let us first prove the following lemmas, which will be useful in the proof of Theorem 5.2. Lemma 5.4. Let µ0, µ1 be measures on [a, b]. Assume that 0 < µ0([a, b]) < ∞ and w−1/(p−1) Then: 1 ∈ L1([a, b]). (1) There exists a positive constant c1 such that ∀ c ∈ R, and all f ∈ M([a, b]): (2) If µ1 is finite, then there exists a positive constant c2 such that ∀ c ∈ R, and all f ∈ M([a, b]) we have: a c +Z x (cid:13)(cid:13)(cid:13)(cid:13) c +Z x (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)L∞([a,b]) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],µ1) a ≤ c1 (cid:13)(cid:13)(cid:13)(cid:13) ≤ c2 (cid:13)(cid:13)(cid:13)(cid:13) c +Z x a c +Z x a f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],µ0) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],µ0) + kf kLp([a,b],µ1)! . + kf kLp([a,b],µ1)! . Proof. We just need to prove (1), since (2) is a direct consequence of (1). Let us fix x0 ∈ [a, b]. For any x ∈ [a, b] and f ∈ M([a, b]), using Holder inequality c +Z x0 (cid:12)(cid:12)(cid:12)(cid:12) a p a a a f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) c +Z x ≤(cid:12)(cid:12)(cid:12)(cid:12) c +Z x ≤(cid:12)(cid:12)(cid:12)(cid:12) c +Z x =(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2p−1(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) ≤ c4 (cid:13)(cid:13)(cid:13)(cid:13) c +Z x ≤ c1 (cid:13)(cid:13)(cid:13)(cid:13) c +Z x0 a (cid:12)(cid:12)(cid:12)(cid:12) c +Z x0 (cid:12)(cid:12)(cid:12)(cid:12) c +Z x (cid:13)(cid:13)(cid:13)(cid:13) a p a f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)L∞([a,b]) and all x0 ∈ [a, b]. Therefore, we conclude that for all c ∈ R Since 0 < µ0([a, b]) < ∞, we can integrate in the x-variable on [a, b] with respect to µ0 in order to find a constant c4 > 0 such that for every c ∈ R, f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) c +Z x a +Z b a f w1/p 1 w−1/p 1 + kf kLp([a,b],w1) kw−1/(p−1) 1 k + c3 kf kLp([a,b],w1) , p−1 p L1([a,b]) p + cp 3 kf kp Lp([a,b],µ1)(cid:19) . f (t) dt(cid:12)(cid:12)(cid:12)(cid:12) p a Lp([a,b],µ0) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13) c +Z x f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,b],µ0) a + kf kp Lp([a,b],µ1)! , ∀f ∈ M([a, b]) + kf kLp([a,b],µ1)! , ∀ f ∈ M([a, b]). (cid:3) Lemma 5.5. Let µ0, µ1 be finite measures on [a, b]. Assume that w−1/(p−1) 0 < ε < (b − a)/2 and that Reg([a, b]) = (a, b). Let us fix x0 ∈ (a, b). If µ0 ((a, b)) > 0 and 1 ∈ L1([a + ε, b − ε]) for every Λp,[a,x0],a ((µ1 − kµ0)+, µ1) < ∞ , Λp,[x0,b],b ((µ1 − kµ0)+, µ1) < ∞ , (5.4) for some constant k, then the MO is bounded in P1,p(µ0, µ1). 14 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA Proof. Proposition 3.1 proves that there exists a constant c such that for all f ∈ M([a, b]). Then x0 Z x (cid:13)(cid:13)(cid:13)(cid:13) Z x (cid:13)(cid:13)(cid:13)(cid:13) x0 x0 ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([x0,b],µ0) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,x0],µ0) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([x0,b],µ1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,x0],µ1) kg − g(x0)kLp([x0,b],µ1) ≤ c(cid:16)kg − g(x0)kLp([x0,b],µ0) + kg′kLp([x0,b],µ1)(cid:17) , kg − g(x0)kLp([a,x0],µ1) ≤ c(cid:16)kg − g(x0)kLp([a,x0],µ0) + kg′kLp([a,x0],µ1)(cid:17) , + kf kLp([x0,b],µ1)! , + kf kLp([a,x0],µ1)! , x0 are satisfied for every g ∈ P. Consequently, kgkLp([x0,b],µ1) ≤ c(cid:16)kgkLp([x0,b],µ0) + kg′kLp([x0,b],µ1)(cid:17) + g(x0)(cid:16)µ1([x0, b])1/p + c µ0([x0, b])1/p(cid:17) , kgkLp([a,x0],µ1) ≤ c(cid:16)kgkLp([a,x0],µ0) + kg′kLp([a,x0],µ1)(cid:17) + g(x0)(cid:16)µ1([a, x0])1/p + c µ0([a, x0])1/p(cid:17) , hold for any g ∈ P. (5.5) Since µ0 ((a, b)) > 0, there exist a < a0 < b0 < b with x0 ∈ [a0, b0] and µ0 ([a0, b0]) > 0. Taking into account that w−1/(p−1) 1 ∈ L1([a0, b0]), by Lemma 5.4 there exists a constant c1 > 0 such that By this inequality and (5.5), there exists a constant c2 > 0 such that g(x0) ≤ c1 (cid:16)kgkLp([a0,b0],µ0) + kg′kLp([a0,b0],µ1)(cid:17) , ∀ g ∈ P. as a consequence, the MO is bounded in P1,p(µ0, µ1) by (1.6). kgkLp(µ1) ≤ c2(cid:16)kgkLp(µ0) + kg′kLp(µ1)(cid:17) , ∀ g ∈ P, (cid:3) Proof of Theorem 5.2. • We prove first that if µ0 ([a, b]) = 0, then the MO is not bounded in P1,p(µ0, µ1). By contradiction, let us suppose that the MO is bounded in P1,p(µ0, µ1). Therefore by (1.6) there exists a constant c > 0 such that Taking f = 1, we obtain kf kLp(µ1) ≤ c(cid:16)kf kLp(µ0) + kf ′kLp(µ1)(cid:17) , µ1 ([a, b]) ≤ cpµ0 ([a, b]) = 0 , ∀ f ∈ P. then we conclude µ1 ([a, b]) = 0, which is a contradiction with w−1/(p−1) 0 < ε < (b − a)/2, and we deduce that the MO is not bounded in P1,p(µ0, µ1). In order to prove part (1) we assume that µ0 ([a, b]) > 0. Since w−1/(p−1) 1 1 ∈ L1([a + ε, b − ε]) for every ∈ L1([a, b]), by Lemma 5.4 there exists a constant c2 > 0 such that Therefore the MO is bounded in P1,p(µ0, µ1). kgkLp(µ1) ≤ c2 (cid:16)kgkLp(µ0) + kg′kLp(µ1)(cid:17) , ∀ g ∈ P. • In order to prove part (2), we assume that the MO is bounded in P1,p(µ0, µ1). Then, there exists a constant c > 0 such that In particular we have kgkLp(µ1) ≤ c(cid:16)kgkLp(µ0) + kg′kLp(µ1)(cid:17) , f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ0) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x a Z x (cid:13)(cid:13)(cid:13)(cid:13) a + kf kLp(µ1)! , ∀ g ∈ P. (5.6) ∀ f ∈ P. (5.7) By Theorem 3.9 we know that (5.7) holds for all f ∈ M([a, b]). Hence, applying Corollary 3.6 we obtain (5.1) for some constant k. MUCKENHOUPT INEQUALITY WITH THREE MEASURES 15 Let's assume now that µ0 ([a, b]) > 0 and that (5.1) holds for some constant k. By Proposition 3.1 there exists a constant c > 0 such that Z x (cid:13)(cid:13)(cid:13)(cid:13) a Then Consequently, ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ1) kg − g(a)kLp(µ1) ≤ c(cid:16)kg − g(a)kLp(µ0) + kg′kLp(µ1)(cid:17) , f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ0) + kf kLp(µ1)! , a ∀ f ∈ M([a, b]). ∀ g ∈ P. kgkLp(µ1) ≤ c(cid:16)kgkLp(µ0) + kg′kLp(µ1)(cid:17) + g(a)(cid:16)µ1([a, b])1/p + c µ0([a, b])1/p(cid:17) , Since µ0 ([a, b)) > 0, there exists a < b0 < b with µ0 ([a, b0]) > 0. Taking into account that w−1/(p−1) L1([a, b0]), by Lemma 5.4 there exists a constant c1 > 0 such that 1 ∈ ∀ g ∈ P. (5.8) This inequality jointly with (5.8) give (5.6), and then the MO is bounded in P1,p(µ0, µ1). g(a) ≤ c1(cid:16)kgkLp([a,b0],µ0) + kg′kLp([a,b0],µ1)(cid:17) , ∀ g ∈ P. A similar argument to the one in part (2) allows us to prove part (3). • Finally, let us prove part (4). Fix x0 ∈ (a, b). Assume first that the MO is bounded in P1,p(µ0, µ1). Then (5.6) holds for every polynomial. In particular, We are going to prove that ∀ f ∈ P. x0 x0 x0 + kf kLp(µ1)! , ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x Z x (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µ0) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x Z x (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([x0,b],µ1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([x0,b],µ0) ≤ c (cid:13)(cid:13)(cid:13)(cid:13) Z x Z x (cid:13)(cid:13)(cid:13)(cid:13) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,x0],µ1) f (t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp([a,x0],µ0) (cid:13)(cid:13)f I[x0,b] − h0(cid:13)(cid:13)Lp(µ0+µ1) +(cid:13)(cid:13)f I[x0,b] − h0(cid:13)(cid:13)L1([a,b]) < ε . x0 x0 x0 + kf kLp([x0,b],µ1)! + kf kLp([a,x0],µ1)! (5.9) (5.10) hold for every f ∈ P. By symmetry, it suffices to prove the first inequality. density of the continuous functions in Lp, there exists a function h0 ∈ C([a, b]) with for f ∈ P and ε > 0, by the Weierstrass' Theorem provides a polynomial h with kh0 − hkL∞([a,b]) < ε. Let us define the constant c0 := (µ0 + µ1)([a, b])1/p. We have kh0 − hkLp(µ0+µ1) + kh0 − hkL1([a,b]) < ε c0 + ε (b − a) , f (t) dt I[x0,b](x) −Z x x0 (cid:13)(cid:13)f I[x0,b] − h(cid:13)(cid:13)Lp(µ0+µ1) +(cid:13)(cid:13)f I[x0,b] − h(cid:13)(cid:13)L1([a,b]) < (c0 + b − a + 1) ε , h(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µi) Z x =(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)f I[x0,b] − h(cid:13)(cid:13)L1([a,b]) µi([a, b])1/p < c0 (c0 + b − a + 1) ε . f (t)I[x0,b](t) dt −Z x h(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Lp(µi) x0 x0 Since ε > 0 is arbitrary, these inequalities and (5.9) prove the first one in (5.10). Z x (cid:13)(cid:13)(cid:13)(cid:13) x0 Moreover, taking into account that the inequalities (5.10) hold for all polynomials, then by Theorem 3.9 both inequalities in (5.10) hold for all f ∈ M([a, b]). Hence, by Corollary 3.6 we obtain (5.3) for some constant k. If we assume that (5.3) holds, then by Lemma 5.5 we conclude. (cid:3) Now, we consider a new kind of measures. 16 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA Definition 5.6. Let µ1 be a measure on R. We say that µ1 is piecewise regular if there exist real numbers a0 < a1 < · · · < am verifying the following properties: (a) The convex hull of the support of µ1 is the compact interval [a0, am]. (b) If 1 ≤ j ≤ m we have either w−1/(p−1) ∈ L1([aj−1 + ε, aj − ε]) for every 0 < ε < (aj − aj−1)/2 or 1 µ1(aj−1,aj ) is a finite linear combination of Dirac deltas. (c) For 0 < j < m we have w−1/(p−1) 1 /∈ L1([aj − ε, aj + ε]) for every ε > 0 and we do not have w1 = 0 a.e. in [aj−1, aj+1]. We say that µ1 is strongly piecewise regular if it is piecewise regular and it verifies the following property: /∈ L1([aj, aj +ε]) /∈ L1([aj −ε, aj]), and if w−1/(p−1) /∈ L1([aj −ε, aj]) then w−1/p 1 for 0 ≤ j ≤ m, if w−1/(p−1) then w−1/p /∈ L1([aj, aj + ε]). 1 1 We define J as the set of indices 1 ≤ j ≤ m with w−1/(p−1) aj−1)/2, while H will be the (finite) set of points x ∈ [a0, am] verifying µ1({x}) > 0, w−1/(p−1) and w−1/(p−1) /∈ L1([x, x + ε]) for every ε > 0. 1 1 ∈ L1([aj−1 + ε, aj − ε]) for every 0 < ε < (aj − /∈ L1([x−ε, x]) 1 1 Remarks 5.7. (1) Condition (c) is not a real restriction, since it just guarantees the uniqueness of m and the real numbers a0 < a1 < · · · < am. (2) By condition (b) we have that either w1 = 0 a.e. in [aj−1, aj] or w1 > 0 a.e. in [aj−1, aj], for 1 ≤ j ≤ m. (3) If x ∈ ∪j /∈J (aj−1, aj) verifies µ1({x}) > 0, then x ∈ H. (4) Note also that the practical totality of the measures with compact support in R which one usually deals with in the study of orthogonal polynomials is strongly piecewise regular (see Example 5.11). (5) The class of piecewise regular measures allows us to consider (and this was not the case in the papers [21], [24] and [25]) measures for which the Sobolev space W 1,p([a, b], (µ0, µ1)) is not well defined and P1,p([a, b], (µ0, µ1)) is not a space of functions (e.g., kf kp −1 f (x)pdx + −1 f ′(x)p(1 − x2)p−1dx + αf ′(−1)p + βf ′(1)p with α, β ≥ 0 and α + β > 0, or the example at the introduction kf kp W 1,p([−1,1],µ0,µ1) := R 1 0 f (x)pdx + f (0)p + f ′(0)p). R 1 W 1,p([0,1],µ0,µ1) :=R 1 The following result (see [23, Theorem 4.4]) will be necessary: Theorem 5.8. Let µ0, µ1 be finite measures on [a, b], and α ∈ [a, b]. Assume that w−1/(p−1) and w−1/(p−1) in P1,p([a, b], (µ0, µ1)). /∈ L1([α − ε, α]) /∈ L1([α, α + ε]) for every ε > 0. If µ1({α}) > 0 and µ0({α}) = 0, then the MO is not bounded 1 1 Theorem 5.9. Let µ0, µ1 be finite measures on [a, b]. (1) Assume that µ1 is piecewise regular with a0 < a1 < · · · < am. If the MO is bounded in the space P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])) for each j ∈ J and µ0({x}) > 0 for all x ∈ H, then it is bounded in P1,p(µ0, µ1). (2) Suppose that µ1 is strongly piecewise regular with a0 < a1 < · · · < am. If the MO is bounded in P1,p(µ0, µ1), then it is bounded in P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])) for each j ∈ J and µ0({x}) > 0 for all x ∈ H. Proof. First of all, note that by the definition of piecewise regular measure we have: µ1 =Xj∈J µ1Reg([aj−1 ,aj]) + Xx∈H µ1({x})δx . (5.11) Let's assume that µ1 is piecewise regular, that the MO is bounded in P1,p(µ0Reg([aj−1 ,aj ]), µ1Reg([aj−1,aj ])) for each j ∈ J and that µ0({x}) > 0 for every x ∈ H. Then, for every j ∈ J, there exists a constant cj such MUCKENHOUPT INEQUALITY WITH THREE MEASURES 17 that (see (1.6)) kgkLp(µ1Reg([aj−1 ,aj ])) ≤ cj(cid:16)kgkLp(µ0Reg([aj−1 ,aj ])) + kg′kLp(µ1Reg([aj−1 ,aj ]))(cid:17) ≤ cj(cid:16)kgkLp(µ0) + kg′kLp(µ1)(cid:17) , kgkLp(µ1H ) ≤(cid:18) maxx∈H µ1({x}) ≤(cid:18) maxx∈H µ1({x}) minx∈H µ0({x})(cid:19)1/p minx∈H µ0({x})(cid:19)1/p kgkLp(µ0H ) kgkLp(µ0) , are true for every g ∈ P. These inequalities and (5.11) give (5.6), and then the MO is bounded in P1,p(µ0, µ1). Let's assume now that µ1 is strongly piecewise regular and that the MO is bounded in P1,p(µ0, µ1). Therefore, there exists a positive constant c such that kf kLp(µ1) ≤ c(cid:16)kf kLp(µ0) + kf ′kLp(µ1)(cid:17) , ∀ f ∈ P. (5.12) Since the MO is bounded in P1,p(µ0, µ1), (5.11) and Theorem 5.8 prove that µ0({x}) > 0 for every x ∈ H. Let us prove now that the MO is bounded in P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])) for each j ∈ J. Let us fix j ∈ J. Note that in order to check that there exists a constant c > 0 such that kgkLp(µ1Reg([aj−1 ,aj ])) ≤ c(cid:16)kgkLp(µ0Reg([aj−1 ,aj ])) + kg′kLp(µ1Reg([aj−1 ,aj ]))(cid:17) , ∀ g ∈ P, by (5.12) it suffices to prove that given g ∈ P and ε > 0, there exists g0 ∈ P with (cid:12)(cid:12)(cid:12) kgkLp(µiReg([aj−1 ,aj ])) − kg0kLp(µi)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) 0kLp(µ1)(cid:12)(cid:12)(cid:12) ≤ ε , kg′kLp(µ1Reg([aj−1 ,aj ])) − kg′ ≤ ε . (5.13) Assume first that Reg([aj−1, aj]) = [aj−1, aj]. Then we have w−1/p /∈ L1([aj, aj + ε′]) for every ε′ > 0. Fixed g ∈ P and ε > 0, we define K := max{g(aj−1), g(aj)}. Since µ1, µ2 are finite, there exists 0 < η < εp with /∈ L1([aj−1 − ε′, aj−1]) and w−1/p 1 1 µ0 ((aj−1 − η, aj−1)) , µ1 ((aj−1 − η, aj−1)) , µ0 ((aj , aj + η)) , µ1 ((aj, aj + η)) < εp. Since w−1/p 1 /∈ L1 ([aj−1 − η, aj−1]) and w−1/p 1 /∈ L1([aj , aj + η]), there exist t1, t2 > 0 verifying Z aj−1 1 aj−1−η minnw−1/p Let's define g1 := g(aj−1) minnw−1/p [a, b] with zero Lebesgue measure and such that (µ1)sS = (µ1)s. Let f1(x) :=R x the following , t2o = 1 . , t2o. Fixed a measurable set S ⊆ a f0, where f0 is defined by 1 1 1 aj Z aj +η minnw−1/p , t1o = 1 , , t1o and g2 := g(aj) minnw−1/p   on (−∞, aj−1 − η] , on (aj−1 − η, aj−1) , on [aj−1, aj] , on (aj, aj + η) , on [aj + η, ∞) . g1IR\S , g′ , −g2IR\S , 0 , 0 , f0 := 18 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA We have f1 = g on [aj−1, aj] and f1 = 0 on (−∞, aj−1 − η] ∪ [aj + η, ∞). Hence, (cid:12)(cid:12)(cid:12) kgkLp(µiReg([aj−1 ,aj ])) − kf1kLp(µi)(cid:12)(cid:12)(cid:12) kg′kLp(µ1Reg([aj−1 ,aj ])) − kf0kLp(µ1)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) ≤ kf1kLp(µi(aj−1 −η,aj−1 )) + kf1kLp(µi(aj ,aj +η)) ≤ K µi ((aj−1 − η, aj−1))1/p + K µi ((aj, aj + η))1/p ≤ 2 K ε , ≤ kg1kLp(w1(aj−1 −η,aj−1 )) + kg2kLp(w1(aj ,aj +η)) ≤ K Z aj−1 aj−1−η(cid:12)(cid:12)(cid:12) = 2 K η1/p ≤ 2 K ε . w−1/p 1 w1!1/p + K Z aj +η aj p (cid:12)(cid:12)(cid:12) w−1/p 1 (cid:12)(cid:12)(cid:12) w1!1/p p (cid:12)(cid:12)(cid:12) Since f0 ∈ L1([a, b]), there exists h0 ∈ C([a, b]) with kf0 − h0kLp(µ0+µ1) + kf0 − h0kL1([a,b]) < ε . Weierstrass's Theorem provides a polynomial h with kh0 − hkL∞([a,b]) < ε. Let us define the constant c0 := (µ0 + µ1)([a, b])1/p and the polynomial f2(x) :=R x a h. Then, kh0 − hkLp(µ0+µ1) + kh0 − hkL1([a,b]) < ε c0 + ε (b − a) , kf0 − hkLp(µ0+µ1) + kf0 − hkL1([a,b]) < (c0 + b − a + 1) ε , kf1 − f2kLp(µi) ≤ kf0 − hkL1([a,b]) µi([a, b])1/p < c0 (c0 + b − a + 1) ε , and we conclude (cid:12)(cid:12)(cid:12) kgkLp(µiReg([aj−1 ,aj ])) − kf2kLp(µi)(cid:12)(cid:12)(cid:12) 2kLp(µ1)(cid:12)(cid:12)(cid:12) kg′kLp(µ1Reg([aj−1 ,aj ])) − kf ′ (cid:12)(cid:12)(cid:12) ≤ (2 K + c0 (c0 + b − a + 1)) ε , ≤ (2 K + c0 + b − a + 1) ε . Since ε > 0 is arbitrary, this proves (5.13) when Reg([aj−1, aj]) = [aj−1, aj]. If Reg([aj−1, aj]) = (aj−1, aj), we can apply the same argument considering now the functions g1 and g2 in the intervals (aj−1, aj−1 + η) and (aj − η, aj), respectively. In the case Reg([aj−1, aj]) = [aj−1, aj), we consider g1 and g2 in (aj−1 − η, aj−1) and (aj − η, aj), respectively. Finally, for Reg([aj−1, aj]) = (aj−1, aj], we take g1 and g2 in (aj−1, aj−1 + η) and (aj, aj + η), respectively. (cid:3) Theorem 5.9 has the following consequence. Corollary 5.10. Let µ0, µ1 be finite measures on [a, b] such that µ1 is strongly piecewise regular with a0 < a1 < · · · < am. Then the MO is bounded in P1,p(µ0, µ1) iff it is bounded in P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])) for each j ∈ J and µ0({x}) > 0 for all x ∈ H. As we mentioned in the introduction, 5.2 and Corollary 5.10 together characterize the boundedness of the MO for a large class of measures which includes the most usual examples in the literature of orthogonal polynomials. It is remarkable that we require the hypothesis of strongly piecewise regular just for µ1. The following example shows a large class of measures verifying the hypotheses in Corollary 5.10. Example 5.11. The measure µ1 below is finite and strongly piecewise regular r dµ1 := x − a0α0 x − a1α1 · · · x − amαm v(x)I[a0,am](x) dx + cjdδxj , if c1, . . . , cr ≥ 0, x1, . . . , xr ∈ [a0, am], α0, α1, . . . , αm > −1, α0, α1, . . . , αm /∈ [p − 1, p), and there exists a constant C ≥ 1 with C−1 ≤ v(x) ≤ C for x ∈ [a0, am]. If we study a particular (although very large) class of measures, it is possible to improve the first conclusion in Theorem 5.9. Xj=1 MUCKENHOUPT INEQUALITY WITH THREE MEASURES 19 Definition 5.12. Let µ1 be a measure on R. We say that µ1 is piecewise monotone if there exist real numbers b0 < b1 < · · · < bn verifying the following properties: (a) The convex hull of the support of µ1 is the compact interval [b0, bn]. (b) For each 1 ≤ j ≤ n the weight w1 is comparable to a (non-strictly) monotone function on (bj−1, bj). (c) The singular part of µ1 is a finite linear combination of Dirac deltas. If µ1 is piecewise monotone, then it is piecewise regular with constants a0 < a1 < · · · < am (see Definition 5.6). We say that a0 < a1 < · · · < am are the parameters of µ1. Note that if µ1 is piecewise monotone, then b0 = a0 and bn = am, but it is possible that {b0, b1, . . . , bn} 6= {a0, a1, . . . , am}. Also, it is possible that w1 = 0 in some (bj−1, bj). The following results are specially useful in the study of SOP, as Example 5.14 below shows. Theorem 5.13. Let µ0, µ1 be finite measures on [a, b], where µ1 is piecewise monotone with parameters a0 < a1 < · · · < am. If µ0 (Reg([aj−1, aj])) > 0 for each j ∈ J and µ0({x}) > 0 for all x ∈ H, then the MO is bounded in P1,p(µ0, µ1). Proof. By Theorem 5.9, it suffices to prove for every j ∈ J that the MO is bounded in the space P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])). Fixed j ∈ J, since µ1 is piecewise monotone, there exists 0 < ε < (aj −aj−1)/2 such that w1 is comparable to a (non-strictly) monotone function on (aj−1, aj−1 + ε) and on (aj − ε, aj), and (µ1)s ((aj−1, aj−1 + ε)) = (µ1)s ((aj − ε, aj)) = 0. Assume that Reg([aj−1, aj]) = (aj−1, aj), since the other cases are similar and easier. Using that µ0 (Reg([aj−1, aj])) > 0, by Lemma 5.5 the MO is bounded in P1,p(µ0Reg([aj−1,aj ]), µ1Reg([aj−1,aj ])) if we have Λp,[aj−1,x0],aj−1 ((µ1 − k0µ0)+, µ1) < ∞ , Λp,[x0,aj ],aj ((µ1 − k0µ0)+, µ1) < ∞ , for some constant k0 and some point x0 ∈ (aj−1, aj). By Lemma 3.5 these inequalities are equivalent to Λp,[aj−1,aj−1+ε],aj−1 ((µ1 − kµ0)+, µ1) < ∞ , Λp,[aj−ε,aj ],aj ((µ1 − kµ0)+, µ1) < ∞ , for some constant k. Applying Theorem 3.8 we obtain these inequalities, so the proof is finished. (cid:3) The following example shows the large class of measures verifying the hypotheses in Theorem 5.13. Example 5.14. Given a ∈ R, let us consider the set Wa of weights obtained by the products of: α4 α3 x − aα1 , exp(cid:0)−βx − a−α2(cid:1) , , log (cid:12)(cid:12)(cid:12)(cid:12) 1 x − a(cid:12)(cid:12)(cid:12)(cid:12) log (cid:12)(cid:12)(cid:12)(cid:12) log(cid:12)(cid:12)(cid:12)(cid:12) log(cid:12)(cid:12)(cid:12)(cid:12) · · ·(cid:12)(cid:12)(cid:12)(cid:12) 1 x − a(cid:12)(cid:12)(cid:12)(cid:12) · · ·(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) , in such a way that the weights are integrable in some neighborhood of a, and denote by W the class of weights w for which there exist a0 < a1 < · · · < am and weights vj ∈ Waj such that w is comparable to vj in some neighborhood Vj of aj for j = 0, 1, . . . , m, and w is comparable to the constant function 1 in [a0, am]\ ∪m j=0Vj. We say that a0 < a1 < · · · < am are the parameters of w. If dµ1 := w(x)I[a0 ,am](x) dx + cjdδxj , r Xj=1 where w ∈ W with parameters a0 < a1 < · · · < am, c1, . . . , cr ≥ 0, and x1, . . . , xr ∈ [a0, am], then µ1 is finite and piecewise monotone. Note that this class of measures is wider than the one in Example 5.11. As a consequence of Theorem 5.13 we have the following result. Corollary 5.15. Let µ0, µ1 be finite measures on [a, b], where µ1 = µ1,1 + µ1,2, µ1,1 is piecewise monotone with parameters a0 < a1 < · · · < am and µ1,2 ≤ kµ0 for some constant k. If µ0 (Reg([aj−1, aj])) > 0 for each j ∈ J and µ0({x}) > 0 for all x ∈ H, then the MO is bounded in P1,p(µ0, µ1). 20 E. COLORADO, D. PESTANA, J. M. RODR´IGUEZ, AND E. ROMERA References [1] Adams, R.A., Fournier, J.F.: Sobolev spaces. Second edition. Pure and Applied Mathematics (Amsterdam), 140. Else- vier/Academic Press, Amsterdam (2003) [2] Alvarez, V., Pestana, D., Rodr´ıguez, J.M., Romera, E.: Weighted Sobolev spaces on curves, J. Approx. Theory 119, 41-85 (2002) [3] Castro, M., Dur´an, A.J.: Boundedness properties for Sobolev inner products, J. Approx. Theory 122, 97-111 (2003) [4] Fefferman, C.L.: The uncertainty principle, Bull. Amer. Math. Soc. 9, no. 2, 129-206 (1983) [5] Garc´ıa-Cuerva, J., Rubio de Francia, J.L.: Weighted norm inequalities and related topics, volume 116 of North-Holland Mathematics Studies. North-Holland Publishing Col, Amsterdam (1985 [6] Heinonen, J., Kilpelainen, T., Martio, O.: Nonlinear potential theory of degenerate elliptic equations, Oxford Science Publ., Clarendon Press (1993) [7] Heisenberg, W.: On the Perceptual Content of Quantum Theoretical Kinematics and Mechanics. Zeitschrift fur Physik, vol. 43 (1927): 172-198. English Translation by John A. Wheeler and Wojciech Zurek, eds. Quantum Theory and Measurement. Princeton: Princeton University Press (1983): 62-84. [8] Iserles, A., Koch, P.E., Norsett, S.P., Sanz-Serna, J.M.: Orthogonality and approximation in a Sobolev space, in Algorithms for Approximation. J. C. Mason and M. G. Cox, Chapman & Hall, London (1990) [9] Iserles, A., Koch, P.E., Norsett, S.P., Sanz-Serna, J.M.: On polynomials orthogonal with respect to certain Sobolev inner products, J. Approx. Theory 65, 151-175 (1991) [10] Kufner, A.: Weighted Sobolev spaces, Teubner Verlagsgesellschaft, Teubner-Texte zur Mathematik (Band 31), Leipzig (1980). Also published by John Wiley & Sons, New York (1985) [11] Kufner, A., Opic, B.: How to define reasonably weighted Sobolev Spaces, Commentationes Mathematicae Universitatis Caroline 25(3), 537-554 (1984) [12] L´opez Lagomasino, G., P´erez Izquierdo, I., Pijeira, H.: Asymptotic of extremal polynomials in the complex plane, J. Approx. Theory 137, 226-237 (2005) [13] L´opez Lagomasino, G., Pijeira, H.: Zero location and n-th root asymptotics of Sobolev orthogonal polynomials, J. Approx. Theory 99, 30-43 (1999) [14] Mart´ınez-Finkelshtein, A.: Bernstein-Szego's theorem for Sobolev orthogonal polynomials. Constr. Approx. 16, 73-84 (2000) [15] Maz'ja, V.G.: Sobolev spaces, Springer-Verlag, New York (1985) [16] Meyers, N., Serrin, J.: H = W . Proc. Nat. Acad. Sci. USA 51, 1055-1056 (1964) [17] Muckenhoupt, B.: Hardy's inequality with weights. Studia Math. 44, 31-38 (1972) [18] Portilla, A., Quintana, Y., Rodr´ıguez, J.M., Tour´ıs, E.: Zero location and asymptotic behavior for extremal polynomials with non-diagonal Sobolev norms. J. Approx. Theory 162, 2225-2242 (2010) [19] Portilla, A., Rodr´ıguez, J.M., Tour´ıs, E.: The multiplication operator, zero location and asymptotic for non-diagonal Sobolev norms. Acta Appl. Math. 111, 205-218 (2010) [20] Rodr´ıguez, J.M., Alvarez, V., Romera, E., Pestana, D.: Generalized weighted Sobolev spaces and applications to Sobolev orthogonal polynomials I. Acta Appl. Math. 80, 273-308 (2004) [21] Rodr´ıguez, J.M., Alvarez, V., Romera, E., Pestana, D.: Generalized weighted Sobolev spaces and applications to Sobolev orthogonal polynomials II. Approx. Theory Appl. 18:2, 1-32 (2002) [22] Rodr´ıguez, J.M., Alvarez, V., Romera, E., Pestana, D.: Generalized weighted Sobolev spaces and applications to Sobolev orthogonal polynomials: a survey. Electr. Trans. Numer. Anal. 24, 88-93 (2006) [23] Rodr´ıguez, J.M.: The multiplication operator in Sobolev spaces with respect to measures. J. Approx. Theory 109, 157-197 (2001) [24] Rodr´ıguez, J.M.: A simple characterization of weighted Sobolev spaces with bounded multiplication operator. J. Approx. Theory 153, 53-72 (2008) [25] Rodr´ıguez, J.M., Sigarreta, J.M.: Sobolev spaces with respect to measures in curves and zeros of Sobolev orthogonal polynomials. Acta Appl. Math. 104, 325-353 (2008) Departamento de Matem´aticas, Universidad Carlos III de Madrid, Avenida de la Universidad 30, 28911 Legan´es, Madrid, Spain E-mail address: [email protected] Departamento de Matem´aticas, Universidad Carlos III de Madrid, Avenida de la Universidad 30, 28911 Legan´es, Madrid, Spain E-mail address: [email protected] Departamento de Matem´aticas, Universidad Carlos III de Madrid, Avenida de la Universidad 30, 28911 Legan´es, Madrid, Spain E-mail address: [email protected] Departamento de Matem´aticas, Universidad Carlos III de Madrid, Avenida de la Universidad 30, 28911 Legan´es, Madrid, Spain E-mail address: [email protected]
1504.04585
1
1504
2015-04-17T17:26:09
Decomposability of Nonnegative r-Potent Matrices
[ "math.FA" ]
We consider nonnegative r-potent matrices with finite dimensions and study their decomposability. We derive the precise conditions under which an r-potent matrix is decomposable. We further determine a general structure for the r-potent matrices based on their decomposability. Finally, we establish that semigroups of r-potent matrices are also decomposable.
math.FA
math
DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA Abstract. We consider nonnegative r-potent matrices with finite dimensions and study their decomposability. We derive the precise conditions under which an r-potent matrix is de- composable. We further determine a general structure for the r-potent matrices based on their decomposibility. Finally, we establish that semigroups of r-potent matrices are also decom- posable. 1. Introduction A square matrix A is said to be idempotent [1] if and only if A2 = A. The concept of r-potent matrices [2] is a generalization of idempotent matrices where a matrix A is said to be r-potent, for some natural number r, if and only if Ar = A. While every idempotent matrix is r-potent, the reverse is not necessarily true. That is, an r-potent matrix may or may not be idempotent. For example, the matrix (1.1)   0 1 0 0 0 1 1 0 0   is 4-potent (commonly known as quadripotent), but not idempo- tent. Therefore, it makes sense to study r-potent matrices sepa- rately. Several properties of r-potent matrices have been studied by Mc- Closkey [2]. Decomposability of general r-potent matrices has how- ever not been studied so far. On the other hand, decomposability of idempotent matrices has been studied by Marwaha [3]. The tools developed in [3] for idempotent matrices do not apply to r-potent Date: April 13, 2014. 2000 Mathematics Subject Classification. Primary 47D03; Secondary 15B99. Key words and phrases. Decomposition, r-Potent. 1 2 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA matrices. The main focus of this paper is therefore to develop tools so that we can study decomposability of r-potent matrices. Outline of the paper: The rest of the paper is organized as follows. We provide an overview of the known results used in this paper in Section 2. In Section 3, we state our main result on decomposability of r-potent matrices and provide its proof. We provide our results on the structure of r-potent matrices in Section 4. In Section 5, we establish decomposability of the kronecker products of r-potent ma- trices. Section 6 states our main results and corresponding proofs on decomposability of semigroups of r-potent matrices. Finally, we state our results on decomposability of permutation matrices in Section 7. Notation: Capital letters A, B, · · · and small letters a, b, . . . are used to denote matrices and vectors, respectively, over R, where R stands for the space of real numbers. For any set of vectors {v1, v2, . . .}, ∨{v1, v2, . . .} denotes the (closed) linear span of the vectors {v1, v2, . . .}. Mn(R) stands for the space of all n × n ma- trices with entries from R. A matrix A = (aij) ∈ Mn(R) is called nonnegative if aij ≥ 0, ∀i, j = 1, 2, . . . , n. A nonnegative semigroup in Mn(R) is a semigroup of nonnegative matrices. Given two ma- trices A and B, A ⊗ B denotes their kronecker product ([4], pp.3). Since every n × n real matrix represents a linear operator on Rn and vice versa (see [5], pp.276), we use capital letters A, B, · · · to denote finite dimensional linear transformation on Rn and n×n real matrices interchangeably. R(A) and N(A) are used to denote the range space and the null space of linear transformation A. Further, rank(A) denotes the rank of A and Nullity(A) denotes the dimen- sion of the null space of A. Finally, A−1 stands for the inverse of square matrix A. 2. Definitions and Overview of Known Results Definition 2.1. [3] A matrix A ∈ Mn(R) is said to be decompos- able if there exists a proper subset {i1, i2, . . . , ik} of {1, 2, . . . , n} such that ∨{Aei1, Aei2, · · · , Aeik} is contained in ∨{ei1, ei2, . . . , eik}, where {e1, e2, . . . , en} is the standard ordered basis of Rn. The following equivalent definition of decomposability, given as a proposition in [3], will be used throughout this paper: DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 3 Definition 2.2. A matrix A ∈ Mn(R) is decomposable if and only if there exists a permutation matrix P such that (2.1) P−1AP = (cid:20) B C 0 D (cid:21) where B and D are square matrices. A matrix is said to be inde- composable if it is not decomposable. The definition given above for decomposability of a single matrix is extended in the obvious manner to a semigroup in Mn(R) ([8], pp.104), where a common permutation matrix P decomposes every matrix in the semigroup. Definition 2.3. ([6], pp.7) A linear operator A defined on an n- dimensional vector space V is decomposable if there exists a stan- dard subspace (subspace spanned by a subset {e1, e2, . . . , ek} of standard basis vectors e1, e2, . . . , en) M of V such that M is in- variant under the action of A, that is, A(M) ⊆ M. 2.1. Properties of Nonnegative Matrices. Our focus in this paper is on decomposability of nonnegative matrices in Mn(R). For nonnegative matrices, we have the following properties (see [1], pp.487-528, and [7], pp.661-687): (1) For every pair of indices i and j, there exists a natural num- ber m such that (Am)ij is not equal to zero. (2) Period of an index i is the greatest common divisor of all natural numbers m such that (Am)ii > 0. If A is indecom- posable, then period of every index is the same and is called the period of A. (3) Primitive Matrix: A nonnegative matrix A is called prim- itive if its m-th power, Am, is positive for some natural number m (that is, the same m works for all pairs of in- dices). (4) Primitive matrices are the same as indecomposable aperi- odic nonnegative matrices. Please note that, as a consequence of the above statements, every positive matrix is primitive and every primitive matrix is indecom- posable. In other words, every positive matrix is indecomposable. Therefore, we restrict our analysis in this paper to the study of decomposability of nonnegative matrices. 2.2. Perron-Forbenius Theorem for Indecomposable Non- negative Matrices. 4 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA Theorem 2.4. (see [1], pp.487-528, and [7], pp.661-687) Let A be an n × n nonnegative indecomposable matrix with period h and spectral radius ρ. Then the following statements hold: (1) The number ρ is a positive real number, is an eigenvalue of matrix A, and is referred to as the Perron-Frobenius eigen- value of A. (2) The Perron-Frobenius eigenvalue ρ is simple. (3) Matrix A has exactly h complex eigenvalues with absolute value ρ. Each one of these eigenvalues is a simple root of the characterisitic polynomial and is the product of ρ with an h-th root of unity. (4) If A is a nonnegative primitive matrix in Mn(R), then the square matrix An2 −2n+2 is positive. 2.3. Decomposability of Nonnegative Idempotent Matri- ces. We now summarize the results on decomposability of idem- potent matrices [3]: (1) Every nonnegative idempotent matrix of rank k > 1 is de- composable. (2) For an idempotent matrix A, trace(A) = rank(A). (3) Let A be an n × n nonnegative idempotent matrix of rank k > 1. Then, the following hold: (a) Any maximal standard block triangularization of A has the following two properties: (i) Each diagonal block is either zero or a positive idempotent matrix of rank one. (ii) There are exactly k non-zero diagonal blocks. (b) There exists a standard block triangularization of A with the above stated properties 3(a)i and 3(a)ii such that no two consecutive diagonal blocks are zero (so that the total number of diagonal blocks is less than or equal to 2k + 1). (4) Suppose S is a band (semigroup of nonnegative idempotent matrices) in Mn(R) with nonnegative members such that rank(S) > 1, for all S ∈ S. Then, S is decomposable. 2.4. The following lemma by Heydar Radjavi will be repeatedly used for proving our results in this paper: Lemma 2.5. ([8], pp.106) Let A be a nonnegative matrix such that every positive power of A has at least one diagonal entry equal to zero. Then, A is decomposable and has 0 as an eigenvalue. DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 5 2.5. Spectral Properties of r-Potent Matrices. [2] Any r-potent matrix has xr − x as minimal polynomial. Therefore, eigenvalues of an invertible r-potent matrix are the (r − 1)-th roots of unity. r−1 , k = 0, 1, . . . , r − 2, when not counting That is, the roots are e multiplicities of the roots. For a singular r-potent matrix, zero lies in the spectrum apart from the aforesaid eigenvalues. 2kπι Lemma 2.6. [2] For an r-potent matrix A, rank(A) = trace(Ar−1). Proof. We begin by noticing that Ar−1 is an idempotent matrix because A2r−2 = Ar−2Ar = Ar−2A = Ar−1. (2.2) It then follows from Section 2.3 that trace(Ar−1) = rank(Ar−1). In addition, since the range of Ar is contained in the range of Ar−1, we get rank(Ar) ≤ rank(Ar−1) ≤ · · · ≤ rank(A) = rank(Ar). This however implies that rank(Ar) = rank(Ar−1) = rank(A), which, in turn, yields rank(A) = rank(Ar−1) = trace(Ar−1). (cid:3) 3. Decomposability of r-Potent Matrices Before we analyse decomposability of an r-potent matrix, let us It look at decomposability of idempotent matrices more closely. was shown in [3] (restated in Section 2.3 of this paper) that an idempotent matrix of rank > 1 is decomposable. An idempotent matrix of rank one, however, may or may not be decomposable. For example, the idempotent matrix (cid:18) 1/2 1/2 1/2 1/2 (cid:19) is an indecomposable idempotent matrix of rank one, while (cid:18) 1 0 0 0 (cid:19) is a decomposable idempotent matrix of rank one. Let us now ex- tend the result in [3] for idempotent matrices by stating a condition under which an idempotent matrix of rank one becomes decompos- able: Theorem 3.1. Let A be a nonnegative idempotent matrix of rank one. Then, A is decomposable if and only if it has at least one diagonal entry zero. 6 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA Proof. We first assume that A has at least one diagonal entry zero. Then, every positive power of A has at least one diagonal entry equal to zero because An = A, for all natural numbers n. Therefore, the result in [8] (Lemma 2.5 of this paper) implies that A must be decomposable. We next assume that A is decomposable so that (3.1) P−1AP = (cid:18) B C 0 D (cid:19) for some permutation matrix P. In addition, rank(A) = 1 gives rank(P−1AP) = 1, which implies that either B = 0 or D = 0. It therefore follows that A has a zero diagonal entry. (cid:3) Before proceeding to the statement of our main result on decom- posability of nonnegative r-potent matrices in Theorem 3.2, two comments are in order: First, note that an r-potent matrix of rank ≤ r − 1 may or may not be decomposable. For example, if we define A as Ae1 = e2 Ae2 = e3 ... = ... Aer−1 = e1, that is, we consider the matrix 0 0 · · · 1 1 0 · · · 0 0 1 · · · 0 ... ... 0 0 · · · 0 . . . ...   ,   then A is an indecomposable r-potent matrix of rank r − 1 as we cannot find a nontrivial standard invariant subspace. On the other DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 7 hand, the matrix [1] 0 · · · 0 (cid:20) 1/2 1/2 ... 1/2 1/2 (cid:21) · · · . . . ... 0 0 ...   0 0 ... · · · · · · . . . · · · 1/(r − 1) 1/(r − 1) ... 1/(r − 1)     1/(r − 1) 1/(r − 1) ... 1/(r − 1)   is a decomposable r-potent matrix of rank r − 1. Second, it makes sense to analyse the decomposability of r-potent matrices of rank > r − 1 (in addition to the decomposability of r- potent matrices of rank ≤ r − 1) because there exist an infinite number of such r-potent matrices. The existence of such matrices can be established using the properties of kronecker product. In particular, if A and B are two r-potent matrices of rank (say) r − 1 each, then A ⊗ B would also be an r-potent matrix because: (see [4], pp.38) (3.2) (A ⊗ B)r = Ar ⊗ Br = A ⊗ B. Moreover, ([4], pp.20) (3.3) rank(A ⊗ B) = rank(A) · rank(B) = (r − 1)2 > r − 1. Since the above analysis holds for any kronecker product, there exist an infinite number of r-potent matrices of rank > r − 1. To summarize, the above two observations imply that a nonneg- ative r-potent matrix of rank ≤ r − 1 may or may not be decom- posable and there exists an infinite number of r-potent matrices of rank > r − 1. We next state our main result on decomposability of nonnegative real r-potent matrices: Theorem 3.2. For any nonnegative r-potent matrix A, (1) If rank(A) > r − 1, then A is always decomposable. (2) If rank(A) ≤ r − 1 such that A is singular and A2, A3, . . . , Ar−1 have at least one diagonal entry zero, then A is de- composable. Proof. We shall prove the two cases separately: (1) Let us assume that A is indecomposable. Then, the Perron- Frobenius theorem (Theorem 2.4) for indecomposable non- negative matrices is applicable. This implies (substituting 8 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA ρ = 1) that the largest real positive eigenvalue of A (that is, 1) is a simple root of the characteristic polynomial of A. In addition, all other (r − 1)-th roots of unity are simple roots of the characteristic polynomial. Moreover, A has no other eigenvalues. The above statements imply that rank(A) = r − 1, which is a contradiction, and hence, A must be decomposable. (2) Let us again assume that A is indecomposable and consider three jointly exhaustive cases: Case 1: rank(A) = r − 1 Under the assumption of indecomposability, we can ap- ply the Perron-Frobenius theorem (Theorem 2.4) so that 1, α, α2, . . . , αr−2, where α = e r−1 , are all simple roots of the characteristic polynomial of A. Therefore, 2πι (3.4) trace(A) = 1 + α + α2 + · · · + αr−2 = 0. Since A is nonnegative, we must have all the diago- nal entries of A as nonnegative. However, this non- negativity, in conjunction with Eqn. 3.4, implies that all the diagonal entries of A must be zero. Furthermore, since A is an r-potent matrix, this further implies that all the digaonal entries of A = Ar = A2r−1 = A3r−2 = · · · must be zero. Similarly, combining our condition that A2, A3, . . . , Ar−1 have at least one diagonal entry zero with the fact that Ar = A, we get that each of the following: A2 = Ar+1 = A2r = A3r−1 = · · · A3 = Ar+2 = A2r+1 = A3r = · · · ... Ar−1 = A2r−2 = A3r−3 = A4r−4 = · · · have at least one diagonal entry zero. This, thanks to Lemma 2.5, however implies that A must be decom- posable, which is a contradiction. DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 9 Case 2: rank(A) = 1 As A is indecomposable, we can again apply the Per- ron Frobenius theorem (Theorem 2.4) so that 1, be- ing the Perron Frobenius eigenvalue, is a simple eigen- value. Moreover, rank(A) = 1 implies that there is no other eigenvalue with absolute value 1. Furthermore, the number of such eigenvalues is always equal to the period of the matrix. Therefore, we must have A to be aperiodic and hence primitive (Section 2.1). Now, for every primitive matrix A of order n, An2 −2n+2 is a positive matrix (from Theorem 2.4). Finally, since A is singular and rank(A) = 1, the matrix A must be a square matrix of size 2×2 or higher. We can now argue that for n = 2, An2 for n = 3, An2 for n = 4, An2 −2n+2 = A2 is positive, −2n+2 = A5 is positive, −2n+2 = A10 is positive, ... which is a contradiction to the statement of this theo- rem that A2, A3, . . . , Ar−1 have at least one diagonal entry zero. Case 3: 1 < rank(A) < r − 1 Let rank(A) = k, where 1 < k < r − 1. The num- ber of eigenvalues of A would then be equal to the rank(A) = k. However, the assumed indecomposabil- ity of A implies that the number of eigenvalues of A is also equal to the period of A. Hence, the period of A must be equal to k. This is however a contradiction because Ak has at least one diagonal element as zero due to the condition of the theorem, while Ak should be a positive matrix according to Theorem 2.4. Therefore, in each of the above jointly exhaustive cases, we have a contradiction to our assumption that A is indecom- posable. This concludes the proof that A must be decom- posable under the conditions stated in the theorem. (cid:3) 10 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA 4. Structure of an r-Potent Matrix In this section, we study the structure of decomposable nonneg- ative r-potent matrices. Since a decomposable matrix can always be written in a block triangular form (by Defn. 2.2) via a permu- tation matrix, our focus will be on the properties of the diagonal blocks in such a block triangular form. However, before proceeding to our results on the properties of these diagonal blocks, we briefly digress to state a few comments (adopted from [3]) on block tri- angularization that would be useful in proving our results in this section. Let S be a semigroup of matrices in Mn(R) and Lat′S be the lattice of all standard subspaces which are invariant under every member of S. It can be shown by simple induction that for any semigroup S, Lat′S has a maximal chain of such subspaces. This chain may be non-trivial or trivial according to whether S has a non-trivial standard subspace or not. Each nontrivial chain in Lat′S gives rise to a block triangularization for S and since the members in the chain are standard subspaces, we shall call it a standard block triangularization. Evidently, to say that S has a standard block triangularization is equivalent to saying that there exists a permutation matrix P such that for each S ∈ S, P−1SP has the upper block triangular form. Suppose C is a chain in Lat′S and N1 and N2 are two successive elements in C such that N2 ⊆ N1, then N1 ⊖ N2 is called a gap in the chain. If P is the orthogonal projection onto N1 ⊖ N2, then the restriction of PSP to the range of P is called the compression of S to N1⊖N2. Every such compres- sion corresponds to a diagonal block in the block triangularization of S. We now state our main result in this section: Theorem 4.1. Any maximal standard block triangularisation of a decomposable nonnegative r-potent matrix has the following proper- ties: (1) Each diagonal block is either zero or an indecomposable r- potent matrix of rank ≤ r − 1. (2) If n is the number of non-zero diagonal blocks, then (4.1) k r − 1 ≤ n ≤ k, where k = rank(A). DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 11 (3) Total number of diagonal blocks (including 0 blocks) lies be- tween k r−1 and 2k + 1. Proof. We shall prove the above three statements in order: (1) Let (4.2) P−1AP = A11 ∗ 0 A22 ... ... · · · 0 · · · ∗ ... · · · ... . . . · · · Ann     be a maximal standard block triangularization of A via a permutation matrix P. As Ar = A, we have (P−1AP)r = P−1ArP = P−1AP so that Ar ii = Aii, for all i = 1, 2, . . . , n. Thus, each diagonal block is itself an r-potent matrix. Fur- ther, each such diagonal block Aii would be indecomposable and would have rank(Aii) ≤ r − 1. This can be seen as fol- lows: Suppose some Ajj is decomposable with standard sub- space K. Now, Ajj corresponds to some gap, say N1⊖N2, in the maximal chain of invariant subspaces for the aforesaid block triangularization of A. Then, N2 ⊕ K is a standard subspace, invariant under A which lies strictly between N1 and N2, thus contradicting the maximality of the above tri- angularization. In other words, Ajj must be indecompos- able. Finally, since all r-potent matrices of rank > r − 1 are decomposable by Theorem 3.2, we get rank(Aii) ≤ r − 1, ∀i. (2) We start by noticing that (4.3) (4.4) (4.5) (4.6) (4.7) (4.8) (4.9) trace(Ar−1) = trace(Ar−1 11 ) + · · · + trace(Ar−1 nn ). Therefore, k = rank(A) (from Lemma 2.6) 11 ) + · · · + trace(Ar−1 nn ) = trace(Ar−1) = trace(Ar−1 = rank(A11) + · · · + rank(Ann) ≤ (r − 1) + · · · + (r − 1) = n(r − 1), (from Lemma 2.6) 12 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA and therefore, we get the lower bound: n ≥ k r − 1 . On the other hand, let us consider an extreme case that each non-zero diagonal block has rank 1. Then, the maximum number of non-zero diagonal blocks is k, so that we get the upper bound n ≤ k. Combining the two bounds, we can write k r − 1 ≤ n ≤ k. (4.10) (4.11) (3) We claim that two consecutive diagonal blocks cannot be zero. Suppose that two consecutive diagonal blocks are zero. Then, a 2 × 2 block matrix (cid:18) 0 B 0 0 (cid:19) would be an r-potent if and only if it is zero. Therefore, the total number of diagonal blocks (including 0 blocks) lies between k r−1 and 2k + 1. (cid:3) 5. Decomposability of Kronecker Product of r-potent matrices In this section, we shall discuss the decomposability of kronecker products of r-potent matrices. We start by reiterating that the kronecker product of two r-potent matrices is itself an r-potent matrix because (5.1) (5.2) (A ⊗ B)r = Ar ⊗ Br = A ⊗ B. We can now state the two main results in this section as follows: Theorem 5.1. Let A be a nonnegative r-potent matrix of rank > r − 1 and B be any non-zero nonnegative r-potent matrix. Then, A ⊗ B is a decomposable r-potent of rank > r − 1. Proof. Since (5.3) (5.4) rank(A ⊗ B) = rank(A)rank(B) > (r − 1)rank(B), DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 13 which implies that (5.5) rank(A ⊗ B) > r − 1. Therefore, from Theorem 3.2 of this paper, A⊗B is a decomposable r-potent matrix. (cid:3) Theorem 5.2. Let A be a nonnegative r-potent matrix of rank > r − 1 and B be a non-zero nonnegative idempotent matrix. Then, A ⊗ B is a decomposable r-potent matrix. Proof. Since an idempotent matrix is, by definition, also an r- potent matrix, we have (5.6) (A ⊗ B)r = Ar ⊗ Br = A ⊗ B, and therefore, A ⊗ B is also an r-potent matrix. In addition, (5.7) (5.8) rank(A ⊗ B) = rank(A)rank(B) > (r − 1)rank(B), which implies that (5.9) rank(A ⊗ B) > r − 1, and hence, A⊗B is decomposable by Theorem 3.2 of this paper. (cid:3) 6. Decomposibility of Semi-Groups of r-Potent Matrices Let us first analyse the decomposability of a semigroup generated by a single r-potent matrix. Theorem 6.1. Let A be a nonnegative r-potent matrix. Consider the semigroup S = {A, A2, · · · , Ar−1} generated by A. (1) If rank(A) > r − 1, then S is decomposable. (2) If rank(A) ≤ r − 1, A is singular and A2, A3, · · · , Ar−1 have a zero diagonal entry, then S is decomposable. Proof. We shall prove the two parts seperately. (1) Since rank(A) > r − 1, A is decomposable by Theorem 3.2. Let (6.1) (6.2) P−1AP = (cid:18) B C 0 D (cid:19) be a block-triangular decomposition of A. Then (P−1AP)k = P−1AkP = (cid:18) Bk 0 Dk (cid:19) ∗ 14 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA for all k = 2, 3, . . . , r − 1. Therefore, S is decomposable via permutation matrix P. (2) Under the stated conditions, A is decomposable. Decom- posability of S follows from an argument similar to (1) above. (cid:3) The above theorem motivates us to study decomposability of a general semigroup of r-potent matrices. To this end, we require the following equivalent conditions for decomposability of nonnegative semigroups in Mn(R). Lemma 6.2. [3] For a semigroup S in Mn(R) with nonnegative matrices, the following are equivalent: (1) S is decomposable. (2) There exists a non-zero nonnegative functional on Mn(R) whose restriction to S is zero. (3) S has a common zero entry. (4) S has a common non-diagonal zero entry. (5) There exist A, B ∈ Mn(R), both non-zero and nonnegative such that ASB = {0}. We now proceed to the two main results of this section and detail their respective proofs: Theorem 6.3. Let S be a semigroup of nonnegative r-potent ma- trices of rank > r − 1. Then, S is decomposable. Proof. We start by noticing that a semigroup of r-potent matrices always contains an idempotent matrix. For example, if A is an r-potent matrix in S, then Ar−1 is an idempotent matrix in S (see Eqn. 2.2). Consider now a minimal rank idempotent matrix P in S. We can always choose such a minimal rank idempotent matrix because if we choose any minimal rank matrix B in S, then the rank of the corresponding idempotent matrix Br−1 is upperbounded by rank(B) because ([4], pp.61) (6.3) (6.4) rank(Br−1) ≤ min{rank(Br−2), rank(B)} ≤ rank(B). Now, since rank(P) > r − 1 > 1, P must be decomposable. Let P have the form (cid:18) P1 K 0 P2 (cid:19) DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 15 with respect to some permutation of basis where both P1 and P2 are non-zero. Let us now consider an arbitrary S ∈ S. Then, (PSP)r−1 will be an idempotent matrix, also in S. Further, the range of (PSP)r−1 is contained in the range of P and the null space of (PSP)r−1 contains the null space of P. Therefore, (6.5) rank((PSP)r−1) ≤ rank(P). Considering that P is minimal rank in S, we get (6.6) rank((PSP)r−1) = rank(P). Using rank-nullity theorem of linear algebra ([7], pp.199), we have nullity of (PSP)r−1 is equal to the nullity of P, which, in turn, yields (6.7) (PSP)r−1 = P. This, however, implies that (6.8) PSP = P 1 r−1 = P due to the following lemma: Lemma 6.4. In the given semigroup S, the only nonnegative (r − 1)-th root of minimal rank idempotent P is P itself. Proof. Since (6.9) (6.10) P2 = P ⇒ Pr−1 = P, P is itself a nonnegative (r−1)-th root of P in S. Let P′ be another nonnegative (r − 1)-th root of P in S, that is, (6.11) (P′)r−1 = P. As P′ belongs to S, we also have (6.12) (P′)r = P′. It follows from Eqn. 6.11 and Eqn. 6.12 that (6.13) PP′ = P′P = P′. Using the fact that P is of minimal rank in S, it follows from rank- nullity theorem that R(P) = R(P′) and N(P) = N(P′). Moreover, Eqn. 6.9 implies that (6.14) P = I on R(P). 16 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA which, due to Eqn. 6.11, implies that (6.15) (P′)r−1 = I on R(P) where I is the identity operator. In other words, P′ is a nonnegative (r − 1)-th root of I on R(P). However, the only nonnegative (r − 1)-th root of the identity operator is the identity operator itself. Therefore (6.16) and (6.17) Therefore (6.18) Finally, let P′ = I = P on R(P) = R(P′) P′ = P = 0 on N(P) = N(P′). P′ = P. (cid:18) S11 S12 S21 S22 (cid:19) (cid:3) be the representation of an arbitrary S ∈ S with respect to this permuted basis. Then, PSP = P implies that P2S21P1 = 0. If pij and rkl are non-zero entries in P2 and P1, respectively, then it is easy to see that the (j, k)-th entry in each S ∈ S is zero. This makes use of the fact that P2, S21, and P1 are all nonnegative matrices. By Lemma 6.2, S is decomposable. (cid:3) Theorem 6.5. Let S be a semigroup of nonnegative r-potent ma- trices of rank > r − 1. Then, any maximal standard block triangu- larization of S has the property that each non-zero diagonal block is a semigroup of nonnegative r-potent matrices with at least one element of rank ≤ r − 1. Proof. By Theorem 6.3, S is decomposabable. Consider any maxi- mal chain in Lat′S resulting in a standard block triangularization of S. Consider any two subspaces N1 and N2 in this chain such that N1 ⊆ N2 and N2⊖N1 is a gap. If the compression of S to N2⊖N1 is non-zero, it forms a semigroup of nonnegative r-potents. Further, it must be indecomposable, for otherwise, if it has a standard invari- ant subspace K, then N1 ⊕ K is in Lat′S and lies strictly between N1 and N2, contradicting the maximality of this chain. Thus, every DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 17 non-zero compression (or diagonal-block) constitutes an indecom- posable semigroup of r-potent matrices. By Theorem 6.3, it must contain at least one element of rank ≤ r − 1. (cid:3) 7. Decomposability of permutation matrices In this section, we shall study decomposability (or rather, inde- composability) of permutation matrices. We start by noticing that an n×n circulant matrix generated by e1, e2, . . . , en (standard basis vectors), given by 0 0 0 0 · · · 0 1 0 0 · · · 0 0 1 0 · · · ... . . . 1 0 0 0 · · · ... ... ...   1 0 0 ... 0  n×n , is indecomposable as we cannot find a standard invariant subspace. To generalize this observation as a formal result, we require the following lemma: Lemma 7.1. ([8], pp.106) For a semigroup S of nonnegative ma- trices, the following are equivalent: (1) S is decomposable. (2) Every sum of members of S has a zero entry. We now state the main result of this section. Theorem 7.2. Group G of n × n permutation matrices generated by the standard basis vectors e1, e2, . . . , en, comprises of idempo- tent matrices, tripotent matrices, quadripotent matrices, pentapo- tent matrices,. . . , (n + 1)-potent matrices and is indecomposable. Proof. The group comprises the following: Idempotent Matrices: In×n =   1 0 0 0 · · · 0 1 0 0 · · · 0 0 1 0 · · · ... . . . 0 0 0 0 · · · ... ... ... 0 0 0 ... 1  n×n 18 RASHMI SEHGAL THUKRAL AND DR. ALKA MARWAHA Tripotent Matrices: are formed by permuting two rows at a time. For example A21 =   0 1 0 0 · · · 1 0 0 0 · · · 0 0 1 0 · · · ... . . . 0 0 0 0 · · · ... ... ... 0 0 0 ... 1  n×n Quadripotent Matrices: are formed by permuting three rows at a time. For example A231 =   0 1 0 0 · · · 0 0 1 0 · · · 1 0 0 0 · · · 0 0 0 1 · · · ... . . . 0 0 0 0 · · · ... ... ... 0 0 0 0 ... 1  n×n (n+1)-potent Matrices: are formed by permuting all n rows at a time. For example A234...(n+1)1 =   0 1 0 0 · · · 0 0 1 0 · · · ... . . . 0 0 0 0 · · · 1 0 0 0 · · · ... ... ... 0 0 ... 1 0  . n×n Then, In×n+P A234...(n+1)1 does not have any zero entry. Therefore, by Lemma 7.1, G is indecomposable. (cid:3) References 1. Roger A. Horn and Charles R. Johnson, Matrix Analysis, Cambridge Uni- versity Press, ISBN 0521386322, (1990). 2. Joseph P. McCloskey, Characterizations of r-Potent Matrices, Mathemat- ical Proceedings of the Cambridge Philosophical Society, 96, (1984), 213– 222. 3. Alka Marwaha and Heydar Radjavi, Decomposability and Structure of nonnegative Bands in Mn(R), Linear Algebra and its Applications, 291, (1999), 63-82. 4. Helmut Lutkepohl, Handbook of Matrices, Wiley, New York, ISBN 0-471- 96688-6, (1996). 5. I. N. Herstein, Topics in Algebra: 2nd Edition, John Wiley and Sons (Asia), Singapore, ISBN-13: 9971-51-353-X, (1975). DECOMPOSABILITY OF NONNEGATIVE r-POTENT MATRICES 19 6. Alka Marwaha, Decomposability and Structure of Bands of Nonnegative Operators,PhD Thesis, Dalhausie University, Canada, (1996). 7. Carl Meyer, Matrix Analysis and Applied Algebra, Society for Industrial and Applied Mathematics, Philadelphia, PA, USA, ISBN: 0898714540, (2000). 8. Heydar Radjavi and Peter Rosenthal, Simultaneous Triangularization, Springer-Verlag, ISBN: 0387984674, Canada (2000). Department of Mathematics, Jesus and Mary College, Univer- sity of Delhi, Chanakyapuri, New Delhi 110056 E-mail address: [email protected] Department of Mathematics, Jesus and Mary College, Univer- sity of Delhi, Chanakyapuri, New Delhi 110056 E-mail address: [email protected]
1601.02037
1
1601
2016-01-08T21:53:42
Nonseparable $C(K)$-spaces can be twisted when $K$ is a finite height compact
[ "math.FA" ]
We show that, given a nonmetrizable compact space $K$ having $\omega$-derived set empty, there always exist nontrivial exact sequences $0\to c_0\to E\to C(K)\to 0$. This partially solves a problem posed in several papers: Is $Ext(C(K), c_0)\neq 0$ for $K$ a nonmetrizable compact set?
math.FA
math
NONSEPARABLE C(K)-SPACES CAN BE TWISTED WHEN K IS A FINITE HEIGHT COMPACT JES ´US M. F. CASTILLO Abstract. We show that, given a nonmetrizable compact space K having ω-derived set empty, there always exist nontrivial exact sequences 0 → c0 → E → C(K) → 0. This partially solves a problem posed in several papers: Is Ext(C(K), c0) 6= 0 for K a nonmetrizable compact set? 1. Introduction A general problem in Banach space theory is to determine, given two Banach spaces Y and Z, the existence and properties of Banach spaces X containing Y such that X/Y = Z. The space X is called a twisted sum of Y and Z. When only the trivial situation X = Y ⊕ Z is possible one obtains an interesting structure result asserting that every copy of Y in a space X such that X/Y = Z must be complemented. When some X 6= Y ⊕ Z as above exists, one obtains a usually exotic and interesting space with unexpected properties; perhaps the perfect example could be the Kalton-Peck Z2 space [20] which has no unconditional basis while containing an uncomplemented copy of ℓ2 such that Z2/ℓ2 = ℓ2. Recall that a short exact sequence of Banach spaces is a diagram (1) 0 −−−−→ Y i−−−−→ X q −−−−→ Z −−−−→ 0 where Y , X and Z are Banach spaces and the arrows are operators in such a way that the kernel of each arrow coincides with the image of the preceding one. By the open mapping theorem i embeds Y as a closed subspace of X and Z is isomorphic to the quotient X/i(Y ). Two exact sequences 0 → Y → X → Z → 0 and 0 → Y → X1 → Z → 0 are said to be equivalent if there exists an operator T : X −→ X1 making commutative the diagram 0 −−−−→ Y −−−−→ X −−−−→ Z −−−−→ 0 k k Ty 0 −−−−→ Y −−−−→ X1 −−−−→ Z −−−−→ 0. The sequence (1) is said to be trivial (or that it splits) if i(Y ) is complemented in X (i.e., if it is equivalent to the sequence 0 → Y → Y ⊕ Z → Z → 0). We write Ext(Z, Y ) = 0 to indicate that every sequence of the form (1) splits. Summing all up, the general problem is to determine when there exist nontrivial twisted sums of Y and Z, or else, when Ext(Z, Y ) 6= 0. We pass now to consider the specific case in which both Y and Z are spaces of continuous functions on compact spaces. The twisting of separable C(K)-spaces was treated and, to a large extent, solved in [7]. Nevertheless, the problem of constructing nontrivial twisted sums of large C(K)-spaces is mostly unsolved. The following problem was posed and considered in [7, 8, 10]: Is it true that for every non- metrizable compact K one has Ext(C(K), c0) 6= 0? In this paper we prove that Ext(C(K), C(S)) 6= 0 Thanks are due to a demanding referee who pushed the author to look for nicer arguments and, in the end, to produce a better paper. Additional thanks are due to Daniel Tausk who made a few accurate remarks about the content of the paper. This research was supported by project MTM2013-45643-C2-1-P, Spain. 1 2 JES ´US M. F. CASTILLO for any two compact spaces K, S with ω-derived set empty and K nonmetrizable which, in particular, covers the case above for C(S) = c0. 2. Main result, and its consequences In what follows, the cardinal of a set Γ will be denoted Γ. We will write c0(Γ) or c0(Γ) depending if one needs to stress the set or only the cardinal. We begin with a couple of observations: First of all, that if Γ is an uncountable set then Ext(c0(Γ), c0) 6= 0. To explicitly construct an example it is enough to assume that Γ = ℵ1, pick an isomorphic embedding ϕ : c0(ℵ1) → ℓ∞/c0 and consider the commutative diagram 0 −−−−→ c0 −−−−→ ℓ∞ p −−−−→ ℓ∞/c0 −−−−→ 0 (cid:13)(cid:13)(cid:13) 0 −−−−→ c0 −−−−→ p−1 (ϕ(c0(ℵ1))) ϕ x pϕ−−−−→ c0(ℵ1) −−−−→ 0 x ϕ′ x where pϕ(x) = ϕ−1p(x). Indeed, to see that the lower sequence does not split, one only has to observe that in any commutative diagram 0 −−−−→ c0 −−−−→ ℓ∞ −−−−→ ℓ∞/c0 −−−−→ 0 (2) (cid:13)(cid:13)(cid:13) ϕ x 0 −−−−→ c0 −−−−→ X −−−−→ c0(Γ) −−−−→ 0. q in which ϕ is an isomorphic embedding then also ϕ′ is an isomorphic embedding, which implies that the quotient map q cannot be an isomorphism onto any non separable subspace of c0(Γ) since non- separable c0(Γ′) spaces are not subspaces of ℓ∞. The lower sequence in diagram (2) can however split when ϕ is just an operator. Moreover, the results in [12] imply that in any exact sequence 0 −−−−→ c0 −−−−→ X q −−−−→ c0(J) −−−−→ 0 in which J large enough the operator q becomes an isomorphism on copies of c0(I) for large I; while, on the other hand, in an exact sequence 0 −−−−→ Y −−−−→ X q −−−−→ c0(J) −−−−→ 0, in which Ext(X, c0) = 0 then q cannot be an isomorphism onto any non-separable subspace of c0(J): otherwise, if q becomes an isomorphism on some subspace c0(J ′), this must be complemented in c0(J) by [17], and therefore q−1(c0(J ′)) must be complemented in X, which prevents Ext(X, c0) = 0. It is still open to decide whether every twisted sum of two c0(Γ) must be a C(K)-space (see [13] for related information). We prove two preparatory lemmata of independent interest. The first one is a reformulation of [11, Lemma 1]. Recall that given any exact sequence 0 → A → B → B/A → 0 and an operator φ : A → Y there exists a superspace X of Y such that X/Y = B/A (see [9, Prop. 1.3.a]), thus forming a commutative diagram 0 −−−−→ A −−−−→ B −−−−→ B/A −−−−→ 0 (3) φy y (cid:13)(cid:13)(cid:13) 0 −−−−→ Y −−−−→ X −−−−→ B/A −−−−→ 0. The following lemma shows that the converse is somehow true: NONSEPARABLE C(K)-SPACES CAN BE TWISTED WHEN K IS A FINITE HEIGHT COMPACT 3 Lemma 1. Let 0 → A → B → C → 0 be an exact sequence and let E be a Banach space. If Ext(B, E) = 0 then for every exact sequence 0 → E → X → C → 0 there is an operator φ : A → E so that there is a commutative diagram 0 −−−−→ A −−−−→ B −−−−→ C −−−−→ 0 (4) 0 −−−−→ E −−−−→ X −−−−→ C −−−−→ 0. (cid:13)(cid:13)(cid:13) φy y This result can be seen as a consequence of the homology sequence (see [5]), whose part relevant for us is the existence of an exact sequence: . . . −−−−→ L(A, E) ωE−−−−→ Ext(C, E) −−−−→ Ext(B, E) −−−−→ · · · . Thus, if Ext(B, E) = 0 then the map ωE : L(A, E) → Ext(C, E) is surjective. This map ωE sends operators φ : A → E into the lower sequence in diagram (3). Lemma 2. Let (5) 0 −−−−→ Y −−−−→ X −−−−→ c0(c) −−−−→ 0 be an exact sequence. If Y ∗ ≤ c then Ext(X, c0) 6= 0. Proof. Applying the homology sequence as above to sequence (5) with target E = c0 one gets that the map ωY : L(Y, c0) −−−−→ Ext(c0(c), c0) is surjective when Ext(X, c0) = 0, from where Ext(c0(c), c0) ≤ L(Y, c0). Let us show that, under the assumption Y ∗ ≤ c, this cannot be so: On one hand one has L(Y, c0) ≤ L(ℓ1, Y ∗), which is the set of bounded sequences of the unit ball of Y ∗. Since there are (Y ∗)N countable subsets of Y ∗, each of them admitting c bounded sequences one gets, L(Y, c0) ≤ R × (Y ∗ℵ0)ℵ0 = cℵ0 = c. On the other hand, Marciszewski and Pol show in [21, 7.4] that there exist 2c non-equivalent exact sequences 0 → c0 → ♠ → c0(c) → 0; i.e., Ext(c0(c), c0) ≥ 2c. (cid:3) The reason behind the involved proof to come is that the straightforward proof one would obtain from the argument above does not holds in general since the inequality ℵℵ0 < 2ℵ does not necessarily hold for all cardinals. Indeed, see [18, Thm. 5.15], under GCH, it occurs that when ℵ has cofinality greater than ℵ0 then ℵℵ0 = ℵ; but for ℵ having cofinality ℵ0 then ℵℵ0 = 2ℵ. Moreover, the Continuum Hipothesis, in what follows [CH], is necessary to make c the first case to consider. Indeed, the argument of Marciscewski-Pol could not work even for ℵ1 outside CH since it depends on cardinal arithmetics: for c < 2ℵ1 the same proof in [21] yields that there are 2ℵ1 different exact sequences 0 → c0 → X → c0(ℵ1) → 0. However, if 2ℵ1 = c then the method in [21] does not decide. We are indebted to W. Marciszewki who informed us from these facts. Lemma 2 can be improved to: Lemma 3. Let 0 → Y → X → Z → 0 be an exact sequence. Let B be a Banach space for which L(Y, B) < Ext(Z, B). Then Ext(X, B) ≥ Ext(Z, B). Proof. It follows from the exactness of the sequence L(Y, B) ωB−−−−→ Ext(Z, B) γ −−−−→ Ext(X, B) 4 JES ´US M. F. CASTILLO and the standard factorization Ext(Z, B)/ ker γ = Imγ that Ext(X, B) ≥ Im γ = Ext(Z, B)/ ker γ = Ext(Z, B)/ Im ωB. On the other hand one has Im ωB ≤ L(Y, B) < Ext(Z, B), and thus Ext(Z, B)/ Im ωB = Ext(Z, B). (cid:3) To continue with the preparation for the proof we need now a result that can be considered folklore but which, to the best of our knowledge, cannot be found in the literature: Lemma 4. Let Y be a subspace of X. There is a subspace Y • ⊂ Y with density character dens Y • ≤ dens X/Y and a subspace X • ⊂ X such that X •/Y • = X/Y . Proof. Let (zγ)γ be a dense subset of the unit ball of X/Y having minimal size. By the open mapping theorem there is some constant C > 0 such that one can pick for each zγ an element xγ ∈ X with norm kxγk ≤ Ckzγk so that xγ + Y = zγ. Let L : X/Y → X be a linear (not necessarily continuous) selection for the quotient map q : X → X/Y . Thus xγ − Lzγ ∈ Y and the closed subspace they span Y • = [xγ − Lzγ] has density character smaller or equal than that of Z. Let X • = [Y • + [xγ]] be the closed subspace of X spanned by Y • and all the xγ. It is clear that q : X • → X/Y is onto since the points xγ + Y = zγ are in the image of the ball of radius C of X •. Moreover, ker qX • = Y •: indeed, since X • = [Y • + [xγ]] = [xγ − Lzγ, xγ] = [xγ, Lzγ], pick x ∈ X • of the form x = P λγxγ + P λµLzµ so that qx = 0; which means P λγzγ = −P λµzµ. Thus x = X λγxγ + X λµLzµ = X λγxγ + L(cid:16)X λµzµ(cid:17) = X λγxγ − L(cid:16)X λγzγ(cid:17) = X λγ(xγ − Lzγ) and thus x ∈ Y •. (cid:3) Lemma 4 can be reformulated as: given an exact sequence 0 → Y → X → Z → 0 there are subspaces Y • ⊂ Y and X • ⊂ X with dens Y • ≤ dens Z that form a commutative diagram 0 −−−−→ Y • −−−−→ X • −−−−→ Z −−−−→ 0 y iy y i• y (cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) 0 −−−−→ Y −−−−→ X −−−−→ Z −−−−→ 0. In particular, given an exact sequence 0 → c0(Γ) → X → c0(c) → 0, since for every subspace E ⊂ c0(Γ) with dens E ≤ ℵ ≤ Γ there is a copy of c0(ℵ) for which E ⊂ c0(ℵ) ⊂ c0(Γ), it is clear that there is a commutative diagram 0 −−−−→ c0(c) −−−−→ X • −−−−→ c0(c) −−−−→ 0 0 −−−−→ c0(Γ) −−−−→ X −−−−→ c0(c) −−−−→ 0, where i, i• are the canonical inclusions. We are ready for the action: Basic case: J = c: Lemma 5. Given an exact sequence 0 → c0(Γ) → X → c0(c) → 0 then Ext(X, c0) 6= 0. Proof. If Ext(X, c0) = 0 then the map ωc0 : L(c0(Γ), c0) → Ext(c0(c), c0) is surjective and thus, via composition with i, there is also a surjective map L(c0(c), c0) → Ext(c0(c), c0), which is impossible. (cid:3) Reduction to the case J = c: Lemma 6. Under CH, given an exact sequence 0 → c0(Γ) → X → c0(J) → 0 then Ext(X, c0) 6= 0 when X is nonseparable. NONSEPARABLE C(K)-SPACES CAN BE TWISTED WHEN K IS A FINITE HEIGHT COMPACT 5 Proof. If J ≤ ℵ0 then the sequence splits (the spaces c0(Γ) are separably injective, which means that they are complemented in every superspace X so that X/c0(Γ) is separable (see, e.g., [3]). In which case, X is isomorphic to c0(Γ). Being X nonseparable, Γ must be uncountable and Ext(X, c0) 6= 0. If J = c, we are in the case of Lemma 5. If J > c, take a set of size c in J and consider the decomposition c0(J) = c0(c) ⊕ c0(H) with canonical embedding u : c0(c) → c0(J). Take an arbitrary exact sequence 0 → c0 → A → c0(c) → 0 and multiply on the right by c0(H) to get the exact sequence 0 → c0 → A ⊕ c0(H) → c0(c) ⊕ c0(H) → 0. Form then the commutative diagram 0 −−−−→ c0(c) −−−−→ q−1(u(c0(c)))• −−−−→ c0(c) −−−−→ 0 0 −−−−→ c0(Γ) −−−−→ q−1(u(c0(c))) 0 −−−−→ c0(Γ) −−−−→ i• u′ φ′ y y y X iy (cid:13)(cid:13)(cid:13) φy qu−−−−→ q −−−−→ −−−−→ 0 −−−−→ 0 c0(c u c0(J) (cid:13)(cid:13)(cid:13) y (cid:13)(cid:13)(cid:13) 0 −−−−→ c0 −−−−→ A ⊕ c0(H) −−−−→ c0(c) ⊕ c0(H) −−−−→ 0. where the operator φ exist because Ext(X, c0) = 0. If pA : A ⊕ c0(H) → A denotes the canonical projection onto A then there is a commutative diagram 0 −−−−→ c0(c) −−−−→ q−1(u(c0(c)))• −−−−→ c0(c) −−−−→ 0 φiy 0 −−−−→ c0 −−−−→ pAφ′u′i• y A (cid:13)(cid:13)(cid:13) −−−−→ c0(c) −−−−→ 0. This again implies that there is a surjective map L(c0(c), c0) → Ext(c0(c), c0), which we have already shown it is impossible. (cid:3) We are now ready to show: Proposition 1. Let Xn be a sequence of Banach spaces in which X0 is a nonseparable twisted sum of c0(Γ0) and c0(J0), and Xn is a twisted sum of c0(Γn) and Xn−1 for all n ≥ 1. Under CH, Ext(Xn, c0) 6= 0 for all n ≥ 0. Proof. We proceed (formally) inductively on n. That Ext(X0, c0) 6= 0 has been already shown. We show now that also Ext(X1, c0) 6= 0. By the arguments preceding, there is no loss of generality 0) → X0 → c0(c) → assuming that J0 ≥ c, in which case there is also an exact sequence 0 → c0(Γ′ 6 JES ´US M. F. CASTILLO and then X1 and X0 are connected as in the commutative diagram 0 0 c0(c) = c0(c) (6) 0 −−−−→ c0(Γ1) −−−−→ X1 q −−−−→ X0 −−−−→ 0 0 −−−−→ c0(Γ1) −−−−→ q−1(u(c0(Γ′ 0))) −−−−→ c0(Γ′ 0) −−−−→ 0 qu x x x x 0 x x x x 0 p u x x x x 0. p u x x x x 0. (cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) Now, reasoning as in Lemma 5 and Lemma 6 we can assume that Γ′ 0 = c, and thus that also Γ1 ≤ c; in which case q−1(u(c0(Γ′ 0)))∗ ≤ c and Lemma 2 applies to yield Ext(X1, c0) 6= 0. The result for X2 can be obtained via the same method: call P1 = q−1(u(c0(Γ′ 0))) from diagram (6). We get the diagram 0 0 (7) 0 −−−−→ c0(Γ2) −−−−→ X2 −−−−→ X1 −−−−→ 0 0 −−−−→ c0(Γ2) −−−−→ q−1(u(P1)) −−−−→ P1 −−−−→ 0 c0(c) = c0(c) q qu After the appropriate size reduction we can assume that P ∗ 1 ≤ c and Γ2 ≤ c which implies that q−1(u(c0(Γ1)))∗ ≤ c and thus Lemma applies to yield Ext(X2, c0) 6= 0. Call now Pn+1 = q−1(u(Pn)) for n = 1, 2, . . . and iterate the argument. (cid:3) Let us recall that given a compact set K, the derived set K (1) is the set of its accumulation points. If one calls K (n+1) = (K (n))′, the ω-derived set is K ω = ∩nK (n). Theorem 1. [CH] Given two compact spaces K, S with ω-derived set empty and K nonmetrizable then Ext(C(K), C(S)) 6= 0. Proof. The result is clear after recalling a few well-known facts: • That C(S) spaces in which S is a compact with Sω = ∅ contain c0 complemented as it follows from the existence of a convergent sequence in S. NONSEPARABLE C(K)-SPACES CAN BE TWISTED WHEN K IS A FINITE HEIGHT COMPACT 7 • If IK is the set of isolated points of K the restriction operator C(K) → C(K (1)) induces a short exact sequence 0 −−−−→ c0(IK ) −−−−→ C(K) −−−−→ C(K (1)) −−−−→ 0. • That, after a finite number N +2 of derivations of K, one arrives to a finite compact, obtaining thus an exact sequence 0 −−−−→ c0(IK (N ) ) −−−−→ C(K (N )) −−−−→ c0(J) −−−−→ 0. We are therefore in the hypotheses of Proposition 1 and thus Ext(C(K), c0) 6= 0 and therefore Ext(C(K), C(S)) 6= 0. (cid:3) Godefroy, Kalton and Lancien show in [16, Thm. 2.2] that a C(K) space is Lipschitz isomorphic to c0(N) if and only if it is linearly isomorphic to c0. They also show that the result is no longer true in the nonseparable case, showing that a C(K) space is Lipschitz isomorphic to some c0(I) if and only if the ω derived set of K is empty. Therefore, Corollary 1. [CH]. If C(K) and C(S) are Lipschitz isomorphic to some (probably different) spaces c0(Γ) then Ext(C(K), C(S)) 6= 0. For the sake of completeness, let us gather in an omnibus theorem all known results about the problem considered. Recall that a compact space K is said to have countable chain condition (ccc in short) when every family of pairwise disjoint open sets is countable. It is well known [25, Thm. 4.5] that K has ccc if and only if C(K) does not contain c0(I) for I uncountable. Proposition 2. One has Ext(C(S), c0) 6= 0 whenever C(S) contains a complemented non-separable C(K) space for which at least one of the following conditions hold: (1) K is an Eberlein compact. (2) K is a Valdivia compact without ccc. (3) [CH] K has finite height. (4) C(K) admits a continuous injection into C(N∗) but not into ℓ∞. (5) K is an ordinal space. (6) C(K) contains ℓ∞. Proof. The proof of (1) appears in [10]: According to Reif [24], since the space C(K) is WCG it admits a Markusevic basis (xγ, fγ)γ ∈ Γ, for which it can be assumed that (fγ) is bounded. It is possible to define a dense-range operator T : C(K) → c0(Γ) as T (f ) = (fγ(f )). On the other hand, q in any nontrivial exact sequences 0 → c0 → X → c0(Γ) → 0 the space X cannot be WCG (Weakly Compactly Generated) since c0 is complemented in WCG spaces. Thus, one has a commutative diagram 0 −−−−→ c0 −−−−→ X −−−−→ c0(Γ) −−−−→ 0 (cid:13)(cid:13)(cid:13) tx q π T x 0 −−−−→ c0 −−−−→ PT −−−−→ C(K) −−−−→ 0 where PT = {(f, g) ∈ X ⊕∞ C(K) : q(f ) = T (g)}, t(f, g) = f and π(f, g) = g (indeed, when T is not an embedding one cannot expect the middle operator t to be an embedding; i.e., one cannot expect the lower twisted sum space to be a subspace of the upper twisted sum space). The "dense range" version of the 3-lemma implies that the operator t has dense range; hence, the lower sequence cannot split since otherwise PT ≃ c0 ⊕ C(K) would be WCG, as well as X. (2) follows from [1, Thm. 1.2]: if K is a Valdivia compact and c0(I) ⊂ C(K) then there is a subset J ⊂ I such that J = I and c0(J) is complemented in C(K). 8 JES ´US M. F. CASTILLO Assertion (3) is contained in the main results of the paper. Assertion (4) follows from the commutative diagram 0 −−−−→ c0 −−−−→ ℓ∞ q −−−−→ C(N∗) −−−−→ 0 (cid:13)(cid:13)(cid:13) tx T x qT−−−−→ C(K) −−−−→ 0 0 −−−−→ c0 −−−−→ q−1(ϕ(C(K))) in which T is the continuous injection claimed in he hypothesis. If the lower sequence splits, tC(K) would be a continuous injection C(K) → ℓ∞. Instances of the situation just described appear when, say, K contains a dense set of weight at most ℵ1 but C(K) is not a subspace of ℓ∞, as it follows from Parovicenko's theorem [4], which implies the existence of a continuous surjection ϕ : N∗ → K, which in turn provides the injection ϕ◦ : C(K) → C(N∗). See also [27, Prop. 0.1]. Or else, C(K) spaces with non weak*-separable dual, but admitting continuous injections into C(N∗). (5) The case of uncountable ordinal spaces can be reduced to K = [0, ω1], and this is contained in (4). (6) is obvious since Ext(ℓ∞, c0) 6= 0 (see [6]) and ℓ∞ will be complemented in C(K). (cid:3) In [7, p. 4539-4540] we claimed that the argument in (1) can be extended to cover the case of Corson compacta. We have been informed by Daniel Tausk that such result does not immediately follows with the same arguments: on one side, consider any nontrivial exact sequence 0 → c0 → X → c0(I) → 0. This prevents BX ∗ from being a Valdivia compact. If we assume K is a Corson compact then from [2, Thm. 1.7] it follows that there is an injective operator T : C(K) → c0(I), or a dense-range operator T : C(K) → c0(I) if one uses either Markushevich basis for C(K) or increasing PRI [26], argument also valid for K Valdivia. Thus, one gets a commutative diagram 0 −−−−→ c0 −−−−→ X q −−−−→ c0(Γ) −−−−→ 0 (cid:13)(cid:13)(cid:13) tx T x 0 −−−−→ c0 −−−−→ PT −−−−→ C(K) −−−−→ 0. But some argument is required now to conclude that no continuous lifting C(K) → X exists in order to make the lower sequence nontrivial. If T has been chosen having dense range, then t∗ : X ∗ → C(K)∗ is an imbedding; thus, if BC(K)∗ is a Corson compact then also BX ∗ would be a Corson compact, which is a contradiction. However, it is consistent with ZFC the existence of Corson compacta K so that the unit ball of C(K)∗ is not Corson (see [22, Thm. 5.9], or else [2, Thm.3.5, Ex.3.10]). The situation for Valdivia compacta is not much better: while now K Valdivia implies that BC(K)∗ is Valdivia [19, Thm. 5.2], it is no longer true that closed subspaces of Valdivia compacta are Valdivia. Therefore, it is worth to state the problem: Problem. Assume that K is a (nonmetrizable) Corson or Valdivia compact. Does one have Ext(C(K), c0) 6= 0? Added in proof. In the meantime Correa and Tausk have shown in [15] that under CH all Corson compact admit a nontrivial twisting with c0 and that the same holds for "most" Valvidia compact. [1] S. Argyros, J.M.F. Castillo, A.S. Granero, M. Jimenez and J.P. Moreno, Complementation and embeddings of c0(I) in Banach spaces. Proc. London Math. Soc. 85 (2002) 742 -- 772. References NONSEPARABLE C(K)-SPACES CAN BE TWISTED WHEN K IS A FINITE HEIGHT COMPACT 9 [2] S. Argyros, S. Merkouriakis and S. Negrepontis, Functional analytic properties of Corson-compact spaces, Studia Math. 89 (1988) 197-229. [3] A. Avil´es, F. Cabello, J.M.F. Castillo, M. Gonz´alez, Y. Moreno, On separably injective Banach spaces, Advances in Math. 234 (2013) 192-216. [4] A. B laszczyk and A.R. Szyma´nski. Concerning Perovicenko's theorem. Bull. Acad. Polon. Sci. Math. 28 (1980) 311-314. [5] F. Cabello S´anchez and J.M.F. Castillo. The long homology sequence for quasi-Banach spaces, with appli- cations. Positivity 8 (2004), 379 -- 394. [6] F. Cabello and J.M.F. Castillo. Uniform boundedness and twisted sums of Banach spaces. Houston J. Math. 30 (2004), 523 -- 536. [7] F. Cabello S´anchez, J.M.F. Castillo, N.J. Kalton and D.T. Yost. Twisted sums with C(K)-spaces. Trans. Amer. Math. Soc. 355 (2003), 4523 -- 4541. [8] F. Cabello S´anchez, J.M.F. Castillo and D.T. Yost. Sobczyk's theorems from A to B. Extracta Math. 15 (2000) 391 -- 420. [9] J.M.F. Castillo and M. Gonz´alez, Three-space problems in Banach space theory. Lecture Notes in Math. 1667, Springer 1997. [10] J.M.F. Castillo, M. Gonz´alez, A. Plichko and D. Yost. Twisted properties of Banach spaces. Math. Scand. 89 (2001), 217 -- 244. [11] J.M.F. Castillo, Y. Moreno, On the Lindenstrauss-Rosenthal theorem. Israel J. Math. 140 (2004) 253 -- 270. [12] J.M.F. Castillo and Y. Moreno. Singular and cosingular exact sequences of quasi-Banach spaces. Arch. Math. 88 (2007) 123-132. [13] J.M.F. Castillo and M. A. Simoes, Property (V ) still fails the 3-space property. Extracta Math. 27 (2012) 5-11. [14] C. Correa and Daniel V. Tausk, Extension property and complementation of isometric copies of continuous functions spaces, Results in Mathematics, published online September 2014. DOI:10.1007/s00025-014-0411-5. [15] C. Correa and Daniel V. Tausk, Nontrivial twisted sums of c0 and C(K), Journal Funct. Anal. (to appear). [16] G. Godefroy, N.J. Kalton and G. Lancien. Subspaces of c0(N) and Lipschitz isomorphisms. Geom. Funct. Anal. 10 (2000) 798 - 820. [17] A.S. Granero. On the complemented subspaces of c0(I). Atti Sem. Mat. Fis. Univ. Modena 46 (1998), no. 1, 35 -- 36. [18] T. Jech, Set theory. The third millennium edition, revised and expanded. Springer Monographs in Mathe- matics, Springer-Verlag, Berlin, 2003. [19] O. Kalenda, Valdivia compact spaces in topology and Banach space theory, Extracta Math. vol. 15, (2000), 1 -- 85. [20] N.J. Kalton and N.T. Peck. Twisted sums of sequence spaces and the three-space problem. Trans. Amer. Math. Soc. 255 (1979), 1 -- 30. [21] W. Marciszewski and R. Pol, On Banach spaces whose norm-open sets are Fσ sets in the weak topology. J. Math. Anal. Appl. 350 (2009) 708-722. [22] S. Negrepontis, Banach spaces and topology, in "Handbook of Set Theoretic Topology" (K. Kunen and J.E. Vaughan, eds.) pp.1045-1142, North-Holland 1984. [23] A.Pe lczy´nski, Linear extensions, linear averagings, and their applications to linear topological classification of spaces of continuous functions,, Diss. Math. 58 (1968). [24] J. Reif, A note on Markusevic bases in weakly compactly generated Banach spaces Comment. Mat. Univ. Carolinae 15 (1974) 83-111. [25] H.P. Rosenthal. On injective Banach spaces and the spaces L∞(µ) for finite measures µ. Acta Math. 124 (1970), 205-248. [26] D.P. Sinha, On strong M-bases in Banach spaces with PRI, Collectanea Mathematica 51 (2000), 277 -- 284). [27] D. Yost, A different Johnson-Lindenstrauss space, New Zealand J. of Math. 36 (2007) 1-3. Departamento de Matem´aticas, Universidad de Extremadura, Avenida de Elvas, 06011-Badajoz, Spain E-mail address: [email protected]
1512.04790
1
1512
2015-12-15T14:07:18
Absolutely summing operators and atomic decomposition in bi-parameter Hardy spaces
[ "math.FA" ]
For $f \in H^p(\delta^2)$, $0<p\leq 2$, with Haar expansion $f=\sum f_{I \times J}h_{I\times J}$ we constructively determine the Pietsch measure of the $2$-summing multiplication operator \[\mathcal{M}_f:\ell^{\infty} \rightarrow H^p(\delta^2), \quad (\varphi_{I\times J}) \mapsto \sum \varphi_{I\times J}f_{I \times J}h_{I \times J}. \] Our method yields a constructive proof of Pisier's decomposition of $f \in H^p(\delta^2)$ \[|f|=|x|^{1-\theta}|y|^{\theta}\quad\quad \text{ and }\quad\quad \|x\|_{X_0}^{1-\theta}\|y\|^{\theta}_{H^2(\delta^2)}\leq C\|f\|_{H^p(\delta^2)}, \] where $X_0$ is Pisier's extrapolation lattice associated to $H^p(\delta^2)$ and $H^2(\delta^2)$. Our construction of the Pietsch measure for the multiplication operator $\mathcal{M}_f$ involves the Haar coefficients of $f$ and its atomic decomposition. We treated the one-parameter $H^p$-spaces in [P.F.X M\"uller, J.Penteker, $p$-summing multiplication operators, dyadic Hardy spaces and atomic decomposition, Houston Journal Math.,41(2):639-668,2015.].
math.FA
math
ABSOLUTELY SUMMING OPERATORS AND ATOMIC DECOMPOSITION IN BI-PARAMETER HARDY SPACES. PAUL F.X. M ULLER AND JOHANNA PENTEKER Abstract. For f ∈ H p(δ2), 0 < p ≤ 2, with Haar expansion f = P fI×J hI×J we constructively determine the Pietsch measure of the 2-summing multiplication operator Our method yields a constructive proof of Pisier's decomposition of f ∈ H p(δ2) Mf : ℓ∞ → H p(δ2), (ϕI×J ) 7→X ϕI×J fI×J hI×J . f = x1−θyθ and kxk1−θ X0 kykθ H 2(δ2) ≤ Ckf kHp(δ2), where X0 is Pisier's extrapolation lattice associated to H p(δ2) and H 2(δ2). Our construction of the Pietsch measure for the multiplication operator Mf involves the Haar coefficients of f and its atomic decomposition. We treated the one-parameter H p-spaces in [Houston Journal Math. 2015]. 1. Introduction Let Y0, Y be Banach spaces. An operator T ∈ L(Y0, Y ) is called 2-summing if there is a constant C such that for every choice of finite sequences (ϕi) in Y0, we have (1.1) (cid:18) n Xi=1 1 2 kT ϕik2(cid:19) ≤ C sup(cid:26)(cid:16) n Xi=1 1 2 ϕ∗(ϕi)2(cid:17) : ϕ∗ ∈ BY ∗ 0 ≤ 1(cid:27). In the early 70's the concepts of type and cotype were mainly developed by J. Hoffmann-Jørgensen, S. Kwapien, B. Maurey and G. Pisier, see [HJ74, Kwa72, Mau72a, Mau72b, MP73, MP76, Pis73]. A Banach space Y is called of cotype 2 if there is a constant C such that for all finite sequences (yi) in Y (cid:18) n Xi=1 1 2 kyik2(cid:19) ≤ C(cid:18)Z 1 0 (cid:13)(cid:13)(cid:13)(cid:13) n Xi=1 ri(t)yi(cid:13)(cid:13)(cid:13)(cid:13) 2 X 1 2 , dt(cid:19) where (ri)i∈N denotes the independent Rademacher system. One famous theorem due to Maurey ([Mau73, Mau74], see also [Pis78]) combining absolutely summing operators and the concept of cotype states that every bounded operator is 2-summing, whenever Y is of cotype 2. In particular, if T : ℓ∞ → Y kT ϕkY ≤ sup i∈N ϕi, Date: July 16, 2021. 2010 Mathematics Subject Classification. 42B30 46B25 46B09 46B42 46E40 47B10 60G42. 1 2 PAUL F.X. M ULLER AND JOHANNA PENTEKER then T satisfies (1.1) and by Pietsch's factorization theorem (cf. [Woj91]) there exists a constant C such that (1.2) kT ϕkY ≤ C(cid:18)ZΩ ϕ2dµ(cid:19) 1 2 , where µ is a Borel probability measure on Ω = B(ℓ∞)∗, called Pietsch measure. Another concept going back to the 70's are Hardy spaces of martingales and their atomic decomposition, cf. [FS72, Fef72, Ber79, Bro80, Gun80, CF80]. In our recent paper [MP15] we exhibited a connection between these two con- cepts. In the present work we further extend and exploit these newly found con- nections. We consider operators from ℓ∞ into bi-parameter dyadic Hardy spaces H p(δ2) that act as multipliers on the Haar system. By the above, these multi- plication operators are 2-summing and satisfy therefore (1.2). In our main result (Theorem 3.1) we determine explicit formulae for the Pietsch measure of these mul- tiplication operators. We recall that for general absolutely summing operators the existence of a Pietsch measure is given by a Hahn-Banach argument and is therefore not constructive. Let D be the set of dyadic intervals. Let (fI×J )I×J∈D×D be a real sequence indexed by the dyadic rectangles D × D. The space H p(δ2) consists of all functions where hI×J = hI ⊗ hJ , which satisfy fI×J hI×J , f = XI∈D XJ∈D kf kHp(δ2) = Z[0,1]2(cid:16)XI∈D XJ∈D f 2 I×J 1I×J(cid:17) 1 p p 2 dm! < ∞, where m denotes the Lebesgue measure on [0, 1]2. Every f ∈ H p(δ2) defines a multiplication operator of the form (1.3) Mf : ℓ∞(D × D) → H p(δ2) (ϕI×J ) 7→XI XJ ϕI×J fI×J hI×J . (1.4) For 1 ≤ p ≤ 2 the Hardy spaces H p(δ2) are of cotype 2 and therefore the multiplica- tion operators Mf are 2-summing and have Pietsch measures. In our main theorem (Theorem 3.1) we use the atomic decomposition of f ∈ H p(δ2) to give explicit for- mulae for these Pietsch measures. In particular, we determine ω = (ωI×J )I×J∈D×D with ωI×J ≥ 0 andP ωI×J ≤ 1 such that for all ϕ ∈ ℓ∞(D × D) the following holds 1 2 kMf (ϕ)kHp(δ2) ≤ Ckf kHp(δ2)(cid:18)XI∈D XJ∈D ϕI×J 2ωI×J(cid:19) . The explicit formulae for ω are given by equation (3.1) in Section 3. Multiplication operators such as given in (1.3) played an important role in the development of Banach space theory. See for instance the proof by Lindenstrauss and Pe lczy´nski on the uniqueness of the unconditional basis in ℓ1 ([LT77, LP68]). Bi-parameter Hardy spaces H p(δ2) may be regarded as vector-valued Hardy spaces H p X , where X = H p. In [MP15, Theorem 3.3, (3.19)] we obtained partially constructive formulae for the Pietsch measures of Haar multipliers on ℓ∞ into the vector-valued Hardy spaces H p X . In the scalar-valued case, i.e. X = R, we obtained fully constructive formulae for the Pietsch measures of the multiplication operators, ABSOLUTELY SUMMING OPERATORS AND ATOMIC DECOMPOSITION 3 see [MP15, Theorem 3.1.]. With this in mind, our present theorem (Theorem 3.1) gives fully constructive results for a special class of vector-valued Hardy spaces and simultaneously we extend in a non-trivial way the scalar-valued one-parameter case to the bi-parameter case. Application. The Banach spaces H p(δ2) form Banach lattices whose lattice struc- ture is induced by their unconditional basis (hI×J ) and they are related through Calder´on's product formula (1.5) H p(δ2) =(cid:0)H 1(δ2)(cid:1)1−θ(cid:0)H 2(δ2)(cid:1)θ , 0 < θ < 1, 1 p = 1 − θ + θ 2 . This follows by combining the one-parameter identities (cf. [FJ90, Theorem 8.2.]) with Calder´on's theorem ([Cal64, Paragraph 13.6]). Therefore, Pisier's extrapo- lation statement ([Pis79, Theorem 2.10]) can be adapted to the family of H p(δ2) spaces and reads in this setting as follows H p(δ2) = (X0)1−θ(H 2(δ2))θ, θ = 2 − 2 p . Here X0 is the Banach lattice of all elements x =PIPJ xI×J hI×J for which = sup  (cid:13)(cid:13)(cid:13)(cid:13)XI XJ xI×J 1−θyI×J θhI×J(cid:13)(cid:13)(cid:13)(cid:13)Hp(δ2) < ∞,   (1.6) kxkX0 where the supremum is taken over all y = PI,J yI×J hI×J with kykH 2(δ2) ≤ 1. Specifically, (1) asserts that given f ∈ H p(δ2) there is x ∈ X0 and y ∈ H 2(δ2) such that (1.7) f = x1−θyθ and kxk1−θ X0 kykθ H 2(δ2) ≤ Ckf kHp(δ2). Pisier shows in his proof that the weight ω = (ωI×J ) given by equation (1.4) yields factors for f . Hence, our explicit formulae for ω = (ωI×J ) determined in Theorem 3.1 allow us to give constructive factors of f satisfying (1.7). 2. Preliminaries 2.1. Bi-parameter Hardy spaces H p(δ2). The dyadic intervals D on the unit interval are given by D =(cid:8)(cid:2)2−n(k − 1), 2−nk(cid:2) : n, k ∈ N0, 0 ≤ k < 2n(cid:9) and the dyadic rectangles R on the unit square are given by R = D × D. Let C ⊆ R be a collection of dyadic rectangles. Then we denote by C∗ the pointset covered by the union of all dyadic rectangles in the collection C. The space ℓ∞(R) is the space of all sequences ϕ = (ϕIJ )I×J∈R, indexed by the dyadic rectangles, with kϕk∞ = supI×J∈R ϕIJ < ∞. For every I ∈ D we define the L∞- normalised Haar function hI to be +1 on the left half of I, −1 on the right half of I and zero on [0, 1] \ I. The an-isotropic 2D Haar system (hI×J )I×J∈R indexed by the dyadic rectangles is defined as follows hI×J (s, t) := hI (s)hJ (t), I, J ∈ D, (s, t) ∈ [0, 1]2. 4 PAUL F.X. M ULLER AND JOHANNA PENTEKER Let (fIJ )I×J∈R be a real sequence and f = (fIJ )I×J∈R the real vector indexed by the dyadic rectangles. The square function of f is defined as follows S(f )(s, t) =(cid:18) XI×J∈R f 2 IJ 1I×J (s, t)(cid:19) 1 2 , (s, t) ∈ [0, 1]2. The bi-parameter dyadic Hardy space H p(δ2), 0 < p ≤ 2, consists of vectors f = (fIJ )I×J∈R for which kf kHp(δ2) = Z[0,1]2 Sp(f )(s, t) dm(s, t)! 1 p < ∞, where m is the Lebesgue measure on [0, 1]2. Systematically we use the notation kf k2 = kf kH 2(δ2). For convenience we identify f = (fIJ )I×J∈R ∈ H p(δ2) with its formal Haar series (2.1) f = XI×J∈R fIJ hI×J . 2.2. Atomic decomposition. Let 0 < p ≤ 2 and f ∈ H p(δ2) with Haar expansion (2.1). For every n ∈ Z we define the set and the collection of dyadic rectangles Fn =(cid:8)(s, t) ∈ [0, 1]2 : S(f )(s, t) > 2n(cid:9) Rn =(cid:26)I × J ∈ R : I × J ∩ Fn > I × J 2 , I × J ∩ Fn+1 ≤ I × J 2 (cid:27) . Then f =Pn∈Z fn, where and the following inequalities hold kfnkp kf kp (2.2) Hp(δ2) ≤ Xn∈Z fIJ hI×J fn = XI×J∈Rn Hp(δ2) ≤ Xn∈Z R∗ n1− p 2 kfnkp 2 ≤ Apkf kp Hp(δ2). The family (fn, Rn)n∈Z is called the atomic decomposition of u ∈ H p(δ2). This decomposition originates from [Fef72, Ber79, Gun80, CF80]. Note that the right-hand side inequality in (2.2) results from the following S2(fn)1F c n+1 dm ≤ 2 · 22(n+1)R∗ n kfnk2 (2.3) 2 =Z[0,1]2 ≤ 8 · 22n(cid:12)(cid:12)(cid:12)nMS(1Fn ) > S2(fn) dm ≤ 2Z[0,1]2 2o(cid:12)(cid:12)(cid:12) 1 ≤ C 22nFn. MS(1Fn )(s, t) = sup R∋(s,t) 1 RZR 1Fn dm, Here MS is the strong maximal operator (cf. [JMZ35],[FFW95]) in [0, 1]2 given by where the supremum is taken over all rectangles R in [0, 1]2 with side length parallel to the axes. Boundedness estimates for the strong maximal operator (cf. [JMZ35]) give rise to bi-parameter Fefferman-Stein strong maximal operator estimates (cf. [FS71, Theorem 1.]). We exploit these Fefferman-Stein inequalities in the following form. ABSOLUTELY SUMMING OPERATORS AND ATOMIC DECOMPOSITION 5 Lemma 2.1. Fix ε > 0. Suppose that for each I × J ∈ R the subset EI×J ⊆ I × J is a measurable set with EI×J I×J > ε. Then for any f ∈ H p(δ2), 0 < p < ∞, with Haar expansion f =PI×J∈R fIJ hI×J , the following holds fIJ 21EI×J(cid:17) kf kHp(δ2) ≤ Cp(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16) XI×J∈R . 1 2(cid:13)(cid:13)(cid:13)(cid:13)Lp Frazier and Jawerth ([FJ90, Theorem 2.7.]) give a proof for the one-parameter version of this lemma. Their proof can be adapted to the setting above. 2.3. Modified Holder inequality. See [HLP52, p. 61 (65.)]. Let (Ω, Σ, µ) be a measure space and r > 1 or r < 0. Then for all measurable functions f, g on Ω (2.4) ZΩ f rg1−rdµ ≥(cid:18)ZΩ f dµ(cid:19)r(cid:18)ZΩ g dµ(cid:19)1−r . Let 0 < p ≤ 2. Every f ∈ H p(δ2) defines a multiplication operator of the form 3. The main Theorem Mf : ℓ∞(R) → H p(δ2), (ϕIJ ) 7→ XI×J∈R ϕIJ fIJ hI×J and clearly we have (cid:13)(cid:13)Mf : ℓ∞(R) → H p(δ2)(cid:13)(cid:13) ≤ kf kHp(δ2). Banach space theory as described in the introduction guarantees that these multi- plication operators are 2-summing and satisfy (1.4). In Theorem 3.1 we determine explicit formulae for the weights ω = (ωI×J ) given in (1.4). Every multiplication operator Mf is induced by a function f ∈ H p(δ2). These functions admit an atomic decomposition (fn, Rn)n∈N satisfying the equations in (2.2). This is the input for our construction and the output is equation (3.1) determining ω explicitly. Theorem 3.1. Let 0 < p ≤ 2 and f ∈ H p(δ2) with Haar expansion and atomic decomposition (fn, Rn)n∈Z. Then the sequence (ωIJ )I×J∈R, defined by f = XI×J∈R fIJ hI×J (3.1) ωIJ = 1 Apkf kp Hp(δ2) satisfies R∗ n1− p 2 f 2 IJ IJ kfnk2−p 2 , I × J ∈ Rn, and for each ϕ ∈ ℓ∞(R) the following inequality holds ωIJ ≤ 1 XI×J∈R kMf (ϕ)kHp(δ2) ≤ Apkf kHp(δ2)(cid:16) XI×J∈R ϕIJ 2ωIJ(cid:17) 1 2 . 6 PAUL F.X. M ULLER AND JOHANNA PENTEKER Proof. Note that the left-hand side inequality of (2.2) depends only on the fact that (Rn)n∈Z forms a partition of R. Hence, for all ϕ = (ϕIJ ) ∈ ℓ∞(R) the following estimate holds kfnkp 2R∗ n1− p 2 . n1− p 2 p 2 R∗ Hp(δ2) p ϕIJ fIJ hI×J(cid:13)(cid:13)(cid:13)(cid:13) ϕIJ fIJ hI×J(cid:13)(cid:13)(cid:13)(cid:13) hI×J(cid:13)(cid:13)(cid:13)(cid:13) IJ(cid:17) f 2 IJ kfnk2 2 fIJ kfnk2 ϕ2 IJ 2 p p 2 ϕIJ =(cid:13)(cid:13)(cid:13)(cid:13)Xn∈Z XI×J∈Rn ≤ Xn∈Z(cid:13)(cid:13)(cid:13)(cid:13) XI×J∈Rn = Xn∈Z(cid:13)(cid:13)(cid:13)(cid:13) XI×J∈Rn hI×J(cid:13)(cid:13)(cid:13)(cid:13) =(cid:16) XI×J∈Rn ≤ Xn∈Z(cid:16) XI×J∈Rn ϕ2 IJ 2 p IJkfnkp 2R∗ fIJ kfnk2 Hp(δ2) f 2 IJ kfnk2 2 f 2 IJ ϕ2 IJ ϕ2 IJ kfnk2−p 2 p p 2 f 2 IJ kfnk2 2 IJ(cid:17) n1− p 2(cid:17) 2(cid:16)kfnkp 2(cid:16)Xn∈Z p n1− p 2(cid:17) kfnkp 2R∗ n1− p 2 2R∗ n1− p 2 2(cid:17)1− p 2(cid:17)1− p 2 . IJR∗ kfnkp 2R∗ n1− p p ϕIJ fIJ hI×J(cid:13)(cid:13)(cid:13)(cid:13) Hp(δ2) (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R With and Holder's inequality we get p ϕIJ (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈Rn ϕIJ fIJ hI×J(cid:13)(cid:13)(cid:13)(cid:13) = Xn∈Z(cid:16) XI×J∈Rn ≤(cid:16)Xn∈Z XI×J∈Rn ϕIJ fIJ hI×J(cid:13)(cid:13)(cid:13)(cid:13) 1− p ≤ A p kf k p 2 = Apkf kp (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R Hp(δ2) 2 ) p(1− p Hp(δ2)(cid:16)Xn∈Z XI×J∈Rn Hp(δ2)(cid:16)Xn∈Z XI×J∈Rn ϕ2 IJ f 2 IJ ϕ2 IJ kfnk2−p 2 f 2 IJ kf kp kfnk2−p 2 IJR∗ p 2 n1− p 2(cid:17) Hp(δ2)Ap IJR∗ n1− p 2(cid:17) p 2 . Recall that (3.2) kfnk2 2 = XI×J∈Rn f 2 IJ IJ. By the right-hand side inequality in equation (2.2) and by equation (3.2) we obtain for the sequence (ωIJ )I×J∈R, defined by ωIJ = 1 Apkf kp Hp(δ2) R∗ n1− p 2 f 2 IJ IJ kfnk2−p 2 , I × J ∈ Rn, ABSOLUTELY SUMMING OPERATORS AND ATOMIC DECOMPOSITION 7 the following estimate ωIJ = XI×J∈R = 1 Apkf kp 1 Apkf kp Hp(δ2) Xn∈Z XI×J∈Rn n1− p Hp(δ2) Xn∈Z R∗ R∗ n1− p 2 f 2 IJ IJ kfnk2−p 2 2 kfnkp 2 ≤ 1. (cid:3) 4. Another application of the atomic decomposition Pisier's extrapolation lattice X0 defined in (1.6) is known to coincide with H 1(δ2). This follows by a specialisation of a general theorem of Cwikel and Nilsson (see [CNS03]). Their extrapolation method is applicable, since H p(δ2) spaces are re- lated through Calderon's product formula (cf. equation (1.5)). The space X0 is of particular importance to our work in this paper. Hence, we take the opportunity to complement the work of [CNS03, FJ90] with a direct argument based on the atomic decomposition of H p(δ2). We build our strategy by exploiting the formulae used by [Mul05, GMP05, Bow13] for similar purposes. In particular, we refer to Bownik's paper [Bow13] for the formula (4.3) and the idea of using Lemma 2.1 in the proof of the following theorem. Theorem 4.1. Let f ∈ H p(δ2), 0 < p ≤ 2, with Haar expansion f =P fIJ hI×J . Then for 0 < θ < 1 and q given by (4.1) the following holds: 1 q = 1 − θ p + θ 2 , cpkf k1−θ Hp(δ2) ≤ sup  (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R fIJ 1−θgIJθhI×J(cid:13)(cid:13)(cid:13)(cid:13)Hq (δ2)   ≤ kf k1−θ Hp(δ2), where the supremum is taken over all functions g =P gIJ hI×J with kgk2 ≤ 1. Proof. We start with the proof of the right-hand side inequality. Let fIJ 1−θgIJ θhI×J . h = XI×J∈R Then, by applying Holder's inequality for sequence spaces with 1 − θ + θ = 1, we obtain the following inequality for the square functions Sq(h) ≤ Sq(1−θ)(f ) Sqθ(g). Integrating over [0, 1]2 and applying Holder's inequality with q(1−θ) p + qθ 2 = 1 yields khkHq (δ2) ≤ kf k1−θ Hp(δ2) kgkθ 2. For the left-hand side inequality we show that for every f ∈ H p(δ2) there exists a function g ∈ H 2(δ2) such that kgk2 2 ≤ cpkf kp Hp(δ2) and (4.2) (cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R fIJ 1−θgIJ θhI×J(cid:13)(cid:13)(cid:13)(cid:13) q Hq (δ2) ≥ Cp kf kp Hp(δ2). 8 PAUL F.X. M ULLER AND JOHANNA PENTEKER Let (fn, Rn)n∈Z be the atomic decomposition of f ∈ H p(δ2). Let g =P gIJ hI×J , where gIJ = 2− n 2 (2−p)fIJ , (4.3) (4.4) Then, by equation (2.3), we have kgk2 2 = Xn∈Z 2−n(2−p)kfnk2 2 ≤ CXn∈Z ≤ cpkf kp Hp(δ2). I × J ∈ Rn. 2−n(2−p)22nFn = CXn∈Z 2npFn To prove equation (4.2) we use Lemma 2.1 with sets EI×J = I × J ∩ Fn, for I × J ∈ Rn and obtain (4.5) kf kp ≤ C q Hp(δ2) = Z[0,1]2 Sp(f ) dm! p Z[0,1]2(cid:16)XI×J Let h = (cid:16)PI×J∈R fIJ 21EI×J(cid:17) p Z[0,1]2 (cf. equation (2.4)) we have Z[0,1]2 hp dm! (4.6) 1 2 q q p Z[0,1]2 fIJ 21EI×J(cid:17) p Sp(f ) dm!1− q p Z[0,1]2 dm! p 2 q Sp(f ) dm!1− q p . . Note that by the modified Holder inequality Sp(f ) dm!1− q p ≤Z[0,1]2 hqSp−q(f ) dm. Combining equation (4.5) and (4.6) yields (4.7) kf kp Hp(δ2) ≤ C q q 2 Sp−q(f ) dm fIJ 21EI×J(cid:17) pZ[0,1]2(cid:16) XI×J∈R pZ[0,1]2(cid:16)Xn∈Z XI×J∈Rn fIJ 21I×J 1Fn(cid:17) = C q q 2 Sp−q(f ) dm. We know that S(f )1Fn > 2n1Fn. Since q > p, it follows that (4.8) S(f )p−q1Fn < 2−n(q−p)1Fn. Equation (4.1) gives the identity q − p = qθ into equation (4.7) yields 2 (2 − p). Hence, putting equation (4.8) (4.9) kf kp Hp(δ2) ≤ C q pZ[0,1]2(cid:16)Xn∈Z pZ[0,1]2(cid:16)Xn∈Z pZ[0,1]2(cid:16) XI×J∈R p(cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R ≤ C q = C q = C q q 2 dm fIJ 21I×J 1Fn(cid:17) fIJ 21I×J(cid:17) 2−nθ(2−p) XI×J∈Rn 2−nθ(2−p) XI×J∈Rn fIJ 2(1−θ)gIJ 2θ1I×J(cid:17) dm q 2 q 2 dm q fIJ 1−θgIJ θhI×J(cid:13)(cid:13)(cid:13)(cid:13) . Hq (δ2) ABSOLUTELY SUMMING OPERATORS AND ATOMIC DECOMPOSITION 9 Summarizing equations (4.4) and (4.9) yields kf k1−θ Hp(δ2)kgkθ 2 ≤ c θ 2 p kf k p q Hp(δ)2 ≤ Cp(cid:13)(cid:13)(cid:13)(cid:13) XI×J∈R fIJ 1−θgIJ θhI×J(cid:13)(cid:13)(cid:13)(cid:13)Hq(δ2) . (cid:3) Acknowledgements. We would like to thank M. Bownik for helpful discussions during the preparation of this paper. This research has been supported by the Austrian Science foundation (FWF) Pr.Nr.P22549, Pr.Nr.P23987 and Pr.Nr.P28352. References [Ber79] A. Bernard. Espaces H 1 de martingales `a deux indices. Dualit´e avec les martingales de type "BMO". Bull. Sci. Math. (2), 103(3):297 -- 303, 1979. [Bow13] M. Bownik. Extrapolation of discrete Triebel-Lizorkin spaces. Math. Nachr., 286(5- [Bro80] [Cal64] [CF80] 6):492 -- 502, 2013. J. Brossard. Comparison des "normes" Lp du processus croissant et de la variable maximale pour les martingales r´eguli`eres `a deux indices. Th´eor`eme local correspondant. Ann. Probab., 8(6):1183 -- 1188, 1980. A.-P. Calder´on. Intermediate spaces and interpolation, the complex method. Studia Math., 24:113 -- 190, 1964. S.Y.A. Chang and R. Fefferman. A continuous version of duality of H 1 with BMO on the bidisc. Ann. of Math. (2), 112(1):179 -- 201, 1980. [CNS03] M. Cwikel, P. G. Nilsson, and G. Schechtman. Interpolation of weighted Banach lattices. A characterization of relatively decomposable Banach lattices. Mem. Amer. Math. Soc., 165(787):vi+127, 2003. C. Fefferman. Estimates for double Hilbert transforms. Studia Math., 44:1 -- 15, 1972. Collection of articles honoring the completion by Antoni Zygmund of 50 years of scien- tific activity, I. [Fef72] [FJ90] [FFW95] C. Fefferman, R. Fefferman, and S. Wainger, editors. Essays on Fourier analysis in honor of Elias M. Stein, volume 42 of Princeton Mathematical Series. Princeton Uni- versity Press, Princeton, NJ, 1995. M. Frazier and B. Jawerth. A discrete transform and decompositions of distribution spaces. J. Funct. Anal., 93(1):34 -- 170, 1990. C. Fefferman and E. M. Stein. Some maximal inequalities. Amer. J. Math., 93:107 -- 115, 1971. C. Fefferman and E. M. Stein. H p spaces of several variables. Acta Math., 129(3-4):137 -- 193, 1972. [FS71] [FS72] [GMP05] S. Geiss, P.F.X. Muller, and V. Pillwein. A remark on extrapolation of rearrangement operators on dyadic H s, 0 < s ≤ 1. Studia Math., 171(2):197 -- 205, 2005. [Gun80] R. F. Gundy. In´egalit´es pour martingales `a un et deux indices: l'espace H p. In Eighth Saint Flour Probability Summer School -- 1978 (Saint Flour, 1978), volume 774 of Lec- ture Notes in Math., pages 251 -- 334. Springer, Berlin, 1980. J. Hoffmann-Jørgensen. Sums of independent Banach space valued random variables. Studia Math., 52:159 -- 186, 1974. [HJ74] [HLP52] G.H. Hardy, J.E. Littlewood, and G. P´olya. Inequalities. Cambridge, at the University Press, 1952. 2d ed. [JMZ35] B. Jessen, J. Marcinkiewicz, and A. Zygmund. Note on the differentiability of multiple integrals. Fundamenta Mathematicae, 25(1):217 -- 234, 1935. [Kwa72] S. Kwapie´n. Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients. Studia Math., 44:583 -- 595, 1972. Collection of articles honoring the completion by Antoni Zygmund of 50 years of scientific activity, VI. J. Lindenstrauss and A. Pe lczy´nski. Absolutely summing operators in Lp-spaces and their applications. Studia Math., 29:275 -- 326, 1968. [LP68] 10 [LT77] PAUL F.X. M ULLER AND JOHANNA PENTEKER J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. I. Springer-Verlag, Berlin- New York, 1977. Sequence spaces, Ergebnisse der Mathematik und ihrer Grenzgebiete, Vol. 92. [Mau72a] B. Maurey. Espaces de cotype (p, q) et th´eor`emes de rel`evement. C. R. Acad. Sci. Paris S´er. A-B, 275:A785 -- A788, 1972. [Mau72b] B. Maurey. Espaces de type (p, q), th´eor`emes de factorisation et de plongement. C. R. Acad. Sci. Paris S´er. A-B, 274:A1939 -- A1941, 1972. [Mau73] B. Maurey. Une nouvelle d´emonstration d'un th´eor`eme de Grothendieck. In S´eminaire Maurey-Schwartz Ann´ee 1972 -- 1973: Espaces Lp et applications radonifiantes, Exp. No. 22, page 7. Centre de Math., ´Ecole Polytech., Paris, 1973. [MP73] [Mau74] B. Maurey. Th´eor`emes de factorisation pour les op´erateurs lin´eaires `a valeurs dans les espaces Lp. Soci´et´e Math´ematique de France, Paris, 1974. With an English summary, Ast´erisque, No. 11. B. Maurey and G. Pisier. Caract´erisation d'une classe d'espaces de Banach par des propri´et´es de s´eries al´eatoires vectorielles. C. R. Acad. Sci. Paris S´er. A-B, 277:A687 -- A690, 1973. B. Maurey and G. Pisier. S´eries de variables al´eatoires vectorielles ind´ependantes et propri´et´es g´eom´etriques des espaces de Banach. Studia Math., 58(1):45 -- 90, 1976. Paul F. X. Muller and Johanna Penteker. p-summing multiplication operators, dyadic Hardy spaces and atomic decomposition. Houston J. Math., 41(2):639 -- 668, 2015. [MP15] [MP76] [Pis73] [Mul05] P.F.X. Muller. Isomorphisms between H 1 spaces, volume 66 of Instytut Matematyczny Polskiej Akademii Nauk. Monografie Matematyczne (New Series) [Mathematics In- stitute of the Polish Academy of Sciences. Mathematical Monographs (New Series)]. Birkhauser Verlag, Basel, 2005. G. Pisier. Sur les espaces de Banach qui ne contiennent pas uniform´ement de l1 Acad. Sci. Paris S´er. A-B, 277:A991 -- A994, 1973. G. Pisier. Une nouvelle classe d'espaces de Banach v´erifiant Grothendieck. Ann. Inst. Fourier (Grenoble), 28(1):x, 69 -- 90, 1978. G. Pisier. La m´ethode d'interpolation complexe: applications aux treillis de Banach. In S´eminaire d'Analyse Fonctionnelle (1978 -- 1979), pages Exp. No. 17, 18. ´Ecole Poly- tech., Palaiseau, 1979. le th´eor`eme de n. C. R. [Pis78] [Pis79] [Woj91] P. Wojtaszczyk. Banach spaces for analysts, volume 25 of Cambridge Studies in Ad- vanced Mathematics. Cambridge University Press, Cambridge, 1991. P.F.X. Muller, Institute of Analysis, Johannes Kepler University Linz, Austria, 4040 Linz, Altenberger Strasse 69 E-mail address: [email protected] J. Penteker, Institute of Analysis, Johannes Kepler University Linz, Austria, 4040 Linz, Altenberger Strasse 69 E-mail address: [email protected]
1711.08642
1
1711
2017-11-23T10:32:16
On $\ell^1$-regularization under continuity of the forward operator in weaker topologies
[ "math.FA", "math.NA" ]
Our focus is on the stable approximate solution of linear operator equations based on noisy data by using $\ell^1$-regularization as a sparsity-enforcing version of Tikhonov regularization. We summarize recent results on situations where the sparsity of the solution slightly fails. In particular, we show how the recently established theory for weak*-to-weak continuous linear forward operators can be extended to the case of weak*-to-weak* continuity. This might be of interest when the image space is non-reflexive. We discuss existence, stability and convergence of regularized solutions. For injective operators, we will formulate convergence rates by exploiting variational source conditions. The typical rate function obtained under an ill-posed operator is strictly concave and the degree of failure of the solution sparsity has an impact on its behavior. Linear convergence rates just occur in the two borderline cases of proper sparsity, where the solutions belong to $\ell^0$, and of well-posedness. For an exemplary operator, we demonstrate that the technical properties used in our theory can be verified in practice. In the last section, we briefly mention the difficult case of oversmoothing regularization where $x^\dag$ does not belong to $\ell^1$.
math.FA
math
On (cid:96)1-regularization under continuity of the forward operator in weaker topologies Daniel Gerth∗ and Bernd Hofmann∗ October 24, 2019 Abstract Our focus is on the stable approximate solution of linear operator equa- tions based on noisy data by using (cid:96)1-regularization as a sparsity-enforcing version of Tikhonov regularization. We summarize recent results on sit- uations where the sparsity of the solution slightly fails. In particular, we show how the recently established theory for weak*-to-weak continuous linear forward operators can be extended to the case of weak*-to-weak* continuity. This might be of interest when the image space is non-reflexive. We discuss existence, stability and convergence of regularized solutions. For injective operators, we will formulate convergence rates by exploiting variational source conditions. The typical rate function obtained under an ill-posed operator is strictly concave and the degree of failure of the solution sparsity has an impact on its behavior. Linear convergence rates just occur in the two borderline cases of proper sparsity, where the so- lutions belong to (cid:96)0, and of well-posedness. For an exemplary operator, we demonstrate that the technical properties used in our theory can be verified in practice. In the last section, we briefly mention the difficult case of oversmoothing regularization where x† does not belong to (cid:96)1. 1 Introduction We are going to deal with the stable solution of linear operator equations Ax = y (1) with a bounded linear operator A : (cid:96)1 → Y , mapping from the non- reflexive infinite dimensional space (cid:96)1 of absolutely summable infinite real or complex sequences to an infinite dimensional Banach space Y . Instead of the exact right-hand side y from the range R(A) of A we assume to have only noisy data yδ ∈ Y available which satisfy the deterministic noise model (cid:107)y − yδ(cid:107)Y ≤ δ (2) ∗Chemnitz University of Technology, Faculty of Mathematics, 09107 Chemnitz, Germany, {daniel.gerth,bernd.hofmann}@mathematik.tu-chemnitz.de 1 with prescribed noise level δ > 0. Our focus for solving equation (1) is on the method of (cid:96)1-regularization, where for regularization parameters α > 0 the minimizers xδ α of the extremal problem (cid:107)Ax − yδ(cid:107)p Y + α(cid:107)x(cid:107)(cid:96)1 → min, 1 p subject to x ∈ (cid:96)1, (3) are used as approximate solutions. This method is a sparsity-enforcing version of Tikhonov regularization, possessing applications in different branches of imaging, natural sciences, engineering and mathematical fi- nance. It was comprehensively analyzed with all its facets and varieties in the last fifteen years (cf., e.g., the corresponding chapters in the books [31, 32, 33] and the papers [2, 4, 8, 12, 20, 21, 22, 25, 28, 29]). We re- strict our considerations to injective operators A such that the element 2, ...) ∈ (cid:96)1 denotes the uniquely determined solution to (1). † † x† = (x 1, x For assertions concerning the case of non-injective operators A in the con- text of (cid:96)1-regularization, we refer to [10]. In the non-injective case, even the (cid:96)1-norm minimizing solutions need not be uniquely determined. As a consequence, very technical conditions must be introduced in order to formulate convergence assertions and rates. In our framework, the Propo- sitions 4.3 and 5.4 below would have to be adapted, which however is out of the scope of this paper. With the paper [5] as starting point and preferably based on vari- ational source conditions first introduced in [23], convergence rates for (cid:96)1-regularization of operator equations (1) and variants like elastic-net (cid:107)Ax − yδ(cid:107)p Y + α 1 p (cid:107)x(cid:107)(cid:96)2 + η(cid:107)x(cid:107)(cid:96)1 → min, subject to x ∈ (cid:96)1 (4) (cid:18) 1 2 (cid:19) have been verified under the condition that the sparsity assumption slightly fails (cf. [6, 13, 14]). This means that the solution x† ∈ (cid:96)1 is not sparse, abbreviated as x† /∈ (cid:96)0. Most recently in [11], the first author and Jens Flemming have shown that complicated conditions on A, usually sup- posed for proving convergence rates in (cid:96)1-regularization (cf. [5, Assump- tion 2.2 (c)] and condition (9) below), can be simplified to the requirement of weak∗-to-weak continuity of the injective operator A. This seems to be convincing if Y is a reflexive Banach space. The present paper, however, makes assertions also in the case that A is only weak∗-to-weak∗ continu- ous, which is of interest for non-reflexive Banach spaces Y . Moreover, we complement results from [11], for example with respect to the well-posed situation. The paper is organized as follows. In Section 2 we recall basic prop- erties of (cid:96)1-regularization. We proceed in Section 3 by discussing the ill-posedness of equation (1). We mention that in particular variational source conditions allow us to deal with the ill-posedness and yield con- vergence rates. For our convergence analysis a particular property of the operator is necessary. In Section 4 we show that weak*-to-weak continuity and injectivity imply this property. Interestingly, the same property holds under weak*-to-weak* continuity and injectivity as shown in Section 5. There we also derive the convergence rates which hold for both continuity assumptions. Finally, we demonstrate that even the case of a well-posed 2 operator is reflected in our property in Section 6. There we also hint at the case of oversmoothing regularization, which occurs when one employs (cid:96)1-regularization although the true solution x† does not belong to (cid:96)1. 2 Preliminaries and basic propositions In this paper, we consider the variant (3) of (cid:96)1-regularization with some exponent p > 1 and with a regularization parameter α > 0. Let y ∈ R(A). Then, due to the injectivity of A, there exists a uniquely determined so- lution x† ∈ (cid:96)1 to (1). With the following Proposition 2.1 we recall the assertions of Proposition 2.8 in [5] with respect to existence, stability, con- vergence and sparsity of the (cid:96)1-regularized solutions xδ α. The proof ibidem emphasizes the fact that most of these properties follow directly from the general theory of Tikhonov regularization in Banach spaces (cf., e.g., [23, Section 3] and [33, Section 4.1]). Since for p > 1 the Tikhonov functional to be minimized in (3) is strictly convex, the regularized solutions xδ α, whenever they exist, are uniquely determined for all α > 0. Proposition 2.1. Let A : (cid:96)1 → Y be weak∗-to-weak continuous, i.e., xn (cid:42)∗ x0 in (cid:96)1 implies that Axn (cid:42) Ax0 in Y . Then for all α > 0 and all yδ ∈ Y there exist uniquely determined minimizers xδ α ∈ (cid:96)1 of the Tikhonov functional from (3). These regularized solutions are sparse, α ∈ (cid:96)0, and they are stable with respect to the data, i.e., small i.e., xδ perturbations in yδ in the norm topology of Y lead only to small changes If δn → 0 and if the α with respect to the weak∗-topology in (cid:96)1. in xδ regularization parameters αn = α(δn, yδn ) are chosen such that αn → 0 αn converges in the weak∗-topology of and δp n αn (cid:96)1 to the uniquely determined solution x† of the operator equation (1). αn(cid:107)(cid:96)1 = (cid:107)x†(cid:107)(cid:96)1 , which, as a consequence of the Moreover we have lim weak∗ Kadec-Klee property in (cid:96)1 (see, e.g., [3, Lemma 2.2]), implies norm convergence → 0 as n → ∞, then xδn n→∞(cid:107)xδn n→∞(cid:107)xδn lim αn − x †(cid:107)(cid:96)1 = 0. The weak∗-to-weak continuity of A in combination with the stabilizing property of the penalty functional (cid:107)x(cid:107)(cid:96)1 in (cid:96)1 together with an appropriate choice of the regularization parameter α > 0 represent basic assumptions of Proposition 2.1. In contrast to regularization in reflexive Banach space, where the level sets of the norm functional are weakly compact, we have in (cid:96)1 weak∗ compactness of the corresponding level sets according to the se- quential Banach-Alaoglu theorem (cf., e.g., [30, Theorems 3.15 and 3.17]), which we present in form of the following lemma. Lemma 2.2. The closed unit ball of a Banach space X is compact in the weak∗-topology if there is a separable Banach space Z (predual space) with dual Z∗ = X. Then any bounded sequence {xn}n∈N in X has a weak∗- convergent subsequence {xnk}k∈N such that xnk (cid:42)∗ x0 ∈ X as k → ∞. The occurring kind of compactness of the level sets with X = (cid:96)1 and α of the func- predual space Z = c0 ensures the existence of minimizers xδ tional (3). 3 Throughout this paper, we use the terms 'continuous', 'compact' or 'lower semicontinuous' for an operator, a set or a functional always in the sense of 'sequentially continuous', 'sequentially compact' or 'sequen- tially lower semicontinuous'. As the Lemmas 6.3 and 6.5 from [9] show, there is no reason for a distinction in case of using weak topologies. From Lemma 2.7 and Proposition 2.4 in [5] one can take assertions concerning sufficient conditions for the weak∗-to-weak continuity of A, which we sum- marize in the Proposition 2.3 below. As also indicated in Proposition 2.1, for the choice of α, the so-called regularization property α(δ, yδ) → 0 and δp α(δ, yδ) → 0 δ → 0, as (5) where α tends to zero, but sufficiently slow, plays an important role. In our studies, we consider on the one hand a priori parameter choices αAP RI = α(δ) defined as α(δ) := δp ϕ(δ) , 0 < δ ≤ δ, (6) with concave index functions ϕ. We call ϕ : [0,∞) → [0,∞) an index function if ϕ with ϕ(0) = 0 is continuous and strictly increasing. Ob- viously, an a priori parameter choice αAP RI from (6) with an arbitrary concave index function ϕ satisfies (5) as limδ→+0 ϕ(δ) = 0 is valid for each ϕ(δ) δp−1 → 0 as δ → 0, because δp−1 index function and we have δp δ ϕ(δ) is is an index function for all exponents p > 1 in (3) and the factor bounded whenever ϕ is concave. ϕ(δ) = δ On the other hand, we consider the sequential discrepancy principle, comprehensively analyzed in [1] (see also [24]), as a specific a posteriori parameter choice αSDP = α(δ, yδ) for the regularization parameter. For prescribed τ > 1, 0 < q < 1, and a sufficiently large value α0 > 0, we let ∆q := {αj > 0 : αj = qjα0, j = 1, 2, ...}. Given δ > 0 and yδ ∈ Y , we choose α = αSDP ∈ ∆q according to the sequential discrepancy principle such that (cid:107)Axδ α − yδ(cid:107) ≤ τ δ < (cid:107)Axδ α/q − yδ(cid:107). (7) By Theorem 1 in [1] it has been shown that there is δ > 0 such that αSDP is well-defined for 0 < δ ≤ δ and satisfies (5) whenever data compatibility in the sense of [1, Assumption 3] takes place. Consequently, both regularization parameter choices α = αAP RI and α = αSDP are applicable for the (cid:96)1-regularization in order to get exis- tence, stability and convergence of regularized solutions in the sense of Proposition 2.1. Now we are going to discuss conditions under which weak∗-to-weak continuity of A : (cid:96)1 → Y can be obtained. The occur- ring cross connections are relevant in order to ensure existence, stability and convergence of regularized solutions, but they have also an essential impact on convergence rates which will be discussed in Section 4. 4 Proposition 2.3. Let A : (cid:96)1 → Y with adjoint operator A∗ : Y ∗ → (cid:96)∞ satisfy the condition R(A ∗ ) ⊆ c0, (8) where c0 is the Banach space of real-valued sequences converging to zero equipped with the supremum norm. Then A is weak∗-to-weak continuous. In particular, (8) is fulfilled whenever there exist, for all k ∈ N, source elements f (k) ∈ Y ∗ such that the system of source conditions ∗ e(k) = A f (k) (9) holds true, where {e(k)}k∈N is the sequence of k-th standard unit vectors which forms a Schauder basis in c0. Under the condition (9) we even have the equality (cid:96)∞ R(A∗) = c0. (10) The paper [2] shows that the condition (9), originally introduced by Grasmair in [19], can be verified for a wide class of applied linear inverse problems. But as also the counterexamples in [12] indicate, it may fail if the underlying basis smoothness is insufficient. However, weak∗-to- weak continuity of A can be reformulated in several ways as the following proposition, proven in [10, Lemma 2.1], shows. This proposition brings more order into the system of conditions. Proposition 2.4. The three assertions {Ae(k)}k∈N converges in Y weakly to zero, i.e. Ae(k) k→∞ (cid:42) 0 , (i) (ii) R(A∗) ⊆ c0 , (iii) A is weak∗-to-weak continuous , are equivalent. ∞(cid:80) As outlined in [5], the operator equation (1) with operator A : (cid:96)1 → Y is often motivated by a background operator equation Ax = y with an injective and bounded linear operator A mapping from the infinite dimen- sional Banach space X with uniformly bounded Schauder basis {u(k)}k∈N, i.e. (cid:107)u(k)(cid:107) X ≤ K < ∞, to the Banach space Y . Here, following the set- ting in [19] we take into account a synthesis operator L : (cid:96)1 → X defined xku(k) for x = (x1, x2, ...) ∈ (cid:96)1, which is a well-defined, as Lx := injective and bounded linear operator, and so is the composite operator A = A ◦ L : (cid:96)1 → Y . In particular A is always weak∗-to-weak continu- ous if A has a bounded extension to (cid:96)p, 1 < p < ∞, as this yields (i) in Proposition 2.4. Even more specific, A is weak*-to-weak continuous if X is a Hilbert space. Since this case appears rather often in practice, the continuity property comes "for free" in this situation. k=1 3 Ill-posedness and conditional stability In this section, we discuss ill-posedness phenomena of the operator equa- tion (1) based on Nashed's definition from [27], which we formulate in the following as Definition 3.1 for the simplified case of an injective bounded linear operator. Moreover, we draw a connecting line to the phenomenon 5 of conditional well-posedness characterized by conditional stability esti- mates, which yield for appropriate choices of the regularization parameter convergence rates in Tikhonov-type regularization. Definition 3.1. The operator equation Ax = y with an injective bounded linear operator A : X → Y mapping between infinite dimensional Banach spaces X and Y is called well-posed if the range R(A) of A is a closed In the ill-posed subset of Y , otherwise the equation is called ill-posed. case, we call the equation ill-posed of type I if R(A) contains an infinite dimensional closed subspace and otherwise ill-posed of type II. The following proposition taken from [13, Proposition 4.2 and 4.4] and the associated Figure 1 gives some more insight into the different situations distinguished in Definition 3.1. Y Proposition 3.2. Consider the operator equation Ax = y from Defini- tion 3.1. If this equation is well-posed, i.e., R(A) = R(A) and there is some constant c > 0 such that (cid:107)Ax(cid:107) ≥ c(cid:107)x(cid:107) for all x ∈ X or the equation is ill-posed of type I, then the operator A is non-compact. Consequently, compactness of A implies ill-posedness of type II. More precisely, for an ill-posed equation Ax = y with injective A and infinite dimensional Ba- nach spaces X and Y , ill-posedness of type II occurs if and only if A is strictly singular. This means that the restriction of A to an infinite dimensional subspace of X is never an isomorphism (linear homeomor- phism). If X and Y are both Hilbert spaces and the equation is ill-posed, then ill-posedness of type II occurs if and only if A is compact. well-posed o p e r a t o r n o t s t r i c t l y s i n g u l a r ill-posed of type I ill-posed of type II operator strictly singular operator compact Figure 1: Properties of A for well-posedness and ill-posedness types of equations from Definition 3.1. Now we apply the case distinction of Definition 3.1, verified in detail in Proposition 3.2, to our situation of equation (1) with X := (cid:96)1 and A : (cid:96)1 → Y . We start with a general observation in Proposition 3.3, which motivates the use of (cid:96)1-regularization for the stable approximate solution of (1), because the equation is mostly ill-posed. Below we enlighten the cross connections a bit more by the discussion of some example situations. Proposition 3.3. If Y is a reflexive Banach space, then the operator equation (1) is always ill-posed of type II. Proof. As a consequence of the theorem from [18] we have that every bounded linear operator A : (cid:96)1 → Y is strictly singular if Y is a reflexive 6 Banach space. Hence well-posedness and ill-posedness of type I cannot occur in such case. Example 3.4. Consider that, as mentioned before in Section 2, we have a composition A = A ◦ L with forward operator A : X → Y for reflexive Y and synthesis operator L : (cid:96)1 → X. Then (1) is ill-posed of type II even if A is continuously invertible and hence the equation Ax = y well- posed. This may occur, for example, for Fredholm or Volterra integral equations of the second kind. Similarly, if A as mapping between Hilbert spaces is non-compact with non-closed range and hence Ax = y is ill- posed of type I (which occurs, e.g., for multiplication operators mapping in L2(0, 1)), (1) is still ill-posed of type II . In the frequent case that X is a separable Hilbert space and {u(k)}k∈N an orthonormal basis, then A is compact whenever A : X → Y is compact (occurring for example for Fredholm or Volterra integral equations of the first kind). Example 3.5. If A := Eq with 1 ≤ q < ∞ and Y := (cid:96)q is the embedding operator, then solving equation (1) based on noisy data yδ ∈ (cid:96)q fulfilling (2) is a denoising problem (see also [13, Sect. 5] and [14, Example 6.1]). For 1 < q < ∞ the embedding operator A = Eq is strictly singular with non-closed range but non-compact. Due to Proposition 3.3 the equation is ill-posed of type II. Moreover, we have Ae(k) (cid:42) 0 in (cid:96)q, which due to Proposition 2.4 implies that A is weak∗-to-weak continuous and hence R(A∗) ⊆ c0. The latter is obvious, because the adjoint A∗ is the embed- ding operator from (cid:96)q∗ . In particular, the source condition (9) applies with f (k) = e(k) ∈ (cid:96)q∗ ⊂ c0 for all k ∈ N. Example 3.6. For q = 1 in the previous example we have the contin- uously invertible identity operator A = Id : (cid:96)1 → (cid:96)1 with closed range R(A) = (cid:96)1. Then equation (1) is well-posed, but we have Ae(k) (cid:54)(cid:42) 0 in (cid:96)1 for k → ∞, which due to Proposition 2.4 indicates that the range R(A∗) of the adjoint of A does not belong to c0 and in particular A is not weak∗- to-weak continuous. This is evident, because the adjoint of A = Id is the identity A∗ = Id : (cid:96)∞ → (cid:96)∞ and R(A∗) = (cid:96)∞. We will come back to this example later. to (cid:96)∞ with 1/q + 1/q∗ = 1 and R(A∗) = (cid:96)q∗ For obtaining error estimates in (cid:96)1-regularization on which convergence rates are based, we need some kind of conditional well-posedness in order to overcome the ill-posedness of equation (1). Well-posed varieties of equation (1) yield stability estimates (cid:107)x − x†(cid:107)(cid:96)1 ≤ K(cid:107)Ax − Ax†(cid:107)Y for all x ∈ (cid:96)1, which under (2) and for the choice α = αSDP imply the best possible rate (cid:107)xδ α − x †(cid:107)(cid:96)1 = O(δ) δ → 0 , as (11) which is typical for well-posed situations. We will come back to this in Section 6. We say that a conditional stability estimate holds true if there is a subset M ⊂ (cid:96)1 such that (cid:107)x − x †(cid:107)(cid:96)1 ≤ K(M)(cid:107)Ax − Ax †(cid:107)Y (12) Because M is not known a priori, such kind of stability requires the ad- ditional use of regularization for bringing the approximate solutions to for all x ∈ M . 7 M such that a rate (11) can be verified. This idea was first published in [7] by Cheng and Yamamoto. In the context of (cid:96)1-regularization for our equation (1), we have estimates of the form (12) if the solution x† ∈ (cid:96)0 is sparse, i.e. only a finite number of non-zero components occur in the infinite sequence x†. Then M can be considered as a subset of (cid:96)0 with specific properties, and the sparsity of (cid:96)1-regularized solutions verified in Proposition 2.1 ensures that the corresponding approximate solutions be- long to M. This implies the rate (11) for x† ∈ (cid:96)0, although equation (1) is not well-posed. A similar but different kind of conditional well-posedness estimates are †(cid:107)(cid:96)1 ≤ (cid:107)x(cid:107)(cid:96)1 − (cid:107)x variational source conditions, which attain in our setting the form β (cid:107)x − x for all x ∈ (cid:96)1 , (13) satisfied for a constant 0 < β ≤ 1 and some concave index function ϕ. From [24, Theorems 1 and 2] we find directly the convergence rates results of the subsequent proposition. †(cid:107)(cid:96)1 + ϕ((cid:107)Ax − Ax †(cid:107)Y ) Proposition 3.7. If the variational source condition (13) holds true for a constant 0 < β ≤ 1 and some concave index function ϕ, then we have for (cid:96)1-regularized solutions xδ α the convergence rate †(cid:107)(cid:96)1 = O(ϕ(δ)) as δ → 0 (cid:107)xδ α − x (14) whenever the regularization parameter is chosen either a priori α = αAP RI according to (6) or a posteriori as α = αSDP according to (7). Consequently, for the manifestation of convergence rates results in the next section it remains to find constants β, concave index functions ϕ and sufficient conditions for the verification of corresponding variational inequalities (13). 4 Convergence rates for (cid:96)1-regularization The first step to derive a variational source condition (13) at the solution 2, ...) ∈ (cid:96)1 was taken by Lemma 5.1 in [5], where the † point x† = (x inequality † 1, x (cid:107)x − x †(cid:107)(cid:96)1 ≤ (cid:107)x(cid:107)(cid:96)1 − (cid:107)x †(cid:107)(cid:96)1 + 2 x k + † xk − x k † (15) (cid:32) ∞(cid:88) n(cid:88) k=n+1 k=1 (cid:33) was proven for all x = (x1, x2, ...) ∈ (cid:96)1 and all n ∈ N. Then under the source condition (9) ,valid for all k ∈ N, one directly finds xk − x k = † (cid:104)e(k), x − x †(cid:105)(cid:96)∞×(cid:96)1 = (cid:104)f (k), A(x − x † )(cid:105)Y ∗×Y (16) k=1 k=1 k=1 n(cid:88) n(cid:88) n(cid:88) and hence from (15) that a function of type ϕ(t) = 2 inf n∈N x k + γn t † (cid:33) (17) (cid:32) ∞(cid:88) k=n+1 8 with β = 1 and n(cid:88) k=1 γn = (cid:107)f (k)(cid:107)Y ∗ (18) provides us with a variational inequality (13). Along the lines of the proof of [5, Theorem 5.2] one can show the assertion of the following lemma. Lemma 4.1. If {γn}n∈N is a non-decreasing sequence, then ϕ from (17) is a well-defined and concave index function for all x† ∈ (cid:96)1. Both the decay rate of x k → 0 as k → ∞ and the behaviour of γn as † n → ∞ in (17) have impact on the resulting rate function ϕ. A power- † k leads to Holder convergence rates (see [5, Example 5.3] type decay of x † and [13, Example 3.4]), whereas exponential decay of x k leads to near-to-δ rates slowed down by a logarithmic factor (see [3, Example 3.5] and [13, In the case that x† ∈ (cid:96)0 is sparse with x † k = 0 for all Example 3.5]). k > n0, then the best possible rate (11) is seen. This becomes clear from formula (17), because then ϕ fulfills the inequality ϕ(t) ≤ 2γn0 t. From Proposition 3.7 we have that for all concave index functions ϕ from (17) a convergence rate (14) for the (cid:96)1-regularization takes place in the case of appropriate choices of the regularization parameter α whenever a constant 0 < β ≤ 1 exists such that (13) is valid with ϕ from (17). When the condition (9) is valid, this is the case with β = 1 and γn from (18). Under the same condition the rate was slightly improved in [13] (see also [14]) by showing that γn from (18) can be replaced with (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n(cid:88) k=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y ∗ γn = sup ak∈{−1,0,1} k=1,...,n akf (k) . (19) However, the condition (9) may fail as was noticed first in [12] for a bidiagonal operator. Therefore, assumption (9) was replaced by a weaker (but not particularly eye-pleasing) one in [12]. Ibid the authors assume, in principle, that for each n ∈ N there are elements f (n,k) such that for all 1 ≤ i ≤ n f (n,k)]i = [e(k)]i and (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n(cid:88) k=1 ∗ [A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c ∗ [A f (n,k)]i for all i > n and c < 1. This means that each basis vector e(k) can be approximated exactly up to arbitrary position but with a non-zero tail consisting of sufficiently small elements. Later, in [14], a more clearly formulated property was assumed which implies the one from [12]. We give a slightly reformulated version of this property in the following. In this context, we notice that Pn denotes the projection operator applied to elements x = (x1, x2, ..., xn, xn+1, ...) such that Pnx = (x1, x2, ..., xn, 0, 0, ...). Property 4.2. For arbitrary µ ∈ [0, 1), we have a real sequence {γn}n∈N such that for each n ∈ N and each ξ = ξ(n) ∈ (cid:96)∞, with if k ≤ n, if k > n , (20) ξk (cid:40)∈ [−1, 1], = 0, 9 there exists some η = η(µ, n, ξ) ∈ Y ∗ satisfying (a) PnA∗η = ξ, [(I − Pn)A∗η]k ≤ µ for all k > n, (cid:107)η(cid:107)Y ∗ ≤ γn. (b) (c) It is important to note that it was a substantial breakthrough in the recent paper [11] to show that Property 4.2 follows directly from injectivity and weak∗-to-weak continuity of the operator A. Namely, the following proposition was proven there. Note that we changed the definition of the ξ in (20) slightly. By checking the proofs in the original paper one sees the amendments we made are not relevant. Proposition 4.3. Let A : (cid:96)1 → Y be bounded, linear and weak∗-to-weak continuous. Then the following assertions are equivalent. (i) A is injective, e(k) ∈ R(A∗) (cid:96)∞ for all k ∈ N, (ii) (iii) R(A∗) (iv) (cid:96)∞ = c0, Property 4.2 holds. In other words, for such operators there exist appropriate sequences {γn}n∈N occurring in (17) such that a variational source condition (13) holds for an index function ϕ from (17) and constant β = 1−µ 1+µ (see Propo- sition 5.5 below). Item (b) in Property 4.2 is a generalization of (9). Namely, the canonical basis vectors e(k) do not necessarily belong to the range of A∗ but to its closure. For the proof of Proposition 4.3 we refer to [11]. Most of the steps are identical or at least similar to the proof of Proposition 5.4 which we will give later. 5 Non-reflexive image spaces If the injective bounded linear operator A : (cid:96)1 → Y fails to be weak∗-to- weak continuous, then the results of the preceding section do not apply. In case that Y is a non-reflexive Banach space, it makes sense to consider the weaker property of weak∗-to-weak∗ continuity of A. An already men- tioned example is the identity mapping A = Id for Y = (cid:96)1. In (cid:96)1, weak convergence and norm convergence coincide (Schur property), but there is no coincidence with weak∗ convergence. Thus, the identity mapping cannot be weak∗-to-weak continuous, but it is weak∗-to-weak∗ continu- ous as the following Proposition 5.1 shows. It is a modified extension of Proposition 2.4. Following [9, Lemma 6.5] we formulate this extension and repeat below the relevant proof details. Proposition 5.1. Let Z be a separable Banach space which acts as a predual space for the Banach space Y = Z∗. Then the following four assertions are equivalent. {Ae(k)}k∈N converges in Y weakly∗ to zero , (i) (ii) R(A∗Z ) := {v ∈ (cid:96)∞ : v = A∗z for some z ∈ Z ⊆ Y ∗} ⊆ c0 , (iii) A is weak∗-to-weak∗ continuous , 10 Z ⊆ Y ∗ Y = Z∗ S c0 A = S∗ (cid:96)1 = (c0)∗ Y ∗ = Z∗∗ A∗ = S∗∗ (cid:96)∞ = ((cid:96)1)∗ = (c0)∗∗ Figure 2: Schematic display of the operators and underlying spaces needed in this section. Top and bottom row: the (separable) Banach spaces under consideration. Middle row: the operators we work with. There is a bounded linear operator S : Z → c0 such that A = S∗ . (iv) Proof. Let (i) be satisfied. Then for each A∗z from R(A∗Z ) we have z, e(k)(cid:105)(cid:96)∞×(cid:96)1 = (cid:104)z, Ae(k)(cid:105)Y ∗×Y = (cid:104)Ae(k), z(cid:105)Z∗×Z → 0 z]k = (cid:104)A ∗ ∗ [A as k → ∞. This yields A∗z ∈ c0 and hence (ii) is valid. Now let (ii) be true and take a weakly∗ convergent sequence xn (cid:42)∗ x0 in (cid:96)1 as n → ∞. Then (cid:104)Axn, z(cid:105)Z∗×Z = (cid:104)z, Axn(cid:105)Y ∗×Y = (cid:104)A∗z, xn(cid:105)(cid:96)∞×(cid:96)1 for all z in Z. Because moreover A∗z belongs to c0 and (cid:96)1 is the dual of c0, we may write this as (cid:104)A∗z, xn(cid:105)(cid:96)∞×(cid:96)1 = (cid:104)xn, A∗z(cid:105)(cid:96)1×c0 . Thus, z(cid:105)(cid:96)1×c0 n→∞(cid:104)Axn, z(cid:105)Z∗×Z = lim lim n→∞(cid:104)xn, A ∗ = (cid:104)Ax0, z(cid:105)Z∗×Z z(cid:105)(cid:96)1×c0 = (cid:104)x0, A ∗ for all z ∈ Z, which proves condition (iii). From (iii) and the fact that e(k) (cid:42)∗ 0 in (cid:96)1 as k → ∞ we immediately obtain (i). Finally, the equivalence between (iii) and (iv) can be found, e.g., in [26, Theorem 3.1.11]. As a consequence of item (iv) in Proposition 5.1, each weak∗-to-weak∗ continuous linear operator is automatically bounded. Figure 2 illustrates the connection between the different spaces and operators we juggle around in this section. For the identity mapping A = Id : (cid:96)1 → Y with Y = (cid:96)1 and predual Z = c0, property (i) of Proposition 5.1 is trivially satisfied which yields the weak∗-to-weak∗ continuity of this operator. Note that the case Y = (cid:96)1, A = Id is only of theoretical interest. Precisely, it is a tool for exploring the frontiers of the theoretic framework we have chosen for investigating (cid:96)1-regularization. For practical applications it is irrelevant because one easily verifies that with the choice p = 1 in (3), where we have Y = (cid:96)1, the (cid:96)1-regularized solutions coincide with the data yδ if α < 1 and we have the best possible rate (11). Main parts of the above mentioned Proposition 2.1 on existence, sta- bility and convergence of (cid:96)1-regularized solutions xδ α remain true if the operator A : (cid:96)1 → Y is only weak∗-to-weak∗ continuous. The sparsity α ∈ (cid:96)0, however, will fail in general (consider the example of property xδ the identity as mentioned above). Existence, stability and convergence 11 assertions remain valid, because their proofs basically rely on the fact that the mapping x (cid:55)→ (cid:107)Ax − yδ(cid:107)Y is a weakly∗ lower semicontinuous functional. This is the case in both variants, with or without ∗, since the norm functional is weakly and also weakly∗ lower semicontinuous. For the existence of regularized solutions (minimizers of the Tikhonov func- tional (3)) again the Banach-Alaoglu theorem (Lemma 2.2) is required and yields weakly∗ compact level sets of the (cid:96)1-norm functional. Our goal is to proof an analogue to Proposition 4.3 for weak∗-to-weak∗ continuous operators. We start with a first observation. Proposition 5.2. Let A : (cid:96)1 → Y be injective and weak∗-to-weak∗ con- tinuous and let Y = Z∗ for some Banach space Z. Then R(A∗Z ) (cid:96)∞ = c0. Proof. From item (iv) of Proposition 5.1 we take the operator S : Z → c0 with A = S∗. As A is injective, i.e., N (A) = {0}, it follows R(S) c0 = N (S ∗ )⊥ = N (A)⊥ = c0. There, the subscript denotes the pre-annihilator for a set V , in our situa- tion with (c0)∗ = (cid:96)1 and V ⊆ (cid:96)1 defined as V⊥ := {x ∈ c0 : (cid:104)ζ, x(cid:105)(cid:96)1×c0 = 0 ∀ζ ∈ V }. Let η ∈ Z and recall Y ∗ = Z∗∗ (cf. Figure 2). Then for each x ∈ (cid:96)1 η, x(cid:105)(cid:96)∞×(cid:96)1 = (cid:104)η, A x(cid:105)Y ∗×Y = (cid:104)η, A x(cid:105)Z×Y = (cid:104)S η, x(cid:105)c0×(cid:96)1 (cid:104)A ∗ = (cid:104)S η, x(cid:105)(cid:96)∞×(cid:96)1 , i.e., A∗Z = S. Thus R(A∗Z ) c0 = c0. At this point we emphasize that in both Banach spaces (cid:96)∞ and c0 the same supremum norm applies. = R(S) (cid:96)∞ We will show in Proposition 5.4 that conversely R(A∗Z ) = c0 im- plies injectivity for weak∗-to-weak∗ continuous operators. Before doing so we need the following Proposition which coincides in principle with [11, Proposition 9]. Proposition 5.3. Let A be injective and weak∗-to-weak∗ continuous. Moreover, let ε > 0 and n ∈ N. Then for each ξ ∈ c0 there exists ξ ∈ R(A∗) such that (cid:96)∞ ξk = ξk for k ≤ n and ξk − ξk ≤ ε for k > n. Proof. We proof the proposition by induction with respect to n. For ξ ∈ c0 set ξ+ := (ξ1 + ε, ξ2, ξ3, . . .) and − ξ By Proposition 5.2 we have that c0 = R(A∗Z ) find elements ξ+ ∈ R(A∗) and ξ− ∈ R(A∗) with − − ξ (cid:107) ξ+ − ξ+(cid:107)(cid:96)∞ ≤ ε (cid:107) ξ and −(cid:107)(cid:96)∞ ≤ ε. := (ξ1 − ε, ξ2, ξ3, . . .). (cid:96)∞ ⊂ R(A∗) (cid:96)∞ . Hence we 12 1 ≥ ξ1 ≥ ξ 1 and ξ+ − k − ξk ≤ ε as well as ξ k − ξk ≤ ε − Consequently, ξ+ for k > 1. Thus we find a convex combination ξ of ξ+ and ξ− such that ξ1 = ξ1. This ξ obviously also satisfies ξk − ξk ≤ ε for k > 1, which proves the proposition for n = 1. Let ξ ∈ c0 and set Now let the proposition be true for n = m. We prove it for n = m + 1. ξ+ := (ξ1, . . . , ξm, ξm+1 + ε, ξm+2, ξm+3, . . .), := (ξ1, . . . , ξm, ξm+1 − ε, ξm+2, ξm+3, . . .). − ξ By the induction hypothesis we find ξ+ ∈ R(A∗) and ξ− ∈ R(A∗) with − k = ξk = ξ ξ+ k for k ≤ m and k − ξ+ ξ+ k ≤ ε k − ξ k ≤ ε and ξ − − m+1 ≥ ξm+1 ≥ ξ m+1 and ξ+ − for k > m. k − ξk ≤ ε as well as Consequently, we have ξ+ ξ k − ξk ≤ ε for k > m + 1. Thus we find a convex combination ξ of ξ+ − and ξ− such that ξm+1 = ξm+1. This ξ obviously also satisfies ξk = ξk for k < m + 1 and ξk − ξk ≤ ε for k > m + 1, which proves the proposition for n = m + 1. Now we come to the main result of this section. The proof is similar and in part identical to the one of Proposition 12 in [11]. Proposition 5.4. Let A : (cid:96)1 → Y be bounded, linear and weak∗-to-weak∗ continuous. Then the following assertions are equivalent. A is injective, (i) (ii) R(A∗Z ) (cid:96)∞ = c0, (cid:96)∞ e(k) ∈ R(A∗Z ) Property 4.2 holds. (iii) for all k ∈ N, (iv) Proof. We show (i) ⇒ (iv) ⇒ (iii) ⇒ (ii) ⇒ (i). (i)⇒(iv): Fix µ ∈ (0, 1), n ∈ N and take some ξ as described in Property 4.2. By Proposition 5.3 with ε := µ there exists some η such that A∗η (= ξ in the proposition) satisfies items (a) and (b) in Property In particular we have {ηk}k∈1,...,n such that PnA∗ηk = e(k) and 4.2. [(I − Pn)A∗ηk]i ≤ µ n for all i > n. Since ξ ∈ c0 it is n(cid:88) (cid:33) for coefficients −1 ≤ ck ≤ 1, i.e., ξ = A∗η with η ≤ n(cid:80) (cid:32) n(cid:88) ηk as an ∗ ckA ∗ ηk = A n(cid:88) cke(k) = ckηk , ξ = k=1 k=1 k=1 upper bound for γn. By construction this η also fulfills k=1 [(I − Pn)A ∗ [(I − Pn)A ∗ ηk]i ≤ µ. η]i ≤ n(cid:88) i=1 13 (iv)⇒(iii): Fix k, fix n ≥ k, take a sequence (µm)m∈N in (0, 1) with µm → 0 and choose ξ := e(k) in Property 4.2. Then for a corresponding sequence (ηm)m∈N from Property 4.2 we obtain (cid:107)e(k) − A ∗ ηm(cid:107)(cid:96)∞ ≤ (cid:107)e(k) − PnA ∗ ηm(cid:107)(cid:96)∞ + (cid:107)(I − Pn)A ∗ ηm(cid:107)(cid:96)∞ . The first summand is zero by the choice of ξ and the second summand is bounded by µm. Thus, (cid:107)e(k) − A∗ηm(cid:107)(cid:96)∞ → 0 if m → ∞. (iii)⇒(ii): (e(k))k∈N is a Schauder basis in c0. Thus, c0 ⊆ R(A∗Z ) (cid:96)∞ . Proposition 5.1 yields R(A∗Z ) (ii)⇒(i): One easily shows R(A∗) (cid:96)∞ ⊆ c0. Hence R(A∗Z ) = c0. (cid:96)∞ ⊆ N (A)⊥. Thus c0 ⊆ N (A)⊥. If we have x ∈ (cid:96)1 with Ax = 0, then for each u ∈ c0 ⊆ N (A)⊥ we obtain (cid:104)x, u(cid:105)(cid:96)1×c0 = (cid:104)u, x(cid:105)(cid:96)∞×(cid:96)1 = 0, which is equivalent to x = 0. Since, in the context of both Propositions 4.3 and 5.4, the injectivity of A yields Property 4.2, the consequences with respect to variational source conditions and convergence rate results are identical for a weak∗- to-weak and a weak∗-to-weak∗ continuous operator A. We formulate the following theorem and the subsequent corollary and prove the theorem for a weak∗-to-weak∗ continuous operator A : (cid:96)1 → Y . In particular, the corollary requires the existence of a separable predual space Z of Y in order to apply Lemma 2.2 and to ensure the stabilizing property of the Tikhonov penalty. However, the proof of the theorem repeats point by point the ideas of the proof from [11, Corollary 11] focused on weak∗-to- weak continuous operators A. Theorem 5.5. Let the bounded linear operator A : (cid:96)1 → Y be injective and weak∗-to-weak∗ continuous, and assume additionally that the Banach space Y possesses a separable predual Banach space Z with Z∗ = Y . More- over, let µ ∈ [0, 1) and {γn}n∈N be such that Property 4.2 is fulfilled. Then 1+µ ∈ [0, 1) and a variational source condition (13) with the constant β = 1−µ the concave index function ϕ given by (17) is fulfilled. Proof. Fix n ∈ N and x ∈ (cid:96)1 and let ξ := sgn Pn(x − x†) ∈ (cid:96)∞ be the sequence of signs of Pn(x − x†). Then by Property 4.2 there is some η such that η, x − x xk − x k = (cid:104)ξ, x − x † = (cid:104)PnA η − A η, x − x ∗ ∗ = −(cid:104)(I − Pn)A η, (I − Pn)(x − x ∗ ≤ µ(cid:107)(I − Pn)(x − x † †(cid:105)(cid:96)∞×(cid:96)1 = (cid:104)PnA ∗ †(cid:105)(cid:96)∞×(cid:96)1 + (cid:104)A ∗ )(cid:105)(cid:96)∞×(cid:96)1 + (cid:104)A ∗ )(cid:107)(cid:96)1 + γn(cid:107)Ax − Ax η, x − x †(cid:107)Y . †(cid:105)(cid:96)∞×(cid:96)1 †(cid:105)(cid:96)∞×(cid:96)1 † η, x − x †(cid:105)(cid:96)∞×(cid:96)1 (21) The triangle inequality yields (cid:107)Pn(x−x )(cid:107)(cid:96)1 ≤ µ(cid:0)(cid:107)(I−Pn)x(cid:107)(cid:96)1 +(cid:107)(I−Pn)x † (cid:1)+γn(cid:107)Ax−Ax †(cid:107)Y . (22) †(cid:107)(cid:96)1 14 n(cid:88) k=1 Now β(cid:107)x − x = β(cid:107)Pn(x − x † †(cid:107)(cid:96)1 − (cid:107)x(cid:107)(cid:96)1 + (cid:107)x †(cid:107)(cid:96)1 + (cid:107)Pnx †(cid:107)(cid:96)1 + (cid:107)(I − Pn)x †(cid:107)(cid:96)1 )(cid:107)(cid:96)1 + β(cid:107)(I − Pn)(x − x † )(cid:107)(cid:96)1 − (cid:107)Pnx(cid:107)(cid:96)1 − (cid:107)(I − Pn)x(cid:107)(cid:96)1 together with β(cid:107)(I − Pn)(x − x † )(cid:107)(cid:96)1 ≤ β(cid:107)(I − Pn)x(cid:107)(cid:96)1 + β(cid:107)(I − Pn)x †(cid:107)(cid:96)1 and shows (cid:107)Pnx †(cid:107)(cid:96)1 = (cid:107)Pn(x − x † − x)(cid:107)(cid:96)1 ≤ (cid:107)Pn(x − x † )(cid:107)(cid:96)1 + (cid:107)Pnx(cid:107)(cid:96)1 β(cid:107)x − x †(cid:107)(cid:96)1 − (cid:107)x(cid:107)(cid:96)1 + (cid:107)x †(cid:107)(cid:96)1 ≤ 2(cid:107)(I − Pn)x − (1 − β)(cid:0)(cid:107)(I − Pn)x(cid:107)(cid:96)1 + (cid:107)(I − Pn)x †(cid:107)(cid:96)1 + (1 + β)(cid:107)Pn(x − x )(cid:107)(cid:96)1 †(cid:107)(cid:96)1 † (cid:1). Combining this estimate with the previous estimate (22) and taking into account that β = 1−µ 1+β we obtain for all x ∈ (cid:96)1 1+µ and µ = 1−β β(cid:107)x − x †(cid:107)(cid:96)1 − (cid:107)x(cid:107)(cid:96)1 + (cid:107)x †(cid:107)(cid:96)1 ≤ 2(cid:107)(I − Pn)x ≤ 2(cid:107)(I − Pn)x 2 †(cid:107)(cid:96)1 + †(cid:107)(cid:96)1 + 2γn(cid:107)Ax − Ax γn(cid:107)Ax − Ax †(cid:107)Y . 1 + µ †(cid:107)Y Taking the infimum over all n ∈ N completes the proof. The variational source condition immediately yields convergence rates according to Proposition 3.7. Corollary 5.6. Under the conditions of Theorem 5.5 the (cid:96)1-regularized solutions xδ α as minimizers of (3) fulfil (cid:107)xδ α − x †(cid:107)(cid:96)1 = O(ϕ(δ)) as δ → 0 , with the concave index function ϕ from (17), whenever the regularization parameter α is chosen either a priori as α = αAP RI according to (6) or a posteriori as α = αSDP according to (7). In order to familiarize the reader with the concepts in this work we will look at a particular operator to exemplify our theory. In particular we verify Property 4.2. Example 5.7. Let X = Y = (cid:96)1 and [Ax]k = xk + xk+1 k ∈ N, i.e., A maps x = (x1, x2, x3, . . . ) to Ax = (x1 + x2, x2 + x3, x3 + x4, . . . ). Clearly A is linear. Observe that Ax(cid:96)1 ≤ 2x(cid:96)1 and hence A is bounded. One easily verifies the adjoint A∗ : (cid:96)∞ → (cid:96)∞, ∗ A y = (y1, y2 + y1, y3 + y2, y3 + y4, . . . ). 15 Both A and A∗ are injective. Since R(A) (cid:96)1 = N (A∗)⊥, where N (A ∗ )⊥ := {x ∈ (cid:96)1 : (cid:104)y, x(cid:105)(cid:96)∞×(cid:96)1 = 0 ∀y ∈ N (A ∗ )} = (cid:96)1 (cid:96)1 = (cid:96)1. It is however easy to see that R(A) (cid:54)= (cid:96)1. For we have R(A) example there is no x ∈ (cid:96)1 such that Ax = e(2). Namely, solving Ax = e(2) for x leads to the system of equations x1 = −x2, x2 = 1 − x3, x4 = −x3, x5 = −x4, etc. Due to the alternating character again there is no x ∈ (cid:96)1 that satisfies this system. We have shown that R(A) (cid:96)1 (cid:54)= R(A), i.e., the corresponding operator equation (1) is ill-posed. Next we prove that A is weak∗-to-weak∗ continuous but not weak∗-to- weak continuous. To this end we use the properties (i) in Proposition 2.4 and Proposition 5.1, respectively. First let ξ ∈ c0. Then, with (cid:40) [Ae(k)]i = 1 i = k, k − 1 0 else ∀i ≥ 2 it is (cid:104)ξ, Ae(k)(cid:105)c0×(cid:96)1 = ξk−1 + ξk → 0 , since ξ ∈ c0. For ξ ∈ (cid:96)∞, however, (cid:104)ξ, Ae(k)(cid:105)(cid:96)∞×(cid:96)1 = ξk−1 + ξk does in general not converge to zero (let, e.g., ξ ≡ 1). Summarizing the properties of the forward operator for the present example, we note that A is linear, injective, weak∗-to-weak∗ continuous, but not weak∗-to-weak continuous. Moreover its range is not closed such that the corresponding operator equation (1) is ill-posed. For this particular operator let us investigate Property 4.2. We will see in the following that this actually holds with µ = 0 and γn = n, i.e., (cid:96)∞ e(k) ∈ R(A∗) for all k ∈ N. We will also show that e(k) ∈ R(A∗c0 ) according to item (iii) of Proposition 5.4. Even then we still have γn = n for arbitrary 0 < µ < 1. Fix an arbitrary n ∈ N and let ξ = (ξ1, ξ2, . . . , ξn, 0, . . . ) ∈ (cid:96)∞ with ξi ∈ [−1, 1]. We are looking for an η ∈ (cid:96)∞ satisfying Property 4.2. Observe that, by definition of A∗ and for given ξ, any η satisfying PnA∗η = ξ has the structure η1 = ξ1, η2 = ξ2 − η1, η3 = ξ3 − η2 and so on until ηn = ξn − ηn−1. In other words it is ηn =(cid:80)n i=1(−1)n−iξi and ηn(cid:96)∞ ≤ n, which yields item (c) of Property 4.2 with γn = n. Now fix arbitrary 0 < µ < 1. We have [A∗η]n+1 = ηn+1 + ηn and require ηn+1 + ηn ≤ µ. Thus we may take any ηn+1 with −ηn − µ ≤ ηn+1 ≤ −ηn + µ. Analogously we find that in general (−1)iηn − iµ ≤ ηn+i ≤ (−1)iηn + iµ, i = 1, 2, . . . . 16 Therefore, the choice of the tail of η is ambiguous. For example, a viable pick is ηn+i = (−1)iηn. Then η = (η1, η2, . . . , ηn,−ηn, ηn,−ηn, . . . ) (23) with ηi, 1 ≤ i ≤ n, as above and ∗ A η = (ξ1, ξ2, . . . , ξn, 0, 0, 0, . . . ). In particular, this means that e(k) ∈ R(A∗) (choose ξi = 1 and ξj = 0 for i (cid:54)= j). Note that η ∈ (cid:96)∞ but η /∈ c0 in (23). However, we also see that for any ξ and arbitrary 0 < µ < 1 there are (infinitely many) choices for the tail of η such that η ∈ c0 and item (b) of Property 4.2 holds. Independent of µ, all choices satisfy item (c) of Property 4.2 with γn = n. To set this into relation, we would obtain the same γn for a diagonal operator A : (cid:96)2 → (cid:96)2 with singular values decaying as σi ∼ 1√ , see [17]. i Please note that in practice it is not necessary to verify Property 4.2 in the way we did here. In particular the sequential discrepancy principle (7) does not require the knowledge of any of the parameters from Property 4.2 in order to guarantee the convergence rates implied by that property. For the sake of completeness we mention that there exist bounded linear operators which are not even weak∗-to-weak∗ continuous. Example 5.8. If Y = (cid:96)1 and  ∞(cid:80) l=1 xk, [Ax]k := xl, if k = 1, else, for all k in N and all x in (cid:96)1, then Ae(k) = e(1) + e(k) if k > 1. Thus, Ae(k) (cid:42)∗ e(1) (cid:54)= 0. The same operator A considered as mapping into Y = (cid:96)2 is an example of a not weak∗-to-weak continuous bounded linear operator in the classical Hilbert space setting for (cid:96)1-regularization. Note that, because of the first component, A does not have a bounded extension to any (cid:96)p-space with 1 < p < ∞. 6 Well-posedness and further discussions Proposition 6.1. If the problem (1) is well-posed, i.e. R(A) = R(A) , then under the conditions of Theorem 5.5 the (cid:96)1-regularized solutions xδ α as minimizers of (3) fulfil Y (cid:107)xδ α − x †(cid:107)(cid:96)1 = O(δ) as δ → 0 whenever the regularization parameter α > 0 is chosen either a priori as α = αAP RI ∼ δp−1 or a posteriori as α = αSDP according to the sequential discrepancy principle (7). Proof. The condition R(A) = R(A) (cf. [26, Theorem 3.1.21]) and hence V := R(A∗) is a closed subspace of (cid:96)∞, which can be considered as a Banach space with the same supremum norm as implies R(A∗) = R(A∗) (cid:96)∞ Y 17 (cid:96)∞. Then, for the injective operator A : (cid:96)1 → Y , Banach's theorem concerning the continuity of the inverse operator ensures that the linear operator (A∗)−1 : V → Y ∗ is bounded. Moreover, the elements ξ ∈ R(A∗) in Proposition 5.3 associated to ξ from Property 4.2 satisfy the inequality (cid:107) ξ(cid:107)(cid:96)∞ ≤ 1 + ε, and with ξ = A∗η we have (cid:107)η(cid:107)Y ∗ ≤ (cid:107)(A ∗ −1(cid:107)V →Y ∗ (1 + ε) ≤ K < ∞. ) Hence, the sequence {γn}n∈N in Property 4.2 is uniformly bounded by the finite positive constant K. Taking into account the proof of Theorem 5.5 we have with β = 1−µ β(cid:107)x − x 1+µ and for all x ∈ (cid:96)1 †(cid:107)Y , †(cid:107)(cid:96)1 ≤ (cid:107)x(cid:107)(cid:96)1 − (cid:107)x †(cid:107)(cid:96)1 + 2(cid:107)(I − Pn)x i.e. the variational inequality (13) with †(cid:107)(cid:96)1 + 2 K (cid:107)Ax − Ax (cid:33) (cid:32) ∞(cid:88) k=n+1 ϕ(t) = 2 inf n∈N x k + K t † = K t. This, however, yields by Proposition 3.7 the rate (11) and completes the proof of the proposition. Property 4.2 enables us to show convergence rates for (cid:96)1-regularization for ill-posed and well-posed problems with sparse and non-sparse solu- tions. It has been shown in [17] that the rate function ϕ in (14) does in general not saturate. Even more, the rate is always obtainable either with an a priori or an a posteriori choice of the regularization parameter. This stands in sharp contrast to classical Tikhonov regularization, i.e., (3) with p = 2 and x(cid:96)2 as penalty, which is known to admit convergence rates up to a maximum of δ2/3 for a suitable a priori choice of the regulariza- tion parameter and only a rate of δ1/2 under the discrepancy principle. Since the smoothness of the solution is typically unknown, this makes (cid:96)1- regularization more attractive from the viewpoint of regularization theory than its (cid:96)2 counterpart. One simply uses the discrepancy principle and no longer has to care about the smoothness of the solution. However, one may run into trouble when the solution does not belong to (cid:96)1 but only to (cid:96)2\(cid:96)1 such that x†(cid:96)1 = ∞. In such a case we call the regularization method (3) oversmoothing. There are promising results showing that even in the situation of over- smoothing, (cid:96)1-regularization may lead to convergence rates in a weaker norm. Again, an a priori choice or the discrepancy principle for the choice of α would lead to the optimal rates. Preliminary results have been shown in the preprint [15]. There, a strategy is shown to derive convergence rates in the (cid:96)2-norm for (cid:96)1-regularization for every x† ∈ (cid:96)2. The proof of a the- orem analogously to Proposition 3.7 unfortunately was incomplete. It revolves around approximating x† with Pnx† and letting n = n(δ) → ∞ as δ → 0 with a specific choice of n = n(δ). The open problem was to show that the support of the approximate solutions is not larger than this n(δ). It appears that such a statement is possible by using item (c) in Property 4.2 and the necessary optimality condition for a minimizer of (3), where the latter provides us with the norm ηY ∗ in Property 1 18 corresponding to a ξ = A∗η ∈ ∂xδ α(cid:96)1 . Since we are not able to use a variational source condition when x†(cid:96)1 = ∞ we need to use a different approach to show convergence rates. To this end we seem to have a chance to adapt the strategy of [17] based on elementary steps. Consequently, we hope to complete the detailed proof in an upcoming paper [16]. Acknowledgements The authors were financially supported by Deutsche Forschungsgemein- schaft (DFG-grant HO 1454/10-1). References [1] S. W. Anzengruber, B. Hofmann, and P. Math´e. Regularization prop- erties of the sequential discrepancy principle for Tikhonov regular- ization in Banach spaces. Appl. Anal., 93:1382 -- 1400, 2014. [2] S. W. Anzengruber, B. Hofmann, and R. Ramlau. On the interplay of basis smoothness and specific range conditions occurring in sparsity regularization. Inverse Problems, 29:125002 (21pp), 2013. [3] R. I. Bot¸ and B. Hofmann. The impact of a curious type of smooth- ness conditions on convergence rates in (cid:96)1-regularization. Eurasian Journal of Mathematical and Computer Applications, 1:29 -- 40, 2013. [4] K. Bredies and D. A. Lorenz. Regularization with non-convex sepa- rable constraints. Inverse Problems, 25:085011 (14pp), 2009. [5] M. Burger, J. Flemming, and B. Hofmann. Convergence rates in (cid:96)1-regularization if the sparsity assumption fails. Inverse Problems, 29:025013 (16pp), 2013. [6] D. Chen, B. Hofmann, and J. Zou. Elastic-net regularization versus (cid:96)1-regularization for linear inverse problems with quasi-sparse solu- tions. Inverse Problems, 33(1):015004, 17, 2017. [7] J. Cheng and M. Yamamoto. One new strategy for a priori choice of regularizing parameters in Tikhonov's regularization. Inverse Prob- lems, 16(4):L31 -- L38, 2000. [8] I. Daubechies, M. Defrise, and C. De Mol. An iterative threshold- ing algorithm for linear inverse problems with a sparsity constraint. Comm. Pure Appl. Math., 57(11):1413 -- 1457, 2004. [9] J. Flemming. Quadratic Inverse Problems and Sparsity Promoting Regularization -- two subjects, some links between them, and an ap- plication in laser optics. Habilitation thesis, Technische Universitat Chemnitz, Germany, 2017. [10] J. Flemming. injectivity-type assumptions. 2016. Convergence rates for (cid:96)1-regularization without Inverse Problems, 32(9):095001, 19, [11] J. Flemming and D. Gerth. Injectivity and weak∗-to-weak continuity suffice for convergence rates in (cid:96)1-regularization. J. Inv. Ill-Posed 19 Probl., 2017. published ahead of print https://doi.org/10.1515/jiip- 2017-0008. [12] J. Flemming and M. Hegland. Convergence rates in (cid:96)1-regularization when the basis is not smooth enough. Appl. Anal., 94:464 -- 476, 2015. [13] J. Flemming, B. Hofmann, and I. Veseli´c. On (cid:96)1-regularization in light of Nashed's ill-posedness concept. Comput. Methods Appl. Math., 15(3):279 -- 289, 2015. [14] J. Flemming, B. Hofmann, and I. Veseli´c. A unified approach to con- vergence rates for (cid:96)1-regularization and lacking sparsity. J. Inverse Ill-Posed Probl., 24(2):139 -- 148, 2016. [15] D. Gerth. Tikhonov regularization with oversmoothing available at https://www.tu- penalties. chemnitz.de/mathematik/preprint/2016/PREPRINT 08.pdf. Preprint, 2016. [16] D. Gerth. Convergence rates for (cid:96)1-regularization when the exact solution is not in (cid:96)1. Article in preparation, 2017. [17] D. Gerth. Convergence rates for (cid:96)1-regularization without the help of a variational inequality. Electron. Trans. Numer. Anal., 46:233 -- 244, 2017. [18] S. Goldberg and E Thorp. On some open questions concerning strictly singular operators. Proc. Amer. Math. Soc., 14:334 -- 336, 1963. [19] M. Grasmair. Well-posedness and convergence rates for sparse reg- ularization with sublinear lq penalty term. Inverse Probl. Imaging, 3(3):383 -- 387, 2009. [20] M. Grasmair. Generalized Bregman distances and convergence rates for non-convex regularization methods. Inverse Problems, 26:115014 (16pp), 2010. [21] M. Grasmair, M. Haltmeier, and O. Scherzer. Sparse regularization with lq penalty term. Inverse Problems, 24(5):055020, 13, 2008. [22] M. Grasmair, M. Haltmeier, and O. Scherzer. Necessary and sufficient conditions for linear convergence of (cid:96)1-regularization. Comm. Pure Appl. Math., 64:161 -- 182, 2011. [23] B. Hofmann, B. Kaltenbacher, C. Poschl, and O. Scherzer. A conver- gence rates result for Tikhonov regularization in Banach spaces with non-smooth operators. Inverse Problems, 23:987 -- 1010, 2007. [24] B. Hofmann and P. Math´e. Parameter choice in Banach space regu- larization under variational inequalities. Inverse Problems, 28:104006 (17pp), 2012. [25] D. A. Lorenz. Convergence rates and source conditions for Tikhonov regularization with sparsity constraints. J. Inverse Ill-Posed Probl., 16:463 -- 478, 2008. [26] R. E. Megginson. An Introduction to Banach Space Theory, volume 183 of Graduate Texts in Mathematics. Springer, New York, 1998. 20 [27] M. Z. Nashed. A new approach to classification and regularization of ill-posed operator equations. In Inverse and Ill-posed Problems (Sankt Wolfgang, 1986), volume 4 of Notes Rep. Math. Sci. Engrg., pages 53 -- 75. Academic Press, Boston, MA, 1987. [28] R. Ramlau. Regularization properties of Tikhonov regularization with sparsity constraints. Electron. Trans. Numer. Anal., 30:54 -- 74, 2008. [29] R. Ramlau and E. Resmerita. Convergence rates for regularization with sparsity constraints. Electron. Trans. Numer. Anal., 37:87 -- 104, 2010. [30] W. Rudin. Functional Analysis. International Series in Pure and Applied Mathematics. McGraw-Hill, Inc., New York, second edition, 1991. [31] O. Scherzer, editor. Handbook of Mathematical Methods in Imaging. Springer Science+Business Media LCC, New York, 2nd edition, 2011. [32] O. Scherzer, M. Grasmair, H. Grossauer, M. Haltmeier, and F. Lenzen. Variational Methods in Imaging, volume 167 of Applied Mathematical Sciences. Springer, New York, 2009. [33] T. Schuster, B. Kaltenbacher, B. Hofmann, and K. S. Kazimierski. Regularization Methods in Banach Spaces, volume 10 of Radon Series on Computational and Applied Mathematics. Walter de Gruyter, Berlin/Boston, 2012. 21
1101.4553
2
1101
2011-06-13T16:18:26
Hilbertian Jamison sequences and rigid dynamical systems
[ "math.FA", "math.DS" ]
A strictly increasing sequence (n_k) of positive integers is said to be a Hilbertian Jamison sequence if for any bounded operator T on a separable Hilbert space such that the supremum over k of the norms ||T^{n_k}|| is finite, the set of eigenvalues of modulus 1 of T is at most countable. We first give a complete characterization of such sequences. We then turn to the study of rigidity sequences (n_k) for weakly mixing dynamical systems on measure spaces, and give various conditions, some of which are closely related to the Jamison condition, for a sequence to be a rigidity sequence. We obtain on our way a complete characterization of topological rigidity and uniform rigidity sequences for linear dynamical systems, and we construct in this framework examples of dynamical systems which are both weakly mixing in the measure-theoretic sense and uniformly rigid.
math.FA
math
HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS by Tanja Eisner & Sophie Grivaux Abstract. -- A strictly increasing sequence (nk)k≥0 of positive integers is said to be a Hilbertian Jamison sequence if for any bounded operator T on a separable Hilbert space such that supk≥0 T nk < +∞, the set of eigenvalues of modulus 1 of T is at most countable. We first give a complete characterization of such sequences. We then turn to the study of rigidity sequences (nk)k≥0 for weakly mixing dynamical systems on measure spaces, and give various conditions, some of which are closely related to the Jamison condition, for a sequence to be a rigidity sequence. We obtain on our way a complete characterization of topological rigidity and uniform rigidity sequences for linear dynamical systems, and we construct in this framework examples of dynamical systems which are both weakly mixing in the measure-theoretic sense and uniformly rigid. 1. Introduction and main results We are concerned in this paper with the study of certain dynamical systems, in particular linear dynamical systems. Our main aim is the study of rigidity sequences (nk)k≥0 for weakly mixing dynamical systems on measure spaces, and we present tractable conditions on the sequence (nk)k≥0 which imply that it is (or not) a rigidity sequence. Our conditions on the sequence (nk)k≥0 come in part from the study of the so-called Jamison sequences, which appear in the description of the relationship between partial power-boundedness of an operator on a separable Banach space and the size of its unimodular point spectrum. Let us now describe our results more precisely. 2000 Mathematics Subject Classification. -- 47A10, 37A05, 37A50, 47A10, 47B37. Key words and phrases. -- Linear dynamical systems, partially power-bounded operators, point spec- trum of operators, hypercyclicity, weak mixing and rigid dynamical systems, topologically rigid dynamical systems. This work was partially supported by ANR-Projet Blanc DYNOP, the European Social Fund and the Ministry of Science, Research and the Arts Baden-Wurttemberg. 2 TANJA EISNER & SOPHIE GRIVAUX 1.1. A characterization of Hilbertian Jamison sequences. -- Let X be a separable infinite-dimensional complex Banach space, and let T ∈ B(X) be a bounded operator on X. We are first going to study here the relationship between the behavior of the sequence T n of the norms of the powers of T , and the size of the unimodular point spectrum σp(T ) ∩ T, i.e. the set of eigenvalues of T of modulus 1. It is known since an old result of Jamison [19] that a slow growth of T n makes σp(T ) ∩ T small, and vice-versa. More precisely, the result of [19] states that if T is power-bounded, i.e. supn≥0 T n < +∞, then σp(T )∩T is at most countable. For a sample of the kind of results which can be obtained in the other direction, let us mention the following result of Nikolskii [28]: if T is a bounded operator on a separable Hilbert space such that σp(T )∩ T has positive Lebesgue measure, then the series Pn≥0 T n−2 is convergent. This has been generalized by Ransford in the paper [33], which renewed the interest in these matters. In particular Ransford started to investigate in [33] the influence of partial power-boundedness of an operator on the size of its unimodular point spectrum. Let us recall the following definition: Definition 1.1. -- Let (nk)k≥0 be an increasing sequence of positive integers, and T a bounded linear operator on the space X. We say that T is partially power-bounded with respect to (nk) if supk≥0 T nk < +∞. In view of the result of Jamison, it was natural to investigate whether the partial power- boundedness of T with respect to (nk) implies that σp(T ) ∩ T is at most countable. It for was shown in [34] by Ransford and Roginskaya that it is not the case: instance, there exist a separable Banach space X and T ∈ B(X) such that supk≥0 T nk is finite while σp(T ) ∩ T is uncountable. This question was investigated further in [1] and [2], where the following definition was introduced: if nk = 22k Definition 1.2. -- Let (nk)k≥0 be an increasing sequence of integers. We say that (nk)k≥0 is a Jamison sequence if for any separable Banach space X and any bounded operator T on X, σp(T ) ∩ T is at most countable as soon as T is partially power-bounded with respect to (nk). Whether (nk)k≥0 is a Jamison sequence or not depends of course on features of the sequence such as its growth, its arithmetical properties, etc. A complete characterization of Jamison sequences was obtained in [2]. It is formulated as follows: Theorem 1.3. -- Let (nk)k≥0 be an increasing sequence of integers with n0 = 1. The following assertions are equivalent: (1) (nk)k≥0 is a Jamison sequence; (2) there exists a positive real number ε such that for every λ ∈ T \ {1}, k≥0 λnk − 1 ≥ ε. sup Many examples of Jamison and non-Jamison sequences were obtained in [1] and [2]. Among the examples of non-Jamison sequences, let us mention the sequences (nk)k≥0 such that nk+1/nk tends to infinity, or such that nk divides nk+1 for each k ≥ 0 and lim sup nk+1/nk = +∞. Saying that (nk)k≥0 is not a Jamison sequence means that there exists a separable Banach space X and T ∈ B(X) such that supk≥0 T nk < +∞ and HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 3 in the σp(T ) ∩ T is uncountable. But the space X may well be extremely complicated: proof of Theorem 1.3, the space is obtained by a rather involved renorming of a classi- cal space such as ℓ2 for instance. This is a drawback in applications, and this is why it was investigated in [1] under which conditions on the sequence (nk)k≥0 it was possible to construct partially power-bounded operators with respect to (nk)k≥0 with uncountable It was proved in [1] that if the series unimodular point spectrum on a Hilbert space. Pk≥0(nk/nk+1)2 is convergent, there exists a bounded operator T on a separable Hilbert space H such that supk≥0 T nk < +∞ and σp(T ) ∩ T is uncountable. But this left open the characterization of Hilbertian Jamison sequences. Definition 1.4. -- We say that (nk)k≥1 is a Hilbertian Jamison sequence if for any bounded operator T on a separable infinite-dimensional complex Hilbert space which is partially power-bounded with respect to (nk), σp(T ) ∩ T is at most countable. Obviously a Jamison sequence is a Hilbertian Jamison sequence. Our first goal in this paper is to prove the somewhat surprising fact that the converse is true: Theorem 1.5. -- Let (nk)k≥0 be an increasing sequence of integers. Then (nk)k≥0 is a Hilbertian Jamison sequence if and only if it is a Jamison sequence. Contrary to the proofs of [1] and [2], the proof of Theorem 1.5 is completely explicit: the operators with supk≥0 T nk < +∞ and σp(T ) ∩ T uncountable which we construct are perturbations by a weighted backward shift on ℓ2 of a diagonal operator with unimodular diagonal coefficients. The construction here bears some similarities with a construction carried out in a different context in [10] in order to obtain frequently hypercyclic operators on certain Banach spaces. 1.2. Ergodic theory and rigidity sequences. -- Let (X,F, µ) be a finite measure space where µ is a positive regular finite Borel measure, and let ϕ be a measurable transformation of (X,F, µ). We recall here that ϕ is said to preserve the measure µ if µ(ϕ−1(A)) = µ(A) for any A ∈ F, and that ϕ is said to be ergodic with respect to µ if for any A, B ∈ F with µ(A) > 0 and µ(B) > 0, there exists an n ≥ 0 such that µ(ϕ−n(A)∩ B) > 0, where ϕn denotes the nth iterate of ϕ. Equivalently, ϕ is ergodic with respect to µ if and only if µ(ϕ−n(A) ∩ B) → µ(A)µ(B) as N → +∞ for every A, B ∈ F. This leads to the notion of weakly mixing measure-preserving transformation of (X,F, µ): ϕ is weakly mixing if 1 N N Xn=1 1 N N Xn=1 µ(ϕ−n(A) ∩ B) − µ(A)µ(B) → 0 as N → +∞ for every A, B ∈ F. It is well-know that ϕ is weakly mixing if and only if ϕ× ϕ is an ergodic transformation of X × X endowed with the product measure µ × µ. We refer the reader to [9], [30] or [38] for instance for more about ergodic theory of dynamical systems and various examples. Our interest in this paper lies in weakly mixing rigid dynamical systems: 4 TANJA EISNER & SOPHIE GRIVAUX Definition 1.6. -- A measure-preserving transformation of (X,F, µ) is said to be rigid if there exists a sequence (nk)k≥0 of integers such that for any A ∈ F, µ(ϕ−nk (A)△A) → 0 as k → +∞. If Uϕ denotes the isometry on L2(X,F, µ) defined by Uϕf := f ◦ϕ for any f ∈ L2(X,F, µ), it is not difficult to see that ϕ is rigid with respect to the sequence (nk)k≥0 if and only ϕ f − f → 0 as k → +∞ for any f ∈ L2(X,F, µ). The function f itself is said to if U nk be rigid with respect to (nk)k≥0 if U nk ϕ f − f → 0. Rigid functions play a major role in the study of mildly mixing dynamical systems, as introduced by Furstenberg and Weiss in [12], and rigid weakly mixing systems are intensively studied, see for instance the works [15], [26], [16] or [36] as well as the references therein for some examples of results and methods. Let us just mention here the fact that weakly mixing rigid transformations of (X,F, µ) form a residual subset of the set of all measure-preserving transformations of (X,F, µ) for the weak topology [23]. A rigidity sequence is defined as follows: Definition 1.7. -- Let (nk)k≥0 be a strictly increasing sequence of positive integers. We say that (nk)k≥0 is a rigidity sequence if there exists a measure space (X,F, µ) and a measure-preserving transformation ϕ of (X,F, µ) which is weakly mixing and rigid with respect to (nk)k≥0. Remark 1.8. -- In the literature one often defines rigidity sequences as sequences for which there exists an invertible measure-preserving transformation which is weakly mixing and rigid with respect to (nk)k≥0. In fact, these two definitions are equivalent since every rigid measure-preserving transformation ϕ is invertible (in the measure-theoretic sense). An easy way to see it is to consider the induced isometry Uϕ defined above. Since ϕ is invertible if and only if Uϕ is so, it suffices to show that Uϕ is invertible. By the decomposition theorem for contractions due to Sz.-Nagy, Foia¸s [37], Uϕ can be decomposed into a direct sum of a unitary operator and a weakly stable operator. Since limk→∞ U nk ϕ = I in the weak operator topology (see Fact 3.2 below), the weakly stable part cannot be present and thus Uϕ is a unitary operator and ϕ is invertible. Rigidity sequences are in a sense already characterized: (nk)k≥0 is a rigidity sequence if and only if there exists a continuous probability measure σ on the unit circle T such that ZT λnk − 1dσ(λ) → 0 as k → +∞ (see Section 3.1 for more details). Still, there is a lack of practical conditions which would enable us to check easily whether a given sequence (nk)k≥0 is a rigidity sequence. It is the second aim of this paper to provide such conditions. Some of them can be initially found in the papers [1] and [2] which study Jamison sequences in the Banach space setting, and they turn out to be relevant for the study of rigidity. We show for instance that if nk+1/nk tends to infinity as k tends to infinity, (nk)k≥0 is a rigidity sequence (see Example 3.4 and Proposition 3.5). If (nk)k≥0 is any sequence such that nk divides nk+1 for any k ≥ 0, (nk)k≥0 is a rigidity sequence (Propositions 3.8 and 3.9). We also give some examples involving the denominators of the partial quotients in the continuous fraction expansion of some irrational numbers (Examples 3.15 and 3.16), as well as an example of HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 5 a rigidity sequence such that nk+1/nk → 1 (Example 3.17). In the other direction, it is not difficult to show that if nk = p(k) for some polynomial p ∈ Z[X] with p(k) ≥ 0 for any k, (nk)k≥0 cannot be a rigidity sequence (Example 3.12), or that the sequence of prime numbers cannot be a rigidity sequence (Example 3.14). Other examples of non-rigidity sequences can be given (Example 3.13) when the sequences (nkx)k≥0, x ∈ [0, 1], have suitable equirepartition properties. 1.3. Ergodic theory and rigidity for linear dynamical systems. -- If T is a bounded operator on a separable Banach space X, it is sometimes possible to endow the space X with a suitable probability measure m, and to consider (X,B, m, T ) as a measurable dynamical system. This was first done in the seminal work [11] of Flytzanis, and the study was continued in the papers [4] and [5]. If X is a separable complex Hilbert space which we denote by H, T ∈ B(H) admits a non-degenerate invariant Gaussian measure if and only if its eigenvectors associated to eigenvalues of modulus 1 span a dense subspace of H, and it is ergodic (or here, equivalently, weakly mixing) with respect to such a measure if and only if it has a perfectly spanning set of eigenvectors associated to unimodular eigenvalues (see Section 2.1 for the definitions) -- this condition very roughly means that T has "plenty" of such eigenvectors, "plenty" being quantified by a continuous probability measure on the unit circle. It comes as a natural question to describe rigidity sequences in the framework of linear dynamics, and it is not difficult to show that if (nk)k≥0 is a rigidity sequence, there exists a bounded operator on H which is weakly mixing and rigid with respect to (nk)k≥0 (see Section 4.1). Thus, every rigidity sequence can be realized in a linear Hilbertian measure- preserving dynamical system. However, the answer is not so simple when one considers topological and uniform rigidity, which are topological analogues of the (measurable) no- tion of rigidity. These notions were introduced by Glasner and Maon in the paper [14] for continuous dynamical systems on compact spaces: Definition 1.9. -- Let (X, d) be a compact metric space, and let ϕ be a continuous self-map of X. We say that ϕ is topologically rigid with respect to the sequence (nk)k≥0 if ϕnk (x) → x as k → +∞ for any x ∈ X, and that ϕ is uniformly rigid with respect to (nk)k≥0 if sup x∈X d(ϕnk (x), x) → 0 as k → +∞. It is easy to check, using the Lebesgue dominated convergence theorem, that a topolog- ically or uniformly rigid dynamical system is rigid. Uniform rigidity is studied in [14], where in particular uniformly rigid and topologically weakly mixing transformations are constructed, see also [8], [24] and [20] for instance. Recall that ϕ is said to be topologically weakly mixing if for any non-empty open subsets U1, U2, V1, V2 of X, there exists an integer n such that ϕ−n(U1)∩ V1 and ϕ−n(U2)∩ V2 are both non-empty (topological weak mixing is the topological analogue of the notion of measurable weak mixing). Uniform rigidity sequences are defined in [20]: Definition 1.10. -- Let (nk)k≥0 be a strictly increasing sequence of integers. We say that (nk)k≥0 is a uniform rigidity sequence if there exists a compact dynamical system 6 TANJA EISNER & SOPHIE GRIVAUX (X, d, ϕ) with ϕ a continuous self-map of X, which is topologically weakly mixing and uniformly rigid with respect to (nk)k≥0. The question of characterizing uniform rigidity sequences is still open, as well as the ques- tion [20] whether there exists a compact dynamical system (X, d, ϕ) with ϕ continuous, which would be both weakly mixing with respect to a certain ϕ-invariant measure µ on X and uniformly rigid. We investigate these two questions in the framework of linear dynamics. Of course we have to adapt the definition of uniform rigidity to this setting, as a Banach space is never compact. Definition 1.11. -- Let X be complex separable Banach space, and let ϕ be a continuous transformation of X. We say that ϕ is uniformly rigid with respect to (nk)k≥0 if for any bounded subset A of X, x∈Aϕnk (x) − x → 0 sup as k → +∞. When T is a bounded linear operator on X, T is uniformly rigid with respect to (nk)k≥0 if and only if T nk − I → 0 as k → +∞. We prove the following theorems: Theorem 1.12. -- Let (nk)k≥0 be an increasing sequence of integers with n0 = 1. The following assertions are equivalent: (1) for any ε > 0 there exists a λ ∈ T \ {1} such that λnk − 1 → 0 k≥0 λnk − 1 ≤ ε sup and as k → +∞; (2) there exists a bounded linear operator T on a separable Banach space X such that σp(T ) ∩ T is uncountable and T nk x → x as k → ∞ for every x ∈ X; (3) there exists a bounded linear operator T on a separable Hilbert space H such that T admits a non-degenerate invariant Gaussian measure with respect to which T is weakly mixing and T nkx → x as k → +∞ for every x ∈ H, i.e. T is topologically rigid with respect to (nk)k≥0. We also have a characterization for uniform rigidity in the linear setting: Theorem 1.13. -- Let (nk)k≥0 be an increasing sequence of integers. The following as- sertions are equivalent: (1) there exists an uncountable subset K of T such that λnk tends to 1 uniformly on K; (2) there exists a bounded linear operator T on a separable Banach space X such that σp(T ) ∩ T is uncountable and T nk − I → 0 as k → ∞; (3) there exists a bounded linear operator T on a separable Hilbert space H such that T admits a non-degenerate invariant Gaussian measure with respect to which T is weakly mixing and T nk − I → 0 as k → ∞, i.e. T is uniformly rigid with respect to (nk)k≥0. In particular we get a positive answer to a question of [20] in the context of linear dynamics: HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 7 Corollary 1.14. -- Any sequence (nk)k≥0 such that nk+1/nk tends to infinity, or such that nk divides nk+1 for each k and lim sup nk+1/nk = +∞ is a uniform rigidity sequence for linear dynamical systems, and measure-theoretically weakly mixing uniformly rigid sys- tems do exist in this setting. After this paper was submitted for publication, V. Bergelson, A. Del Junco, M. Lema´nczyk and J. Rosenblatt sent us a preprint "Rigidity and non-recurrence along sequences" [7], in which they independently investigated for which sequences there exists a weakly mixing transformation which is rigid with respect to this sequence. A substantial part of the results of Section 3 of the present paper is also obtained in [7], often with different methods. We are very grateful to V. Bergelson, A. Del Junco, M. Lema´nczyk and J. Rosenblatt for making their preprint available to us. 2. Hilbertian Jamison sequences Our aim in this section is to prove Theorem 1.5. Clearly, if (nk)k≥0 is a Jamison sequence, it is automatically a Hilbertian Jamison sequence, and the difficulty lies in the converse direction: using Theorem 1.3, we start from a sequence (nk)k≥0 such that for any ε > 0 there is a λ ∈ T \ {1} such that supk≥0 λnk − 1 ≤ ε, and we have to construct out of this a bounded operator on a Hilbert space which is partially power-bounded with respect to (nk)k≥0 and which has uncountably many eigenvalues on the unit circle. We are going to prove a stronger theorem, giving a more precise description of the eigenvectors of the operator: Theorem 2.1. -- Let (nk)k≥0 be an increasing sequence of integers with n0 = 1 such that for any ε > 0 there exists a λ ∈ T \ {1} such that k≥0 λnk − 1 ≤ ε. sup Let δ > 0 be any positive number. There exists a bounded linear operator T on the Hilbert space ℓ2(N) such that T has perfectly spanning unimodular eigenvectors and k≥0 T nk ≤ 1 + δ. sup In particular the unimodular point spectrum of T is uncountable. Before embarking on the proof, we need to define precisely the notion of perfectly spanning unimodular eigenvectors and explain its relevance here. 2.1. A criterion for ergodicity of linear dynamical systems. -- Let H be a com- plex separable infinite-dimensional Hilbert space. Definition 2.2. -- We say that a bounded linear operator T on H has a perfectly span- ning set of eigenvectors associated to unimodular eigenvalues if there exists a continuous probability measure σ on the unit circle T such that for any Borel subset B of T with σ(B) = 1, we have sp[ker(T − λI) ; λ ∈ B] = H. 8 TANJA EISNER & SOPHIE GRIVAUX When T ∈ B(H) has a perfectly spanning set of eigenvectors associated to unimodular eigenvalues, there exists a Gaussian probability measure m on H such that: -- m is T -invariant; -- m is non-degenerate, i.e. m(U ) > 0 for any non-empty open subset U of H; -- T is weakly mixing with respect to m. See [5] for extensions to the Banach space setting, and the book [6, Ch. 5]. In the Hilbert space case, the converse of the assertion above is also true: if T ∈ B(H) defines a weakly mixing measure-preserving transformation with respect to a non-degenerate Gaus- sian measure, T has perfectly spanning unimodular eigenvectors. A way to check this spanning property of the eigenvectors is to use the following criterion, which was proved in [17, Th. 4.2]: Theorem 2.3. -- Let X be a complex separable infinite-dimensional Banach space, and let T be a bounded operator on X. Suppose that there exists a sequence (ui)i≥1 of vectors of X having the following properties: (i) for each i ≥ 1, ui is an eigenvector of T associated to an eigenvalue µi of T where µi = 1 and the µi's are all distinct; (ii) sp[ui ; i ≥ 1] is dense in X; (iii) for any i ≥ 1 and any ε > 0, there exists an n 6= i such that un − ui < ε. Then T has a perfectly spanning set of eigenvectors associated to unimodular eigenvalues. 2.2. Proof of Theorem 2.1: the easy part. -- We are first going to define the oper- ator T , and show that it is bounded. We will then describe the unimodular eigenvectors of T , and show that T satisfies the assumptions of Theorem 2.3. ◮ Construction of the operator T . Let (en)n≥1 denote the canonical basis of the space ℓ2(N) of complex square summable sequences. We denote by H the space ℓ2(N). The construction depends on two sequences (λn)n≥1 and (ωn)n≥1 which will be suitably chosen further on in the proof: (λn)n≥1 is a sequence of unimodular complex numbers which are all distinct, and (ωn)n≥1 is a sequence of positive weights. Let j : {2, +∞} → {1, +∞} be a function having the following two properties: • for any n ≥ 2, j(n) < n; • for any k ≥ 1, the set {n ≥ 2 ; j(n) = k} is infinite (i.e. j takes every value k infinitely often). Let D be the diagonal operator on H defined by Den = λnen for n ≥ 1, and let B be the weighted backward shift defined by Be1 = 0 and Ben = αn−1en−1 for n ≥ 2, where the weights αn, n ≥ 1, are defined by and α1 = ω1 λ2 − λj(2) λn+1 − λj(n+1) λn − λj(n) (cid:12)(cid:12)(cid:12)(cid:12) This definition of αn makes sense because j(n) < n, so that λn 6= λj(n). The operators D and B being thus defined, we set T = D + B. αn = ωn (cid:12)(cid:12)(cid:12)(cid:12) for any n ≥ 2. HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 9 ◮ Boundedness of the operator T . The first thing to do is to prove that T is indeed a bounded operator on H, provided some conditions on the λn's and ωn's are imposed. The diagonal operator D being obviously bounded, we have to figure out conditions for B to be bounded. If γ > 0 is fixed, we have B ≤ γ provided ω1 λ2 − λj(2) ≤ γ ≤ γ for any n ≥ 3. and ωn−1 (cid:12)(cid:12)(cid:12)(cid:12) λn − λj(n) λn−1 − λj(n−1)(cid:12)(cid:12)(cid:12)(cid:12) If the weights ωn > 0 are arbitrary, the λn's can be chosen in such a way that these conditions are satisfied: • ω1 > 0 is arbitrary, we take λ1 = 1 for instance (we could start here from any λ1 ∈ T); • we have j(2) = 1: take λ2 such that λ2 − λ1 ≤ γ/ω1 with λ2 6= λ1; • take ω2 > 0 arbitrary; • j(3) ∈ {1, 2}: take λ3 so close to λj(3), λ3 6∈ {λ1, λ2}, that γ ω2 λ2 − λj(2) λ3 − λj(3) ≤ • take ω3 > 0 arbitrary, etc. Thus B ≤ γ provided λn is so close to λj(n) for every n ≥ 2 that λn − λj(n) ≤ γ ωn−1 λn−1 − λj(n−1). No condition on the ωn's needs to be imposed there. ◮ Unimodular eigenvectors of the operator T . The algebraic equation T x = λx with xk x = Pk≥1 xkek is equivalent to the equations λkxk + αkxk+1 = λxk, i.e. xk+1 = λ−λk for any k ≥ 1, i.e. αk xk = (λ − λk−1) . . . (λ − λ1) x1. αk−1 . . . α1 Hence for any n ≥ 1, the eigenspace ker(T−λn) is 1-dimensional and ker(T−λn) = sp[u(n)], where u(n) = e1 + (λn − λk−1) . . . (λn − λ1) αk−1 . . . α1 ek. n Xk=2 Our aim is now to show the following lemma: Lemma 2.4. -- By choosing in a suitable way the coefficients ωn and λn, it is possible to ensure that for any n ≥ 2, u(n) − u(j(n)) ≤ 2−n (the sequence (2−n)n≥2 could be replaced by any sequence (γn)n≥2 going to zero with n). Proof of Lemma 2.4. -- We have: u(n) − u(j(n)) = + j(n) αk−1 . . . α1 Xk=2(cid:18) (λn − λk−1) . . . (λn − λ1) Xk=j(n)+1 (λn − λk−1) . . . (λn − λ1) αk−1 . . . α1 n (λj(n) − λk−1) . . . (λj(n) − λ1) αk−1 . . . α1 − (cid:19) ek ek := v(n) + w(n). 10 TANJA EISNER & SOPHIE GRIVAUX We denote the first sum by v(n) and the second one by w(n). If εn > 0 is any positive number, we can ensure that v(n) < εn by choosing λn such that λn−λj(n) is sufficiently small, because the quantities αk−1 . . . α1 for k ≤ j(n) do not depend on λn. Let us now estimate w(n)2 = = since 2 n n (cid:12)(cid:12)(cid:12)(cid:12) αk−1 . . . α1 (λn − λk−1) . . . (λn − λ1) Xk=j(n)+1(cid:12)(cid:12)(cid:12)(cid:12) Xk=j(n)+1 αk−1 . . . α1 = ωk−1 . . . ω1 λk − λj(k). 1 · (cid:12)(cid:12)(cid:12)(cid:12) k−1 . . . ω2 ω2 1 (λn − λk−1) . . . (λn − λ1) λk − λj(k) 2 (cid:12)(cid:12)(cid:12)(cid:12) We estimate now each term in this sum. We can suppose that λp− λq ≤ 1 for any p and q (this is no restriction), so (λn−λk−1) . . . (λn−λ1) ≤ λn−λj(n) since j(n) ∈ {1, . . . , k−1}. Thus for k = j(n) + 1, . . . , n, 1 ω2 k−1 . . . ω2 (λn − λk−1) . . . (λn − λ1) λk − λj(k) 1 ≤ ω2 k−1 . . . ω2 2 (cid:12)(cid:12)(cid:12)(cid:12) 2 · 1 · (cid:12)(cid:12)(cid:12)(cid:12) λn − λj(n) λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) 1 · (cid:12)(cid:12)(cid:12)(cid:12) If k ∈ {j(n) + 1, . . . , n− 1}, the term on the right-hand side can be made arbitrarily small provided that we choose λn in such a way that λn − λj(n) is very small with respect to the quantities λk − λj(k) . ωk−1 . . . ω1, k < n. However for k = n, we only get the bound ω−2 n−1 . . . ω−2 1 , which has to be made small if we want w(n) to be small. So we have to impose a condition on the weights ωn: we take ωn−1 so large with respect to ω1, . . . , ωn−2 that ω−2 All the conditions on the λn's and the ωn's needed until now can indeed be satisfied simultaneously: at stage n of the construction, we take ωn−1 very large. After this we take λn extremely close to λj(n). Thus we can ensure that w(n) < εn, hence that u(n) − u(j(n)) < 2εn. Taking εn = 2−(n+1) gives our statement. Thanks to Lemma 2.4, it is easy to see that T satisfies the assumptions of Theorem 2.3: is extremely small. n−1 . . . ω−2 1 Proposition 2.5. -- The operator T satisfies the assumptions of Theorem 2.3. Hence it has a perfectly spanning set of eigenvectors associated to unimodular eigenvalues, and in particular its unimodular point spectrum is uncountable. Proof of Proposition 2.5. -- It suffices to show that the sequence (u(n))n≥1 satisfies prop- erties (i), (ii) and (iii). That (i) is satisfied is clear, since the vectors u(n) are eigenvectors of T associated to the eigenvalues λn which are all distinct. Since for each n ≥ 1 the vector u(n) belongs to the span of the first n basis vectors e1, . . . , en and hu(n), eni 6= 0, the linear span of the vectors u(n), n ≥ 1, contains all finitely supported vectors of ℓ2(N), and thus (ii) holds true. It remains to prove (iii). As the function j takes every value in N infinitely often, it follows from Lemma 2.4 that for any n ≥ 1 there exists a strictly increasing sequence (p(n) s )s≥1 of integers such that u(p ) − u(n) tends to 0 as s tends to + ∞. (n) s HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 11 Hence (iii) is true. In order to conclude the proof of Theorem 1.5, it remains to show that T is partially power- bounded with respect to (nk)k≥0, with supk≥0 T nk ≤ 1 + δ. This is the most difficult part of the proof, which uses the assumption that (nk)k≥0 is not a Jamison sequence, and it is the object of the next section. 2.3. Proof of Theorem 2.1: the hard part. -- In order to estimate the norms T nk, we will show that provided the ωn's and λn's are suitably chosen, T nk − Dnk ≤ δ for every k ≥ 1. Since Dnk = 1, this will prove that T nk ≤ 1 + δ for every k ≥ 1. ◮ An expression of (T n − Dn). We first have to compute (T n − Dn)x for n ≥ 1 and x ∈ H. For k, l ≥ 1, let t(n) k,l = hT nel, eki be the coefficient in row k and column l of the matrix representation of T n. If k > l, t(n) k,l = 0 (all coefficients below the diagonal are zero), and if l − k > n, t(n) k,l = 0 (all coefficients which are not in one of the first n upper diagonals of the matrix are zero). We have t(n) k,k = λn k for any k ≥ 1. Lemma 2.6. -- For any k, l ≥ 1 such that 1 ≤ l − k ≤ n, k,l = αl−1αl−2 . . . αk Xjk+...+jl=n−(l−k) t(n) k . . . λjl λjk l . Proof. -- The proof is done by induction on n ≥ 1. • n = 1: in this case l = k + 1, and the formula above gives t(1) • Suppose that the formulas above are true for any m ≤ n. Let k and l be such that 1 ≤ l − k ≤ n + 1 (in particular l ≥ 2). We have k,l t(1) k,k+1 = αk, which is true. l,l = αl−1t(n) k,l−1 + λlt(n) k,l . k,l = t(n) t(n+1) l−1,l + t(n) k,l−1t(1) If 2 ≤ l − k ≤ n, we can apply the induction assumption to the two quantities t(n) t(n) k,l , and we get k,l−1 and t(n+1) k,l = αl−1αl−2 . . . αk k . . . λjl−1 λjk l−1 X jk+...+jl−1=n−(l−1−k) k . . . λjl−1 λjk l−1 λjl+1 l + αl−1αl−2 . . . αk Xjk+...+jl=n−(l−k) k . . . λjl−1 λjk l−1 λjl l k . . . λjl−1 λjk l−1 λjl l = + = jk+...+jl−1+jl=n+1−(l−k), jl=0 jk+...+jl−1+jl=n+1−(l−k), jl≥1 k . . . λjl−1 λjk l−1 λjl l jk+...+jl−1+jl=n+1−(l−k) X X X and the formula is proved for 1 ≤ l − k ≤ n. It remains to treat the cases where l − k = 1 and where l− k = n + 1. If l− k = 1, we have t(n+1) k,k+1. By the induction k,k+1 = αkλn k + λk+1t(n) 12 assumption so that TANJA EISNER & SOPHIE GRIVAUX k λjk+1 λjk k+1 = αk k k+1 − λn λn λk+1 − λk t(n) k,k+1 = αk Xjk+jk+1=n−1 λn k+1 − λn λk+1 − λk(cid:19) = αk k t(n+1) k + λk+1 k,k+1 = αk(cid:18)λn = αk Xjk+jk+1=n which is the formula we were looking for. Lastly, when l−k = n+1, t(n+1) By the induction assumption t(n) formula is proved in this case too. k,n+k = αn+k−1 . . . αk, thus t(n+1) k+1 − λn+1 λn+1 λk+1 − λk k k,n+1+k = αn+kt(n) k,n+k. k,n+1+k = αn+k . . . αk and the k λjk+1 λjk k+1 ◮ A first estimate on the norms (T n − Dn). For x = Pl≥1 xlel, let us estimate (T n − Dn)x2: we have (T n − Dn)x =Xl≥1 k,l ek xl Xk=max(1,l−n)  k,l ek  so that by the Cauchy-Schwarz inequality t(n) t(n) l−1 (T n − Dn)x2 ≤ x2Xl≥2 ≤ x2Xl≥2 We thus have to estimate for each l ≥ 2 and p ≥ 1 the quantities Xk=max(1,l−n) l−1 Xk=max(1,l−n) t(n) k,l 2. xl Xk≥1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 l−1 t(n) xlt(n)  −Xl≥1 l,l el =Xl≥2 k,l ek(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=max(1,l−np) t(np) 2. l−1 k,l For k, l ≥ 1, 1 ≤ l − k ≤ n, let s(n) k,l = We have t(n) k,l = αl−1 . . . αk s(n) so that we have to estimate k . . . λjl λjk l . Xjk+...+jl=n−(l−k) k,l = ωl−1 . . . ωk λl − λj(l) λk − λj(k) s(n) k,l l−1 Xk=max(1,l−n) l−1 . . . ω2 ω2 k λl − λj(l)2 λk − λj(k)2 s(n) k,l 2. We are going to show that the following property holds true: j Lemma 2.7. -- For any 1 ≤ k ≤ l − 1, there exists for each k ≤ j ≤ l − 1 a complex number c(k,l) depending only on λ1, . . . , λl−1 (and k and l of course), but neither on λl nor on n, such that for any n ≥ l − k, Xj=k − λn+1−(l−k) λn+1−(l−k) l λl − λj s(n) k,l = c(k,l) j l−1 · j HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 13 Proof. -- The proof is again done by induction on l ≥ 2. • Let us first treat the case l = 2: we have to show that there exists c(1,2) any n ≥ 2, 1 such that for But s(n) 1,2 = c(1,2) 1 1 2 − λn λn λ2 − λ1 · 1 λn−1−j1 λj1 Xj1=0 n−1 2 = 1 2 − λn λn λ2 − λ1 s(n) 1,2 = Xj1+j2=n−1 λj1 1 λj2 2 = so this holds true with c(1,2) • Suppose that the property is true for some l ≥ 2, and consider for 1 ≤ k ≤ l and n ≥ l + 1 − k the quantities 1 = 1. s(n) k,l+1 = k . . . λjl+1 λjk l+1 jk+...+jl+1=n−(l+1−k) n−(l+1−k) l   λjl+1 If 1 ≤ k ≤ l − 1, we can apply the induction assumption and we get that jk+...+jl=n−(l+1+jl+1−k) Xjl+1=0 k . . . λjl λjk X = l+1 . X   n−(l+1−k) s(n) k,l+1 = = n−(l+1−k) Xjl+1=0 Xjl+1=0 k,l l+1 s(n−1−jl+1) λjl+1 j   Xj=k λjl+1 l+1 c(k,l) l−1 λn−jl+1−(l−k) l − λn−jl+1−(l−k) j λl − λj   depends only on λ1, . . . , λl−1 (the first equality in the display above comes j where c(k,l) from the fact that jl+1 ≤ n − 1 − l + k, i.e. l − k ≤ n − 1 − jl+1). Thus s(n) k,l+1 = l+1 λn−jl+1−(l−k) λjl+1 l+1 λn−jl+1−(l−k) c(k,l) j n−(l+1−k) l−1 j l − λjl+1 1 − (λl+1λl)n−(l−k) 1 − (λl+1λl) − λn−(l−k) l j − λn−(l−k) λn−(l−k) l+1 − λn−(l−k) j λl+1 − λj 1 − (λl+1λj)n−(l−k) 1 − (λl+1λj) !   ! l = = l−1 l−1 c(k,l) j c(k,l) j λl − λj  Xj=k Xjl+1=0  λl − λj λn−(l−k) Xj=k λl − λj λl Xj=k λl − λj! λn−(l−k) Xj=k − +  λn−(l−k) Xj=k  c(k,l) j λl − λj λl  λn−(l−k) l+1 λjc(k,l) l+1 l+1 l−1 l−1 = j λl+1 − λl j − λj − λn−(l−k) ! − λn−(l−k) ! l λl+1 − λl λl+1 − λj 14 i.e where TANJA EISNER & SOPHIE GRIVAUX s(n) k,l+1 = c(k,l+1) j l Xj=k λn+1−(l+1−k) l+1 − λn+1−(l+1−k) j λl+1 − λj c(k,l+1) j = − j λjc(k,l) λl − λj for k ≤ j ≤ l − 1 and c(k,l+1) l = l−1 Xj=k c(k,l) j λl − λj λl depend only on λ1, . . . , λl. This settles the case where 1 ≤ k ≤ l − 1. When k = l, we get s(n) l,l+1 = Xjl+jl+1=n−1 l λjl+1 λjl l+1 = l l+1 − λn λn λl+1 − λl , and the statement is true with c(l,l+1) Let us now go back to the estimate on (T np − Dnp)x2, p ≥ 0: we want to show that if the coefficients λl are suitably chosen, the following holds true: for any p ≥ 0 we have • for any 2 ≤ l ≤ np, = 1. l l−1 • for any l ≥ np + 1, Xk=1 l−1 t(np) k,l 2 ≤ δ2 2−l, k,l t(np) 2 ≤ δ2 2−l. Xk=l−np Let us first consider the case 2 ≤ l ≤ np. ◮ The "easy" estimate on (T np − Dnp). Let us write Xk=1 λl − λj(l) l−1 . . . ω2 ω2 2 = t(np) l−1 l−1 k,l 2 Xk=1 Xk=1 l−1 k(cid:12)(cid:12)(cid:12)(cid:12) λl − λj(l) s(n) k,l 2 2  λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) k(cid:12)(cid:12)(cid:12)(cid:12) k(cid:12)(cid:12)(cid:12)(cid:12) λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) 2 λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) c(k,l) Xj=k,j6=j(l)  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  − λnp+1−(l−k)  λl − λj(l) 2 λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) Xj=k,j6=j(l)  Xj=k (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) λl − λj(l) k(cid:12)(cid:12)(cid:12)(cid:12) l−1 l−1 j(l) 2 j λnp+1−(l−k) l l−1 . . . ω2 ω2 j(l) Xk=1 + c(k,l) j(l) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=j(l)+1 l−1 + l−1 . . . ω2 ω2 ≤ l−1 . . . ω2 ω2   In the sum indexed by j, we have two different cases to consider: either j = j(l) or j 6= j(l). The case j = j(l) can happen only when j(l) ≥ k. Thus the sum can be decomposed as − λnp+1−(l−k) λnp+1−(l−k) l λl − λj(l) λl − λj c(k,l) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) l−1 . j j 2 λnp+1−(l−k) l − λnp+1−(l−k) j λl − λj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) λnp+1−(l−k) l − λnp+1−(l−k) j λl − λj c(k,l) j (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 15 which is less than l−1 λnp+1−(l−k) l − λnp+1−(l−k) j λl − λj l−1 . . . ω2 ω2 k . k(cid:12)(cid:12)(cid:12)(cid:12) 2 +2 j(l) l−1 . . . ω2 ω2 Xk=1 Xk=1 λl − λj(l)2  8 j(l) +2 l−1 . . . ω2 ω2 k . j 1 l−1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) c(k,l) λl − λj(l) λnp+1−(l−k) l 2 λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) Xj=k,j6=j(l)  j(l) 2 . (cid:12)(cid:12)(cid:12) λk − λj(k)2 .c(k,l) λk − λj(k)2 .  Xj=k,j6=j(l)  j(l) 2 . (cid:12)(cid:12)(cid:12) λk − λj(k)2 .c(k,l) λnp+1−(l−k) l l−1 . . . ω2 ω2 k . l−1 1 1 l−1 Xk=1 Xk=1 − λnp+1−(l−k) j(l) 2 1 (cid:12)(cid:12)(cid:12) λl − λj  (cid:12)(cid:12)(cid:12) 2 . c(k,l) j − λnp+1−(l−k) j(l) 2   (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2  and this in turn is less than (1) Suppose (as we may) that λl is so close to λj(l) that λl − λj(l) ≤ 1 2 min j≤l−1,j6=j(l)λj − λj(l). Then for any j ≤ l−1 with j 6= j(l) we have λl−λj ≥ λj−λj(l)−λl−λj(l) ≥ 1 Thus the first term in the expression (1) above is less than λj(l) − λj  λk − λj(k)2 .   32λl − λj(l)2 Xj=k,j6=j(l) l−1 . . . ω2 ω2 k . c(k,l) Xk=1 l−1 l−1 1 1 j 2 . Since the quantity between the brackets depends only on λ1, . . . , λl−1, ω1, . . . , ωl−1 but not on λl, the expression in the display above can be made arbitrarily small if λl − λj(l) is small enough. So we take, for any l ≥ 2, λl with λl − λj(l) so small that 2λj−λj(l). 32λl − λj(l)2 l−1 Xk=1 l−1 . . . ω2 ω2 k . 1 λk − λj(k)2 .   l−1 Xj=k,j6=j(l) c(k,l) j 2 1 λj(l) − λj  Observe that the estimate we get here does not depend on np (it is valid for any n in fact). This takes care of the first term in the sum (1). ◮ The "hard" estimate on T np − Dnp. We have now to estimate the term ≤ δ22−(l+1). (2) 2 We have j(l) l−1 . . . ω2 ω2 k . Xk=1 λnp+1−(l−k) l 1 j(l) 2 . (cid:12)(cid:12)(cid:12) λk − λj(k)2 .c(k,l) λnp+1−(l−k) l − λnp+1−(l−k) j(l) . 2 (cid:12)(cid:12)(cid:12) − λnp+1−(l−k) j(l) ≤ λnp ≤ λnp ≤ λnp l − λnp l − λnp l − λnp − λl−k−1 j(l) + λl−k−1 j(l) + (l − k − 1)λl − λj(l) j(l) + (l − 2)λl − λj(l) j(l) l 16 TANJA EISNER & SOPHIE GRIVAUX so that the quantity in (2) is less than j(l) (3) 4λnp l − λnp j(l)2 c(k,l) j(l) 2 l−1 . . . ω2 ω2 k . Xk=1 +4(l − 2)2λl − λj(l)2 λk − λj(k)2 Xk=1 j(l) l−1 . . . ω2 ω2 k . c(k,l) j(l) 2 λk − λj(k)2· As previously the second term in this sum can be made arbitrarily small for any l ≥ 2 provided λl − λj(l) is sufficiently small, and we can ensure that it is less than δ22−(l+2). The difficult term to handle is the first one, and it is here that we use our assumption on the sequence (np)p≥0 (which was never used in the proof until this point). Our assumption is that for any ε > 0 there exists a λ ∈ T \ {1} such that supp≥0 λnp − 1 ≤ ε. This can be rewritten using the distance on T defined by d(np)(λ, µ) = sup p≥0 λnp − µnp as: for any ε > 0 there exists a λ ∈ T \ {1} such that d(np)(λ, 1) ≤ ε. This means (see [2]) that there exists an uncountable subset K of T such that (K, d(np)) is a separable metric space. Thus K contains a subset K ′ which is uncountable and perfect for the distance d(np). This means that for every ε > 0 and every λ ∈ K ′, there exists a λ′ ∈ K ′, λ′ 6= λ such that d(np)(λ, λ′) < ε. Observe that since n0 = 1, λ − λ′ ≤ d(np)(λ, λ′) < ε. In the construction of the λl's, l ≥ 1, we start by taking λ1 ∈ K ′. Then we take λ2 in K ′ such that d(np)(λ2, λj(2)) is extremely small, which is possible since λj(2) = λ1 is not isolated in K ′. In the same way we can take the λl's for l ≥ 2 to be elements of K ′ such that d(np)(λl, λ(l)) is arbitrarily small, λl 6= λj(l). Then λl − λj(l) is also arbitrarily small. With this suitable choice of the λl's, we can estimate the remaining term in (3): (4) 4λnp l − λnp j(l)2 j(l) l−1 . . . ω2 ω2 k . Xk=1 ≤ 4 d(np)(λl, λj(l))2 c(k,l) j(l) 2 j(l) λk − λj(k)2 Xk=1 l−1 . . . ω2 ω2 k . c(k,l) j(l) 2 λk − λj(k)2· Since the sum in k depends only on λ1, . . . , λl−1, ω1, . . . , ωl−1, but not on λl, by taking λl such that d(np)(λl, λj(l)) is extremely small, we ensure that the right-hand term in (4) is less than δ22−(l+2) for instance, for every l ≥ 2. Let us stress that the restrictions on the size of λl − λj(l) and d(np)(λl, λj(l)) are imposed for any l ≥ 2, and do not depend on a particular np. Thus we have proved that for any p ≥ 1 and any 2 ≤ l ≤ np, we have with these choices of λl l−1 Xk=1 t(np) k,l 2 ≤ δ22−l. HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 17 It remains to treat the case where l ≥ np + 1, where we have to estimate the quantity which is less than (5) l−1 Xk=l−np ω2 l−1 . . . ω2 k . (cid:12)(cid:12)(cid:12)(cid:12) λl − λj(l) λk − λj(k)(cid:12)(cid:12)(cid:12)(cid:12) l−1 Xk=l−np 2  Xj=k l−1 t(np) k,l 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) c(k,l) j λnp+1−(l−k) l − λnp+1−(l−k) j λl − λj 2 . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   We make the same decomposition as above in the sum in j, by separating the cases j = j(l) and j 6= j(l). The case j = j(l) can only happen when j(l) ≥ k, so when j(l) ≥ l − np, i.e. np ≥ l − j(l). The estimates on the term not involving the index j = j(l) are worked out exactly as previously, and this term can be made arbitrarily small provided λl − λj(l) is very small. The other term appears when np ≥ l − j(l), and is equal to c(k,l) j(l) 2 j(l) λnp+1−(l−k) l − λnp+1−(l−k) j(l) 2ω2 l−1 . . . ω2 k . 2 Xk=l−np (6) j(l) ≤ 4 d(np)(λl, λj(l))2 Xk=1 +4 (l − 2)2λl − λj(l)2 λk − λj(k)2 c(k,l) j(l) 2 l−1 . . . ω2 ω2 k . λk − λj(k)2 c(k,l) j(l) 2 λk − λj(k)2· l−1 . . . ω2 ω2 k . j(l) Xk=1 The reasoning is then exactly the same as previously, and if for each l ≥ 2 the quantity d(np)(λl, λj(l)) is sufficiently small we have for any p ≥ 1 and any l ≥ np + 1 that Hence T np − Dnp ≤ δ for any p ≥ 1. For p = 0, T − D = B < δ, so that l−1 Xk=l−np t(np) k,l 2 ≤ δ22−l. p≥0 T np − Dnp ≤ δ. sup This proves that T is partially power-bounded with respect to (np)p≥0, with the estimate supp≥0 T np ≤ 1 + δ, and this finishes the proof of Theorem 2.1. Remark 2.8. -- It is not difficult to see that the operators constructed in Theorem 2.1 are invertible: they are of the form T = D + B , where D is invertible with D = 1, and B can be made arbitrarily small in the construction. 3. Rigidity sequences 3.1. An abstract characterization of rigidity sequences. -- As mentioned already in the introduction, it is in a sense not difficult to characterize rigidity sequences via measures on the unit circle although such a characterization is rather abstract and not so easy to handle in concrete situations. This characterization is well-know, and hinted 18 TANJA EISNER & SOPHIE GRIVAUX at for instance in [12] or [36], but since we have been unable to locate it precisely in the literature, we give below a quick proof of it. A proof is also given in the preprint [7]. Here and later we denote by σ(n) the n-th Fourier coefficient of a measure σ. Proposition 3.1. -- Let (nk)k≥0 be a strictly increasing sequence of positive integers. The following assertions are equivalent: (1) there exists a dynamical system ϕ on a measure space (X,F, µ) which is weakly mixing and rigid with respect to (nk)k≥0; (2) there exists a continuous probability measure σ on T such that σ(nk) → 1 as nk → +∞. Recall that a probability measure σ on T is said to be symmetric if σ(A) = σ(A) for any Borel subset A of T (where A denotes the conjugate set of A). In order to prove Proposition 3.1, we are going to use the following well-known fact: Fact 3.2. -- Let ϕ be a measure-preserving transformation of the space (X,F, µ). The following assertions are equivalent: (a) ϕ is rigid with respect to (nk)k≥0; (b) U nk (c) U nk ϕ → I in the weak operator topology (WOT) of L2(X,F, µ); ϕ → I in the strong operator topology (SOT) of L2(X,F, µ). Proof of Fact 3.2. -- The implication (c) ⇒ (b) is obvious. To prove (b) ⇒ (a) it suffices to apply (b) to the characteristic functions χA of sets A ∈ F: χA(ϕnk (x))χA(x)dµ(x) → µ(A) as nk → +∞. ϕ χA, χAi =ZX hU nk Now χA△ϕ−nk (A) = χA + χϕ−nk (A) − 2χAχϕ−nk (A) so that µ(A △ ϕ−nk (A)) = 2µ(A) − 2ZX χA(ϕnk (x))χA(x)dµ(x) → 0 as nk → +∞. ϕ f − fL2 → 0 for any f ∈ H0, we get in particular that σf0(nk) → f02 = 1 Since U nk as nk → +∞, so (2) holds true. Hence ϕ is rigid with respect to (nk)k≥0. The proof of (a) ⇒ (c) is done using the same kind of argument: thanks to the expression for χA△ϕ−nk (A), we get that U nk ϕ χA − χAL2 → 0 ϕ f − fL2 → 0 for any f ∈ L2(X,F, µ), which is as nk → +∞ for any A ∈ F. Hence U nk assertion (c). Proof of Proposition 3.1. -- Suppose first that (1) holds true, and let σ0 be the reduced maximal spectral type of Uϕ, i.e. the maximal spectral type of the unitary operator U induced by Uϕ on the subspace H0 = {f ∈ L2(X,F, µ) ; RX f (x)dµ(x) = 0} (which is invariant by Uϕ). Note that Uϕ is unitary by Remark 1.8. For the definition and basic properties of the spectral type of a unitary operator see e.g.[27]. Let f0 ∈ H0 with f0 = 1 be such that the spectral measure σf0 of f0 is a representant of the class σ0. Since ϕ is weakly mixing, σf0 is continuous, and it is a probability measure since f0 = 1. For any n ∈ Z we have ϕ f0, f0i =ZT hU n λndσf0(λ) = σf0(n). HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 19 Conversely, let σ be a continuous probability measure on T such that σ(nk) → 1. Then RT λnk − 12dσ(λ) → 0 as nk → +∞, so that in particular we have RT λnk − 1dσ(λ) → 0. Indeed λnk − 12 = 2(1 − ℜe(λnk )), hence RT λnk − 12dσ(λ) = 2ℜe(1 − σ(nk)) → 0. Let now σ be the probability measure on T defined by σ(A) = σ(A) for any A ∈ F. Then σ is a continuous measure on T, and RT λnk − 1dσ(λ) = RT λnk − 1dσ(λ) so that in particular σ(nk) → 1 as nk → +∞. Setting ρ := (σ + σ)/2, we obtain a continuous symmetric probability measure on T such that ρ(nk) → 1. So we can assume without loss of generality that the measure σ given by (2) is symmetric, and we have to construct out of this a weakly mixing measure-preserving transformation of a probability space which is rigid with respect to (nk). The class of transformations which we use for this is the class of stationary Gaussian processes. We refer the reader to one of the references [9], [31] or [29, Ch. 8] for details about this, and in the forthcoming proof we use the notations of [29, Ch. 8]. Since σ is a symmetric probability measure on T, there exists a real-valued stationary Gaussian process (Xn)n∈Z whose spectral measure is σ. This Gaussian process lives on a probability space (Ω, Σ, P) which can be realized as a sequence space: Ω = RZ, Σ is the σ-algebra generated by the sets Θm,A = {(ωn)n∈Z ; (ω−m, . . . , ωm) ∈ A}, m ≥ 0, A is a Borel subset of R2m+1, and P is the probability that (X−m, . . . , Xm) belongs to A, where P is τ -invariant for the shift τ of the space RZ, and (Xn) satisfy Xn+1 = Xn ◦ τ for any n ∈ Z. The shift τ defines a weakly mixing transformation of (Ω, Σ, P) (see [31] for instance), and we have to see that it is rigid with respect to the sequence (nk). Using the same argument as in the proof of [29, Ch. 8, Th. 3.2], it suffices to show that for any functions f, g belonging to Gc, the complexification of the Gaussian subspace of L2(Ω, Σ, P) τ f − f, gi → 0 as nk → +∞. If Φ : Gc → L2(T, σ) spanned by Xn, n ∈ Z, we have hU nk denotes the map defined on the linear span of the Xn's by Φ(P cnXn) := P cnλn, then Φ extends to a surjective isometry of Gc onto L2(T, σ), and we have for any f ∈ Gc that Uτ f = (Φ−1 ◦ Mλ ◦ Φ)f , where Mλ denotes multiplication by the independent variable λ on L2(T, σ). Thus hU nk τ f − f, gi = hM nk λ Φf − Φf, Φgi =ZT (λnk − 1)(Φf )(λ)(Φg(λ))dσ(λ). Now if h is any function in L1(T, σ), we have thatRT λnk−1h(λ)dσ(λ) → 0 as nk → +∞ (it suffices to approximate h by functions h′ ∈ L∞(T, σ) in L1(T, σ)). Since (Φf )(Φg) belongs to L1(T, σ), we get that hU nk τ → I in the WOT of L2(Ω, Σ, P) and τ is rigid with respect to (nk) by Fact 3.2. τ f − f, gi → 0. It follows that U nk Remark 3.3. -- The Gaussian dynamical systems considered in the proof of Proposition 3.1 live on the space of sequences RZ, which is not compact. But by the Jewett-Krieger Theorem (see for instance [30]), such a system is metrically isomorphic to a homeomor- phism of the Cantor set. 3.2. Examples of rigidity and non-rigidity sequences. -- Our first example of rigidity sequences (obtained also in [7]) is the following: Example 3.4. -- Let (nk)k≥0 be a strictly increasing sequence of positive numbers such that nk+1/nk tends to infinity. Then (nk)k≥0 is a rigidity sequence. 20 TANJA EISNER & SOPHIE GRIVAUX This fact follows from the following stronger Proposition 3.5 below, which will allow us to show later on in the paper that any such sequence is a uniform rigidity sequence in the linear framework. The proof of Proposition 3.5 uses ideas from [2]. Proposition 3.5. -- Let (nk)k≥0 be a strictly increasing sequence of positive numbers such that nk+1/nk tends to infinity. There exists a compact perfect subset K of T having the following two properties: (i) for any ε > 0 there exists a compact perfect subset Kε of K such that for any λ ∈ Kε, k≥0 λnk − 1 ≤ ε; sup (ii) λnk tends to 1 uniformly on K. Note that the existence of a compact perfect subset K of T satisfying (ii) above implies that (nk)k≥0 is a rigidity sequence. Indeed, any continuous probability measure σ supported on K satifies assertion (2) in Proposition 3.1. Proof. -- For any k ≥ 1, let γk = 5π supj≥k(nj−1/nj): γk decreases to 0 as k tends to infinity, and let k0 be such that γk ≤ 1 2 [ be such that γk = sin θk for k ≥ k0. The sequence (θk) decreases to 0, and θk ∼ γk as k tends to infinity. Thus there exists a k1 ≥ k0 such that for any k ≥ k1, θk ≥ 4π supj≥k(nj−1/nj) ≥ 4π (nk/nk+1), so that (nk+1/nk) . θk ≥ 4π. Let 2 for any k ≥ k0. Let θk ∈]0, π K0 = {λ ∈ T ; ∀k ≥ k1 λnk − 1 ≤ 2γk}. nk j If we write λ ∈ T as λ = e2iθ, θ ∈ [0, π[, λ belongs to K0 if and only if sin(nkθ) ≤ γk for any k ≥ k1. Let Fk = {θ ∈ [0, π[ ; sin(nkθ) ≤ sin θk}: Fk consists of intervals of the form [− θk ], l ∈ Z. We will construct a Cantor subset K of K0 as K =Tk≥k1Sj∈Ik have the form + lπ nk , θk + lπ nk nk j where the arcs J (k) J (k) j =(eiθ ; θ ∈"− J (k) l(k) j π nk #) l(k) j π nk θk nk θk nk (7) + , + j ∈ Z. Observe that such arcs are disjoint as soon as 2θk for some l(k) , i.e. θk < π 2 , which is indeed the case, and that the arc corresponding to l(k) j = 0 contains the point 1 in its interior. There are 2nk such intervals. We are going to construct by induction on k a collection (J (k) has the form given in (7) and is contained in an arc of the collection (J (k−1) )j∈Ik−1 constructed at step k − 1, and the j collection (J (k) )j∈Ik contains the arc [− θk , θk ] corresponding to the case l = 0. We start nk for k = k1 with the collection of all the 2n1 arcs above. Suppose that the arcs at step k are constructed, and write one of them as )j∈Ik in such a way that each J (k) < π nk nk nk j j j J (k) j ="− θk nk l(k) j π nk , θk nk + + l(k) j π nk # . Let us look for arcs of the form θk+1 nk+1 (cid:20)− + rπ nk+1 , θk+1 nk+1 + rπ nk+1(cid:21) , r ∈ Z, HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 21 j nk , θk nk nk . There are ⌊ 1 π (4π − π θk − θk+1)⌋ = pk+1 such intervals contained in J (k) contained in J (k) π ( nk+1 2 ) − 1 ≥ 2. Remark that in the case where J (k) By construction pk+1 ≥ 1 is the arc ]} we have in the collection (J (k+1) ) the arc {eiθ ; θ ∈ [− θk+1 {eiθ ; θ ∈ [− θk ]} , θk (which is indeed contained in the arc {eiθ ; θ ∈ [− θk ]}). We obtain in this fashion a nk perfect Cantor set K, which contains the point 1 by construction, such that λnk tends to 1 uniformly on K (as λnk − 1 ≤ 2γk for any λ ∈ K and any k ≥ k1). Let ε > 0. There exists an integer κ such that for any k ≥ κ and any λ ∈ K, λnk − 1 ≤ ε. Since 1 belongs to K, the set Kε = {λ ∈ K ; λ − 1 ≤ ε/nκ−1} is a compact perfect subset of T, and for any λ ∈ Kε and any 0 ≤ k ≤ κ − 1, , θk+1 nk+1 . j nk+1 nk j j λnk − 1 ≤ nk ≤ ε. nκ−1 ε Hence supk≥0 λnk − 1 ≤ ε for any λ ∈ Kε, and Proposition 3.5 is proved. Remark 3.6. -- We have shown at the end of the proof of Proposition 3.5 that if K is a compact perfect subset of T such that λnk tends to 1 uniformly on K, and if K contains the point 1, then for any ε > 0 there exists a λ ∈ K \{1} such that supk≥0 λnk − 1 ≤ ε. If we do not suppose that K contains the point 1, the set K = {λµ ; λ, µ ∈ K} is compact, perfect, contains the point 1, and λnk still tends to 1 uniformly on K. We thus have the following fact, which we record here for further use: Fact 3.7. -- The following assertions are equivalent: (i) there exists a compact perfect subset K of T such that λnk tends to 1 uniformly on K; (ii) there exists a compact perfect subset K of T such that λnk tends to 1 uniformly on K, and for any ε > 0 there exists a λ ∈ K \ {1} such that supk≥0 λnk − 1 ≤ ε. Our next examples concern sequences (nk)k≥0 such that nk divides nk+1 for any k ≥ 0 (we write this as nknk+1). We begin with the case where lim supk→∞ nk+1/nk = +∞, since in this case we can derive a stronger conclusion. Recall [1] that such sequences are Jamison sequences. Proposition 3.8. -- Let (nk)k≥0 be a sequence such that nknk+1 for any k ≥ 0 and lim supk→∞ nk+1/nk = +∞. There exists a compact perfect subset K of T containing the point 1 such that λnk → 1 uniformly on K. Proof. -- Since nknk+1 for any k ≥ 0, we have nk+1 ≥ 2nk, so that εp nkp θε =Xp≥1 , and λε = e2iπθε ∈ T. 1 Xk≥1 nk ≤ 1 and Xj≥k+1 1 nj ≤ 2 nk+1 for any k ≥ 1. nkp nkp−1 → +∞ as kp → ∞. Let (kp)p≥1 be a strictly increasing sequence of integers such that For any sequence ε ∈ {0, 1}N of zeroes and ones, ε = (εp)p≥1, consider the real number of [0, 1] 22 TANJA EISNER & SOPHIE GRIVAUX The set K = {λε ; ε ∈ {0, 1}N} is compact, perfect, and contains the point 1. Let us now show that λnk tends to 1 uniformly with respect to ε ∈ {0, 1}N. Fix δ > 0, and let p0 ≥ 1 ε 4π . Let k ≥ kp0, and ε ∈ {0, 1}N. There exists a be such that for any p ≥ p0, p ≥ p0 such that nkp ≤ nk ≤ nkp+1−1. We have < δ nkp−1 nkp nkθε = nk Xj=1 Since nkjnk for any j = 1, . . . , p, nkPp 1 j=1 p εj nkj εj nkj + nk Xj≥p+1 εj nkj · belongs to Z. Hence 1 nj ≤ 4π nk nkp+1 ≤ 4π nkj ≤ 2π nk Xj≥kp+1 e2iπnkθε − 1 ≤ 2π nk Xj≥p+1 ε − 1 < δ for any k ≥ kp0 and ε ∈ {0, 1}N. This proves our statement. so λnk Let us now move over to the case where nknk+1 for every k ≥ 0, but where nk+1/nk is possibly bounded: for instance nk = 2k for any k ≥ 0. Is (nk)k≥0 a rigidity sequence? Somewhat surprisingly, the answer is yes. This was kindly shown to us by Jean-Pierre Kahane, who proved the following proposition: < δ, nkp+1−1 nkp+1 Proposition 3.9. -- Let (nk)k≥0 be a sequence such that nknk+1 for every k ≥ 0. Then (nk)k≥0 is a rigidity sequence. This proposition is also proved in the preprint [7]. Proof. -- Let (ak)k≥1 be a decreasing sequence of positive numbers going to 0 as k goes to infinity, with ak < 1 for every k ≥ 1, such that the series Pk≥1 ak is divergent. Consider the infinite convolution of Bernoulli measures defined on [0, 2π] by µ = ∗j≥1((1 − aj)δ0 + ajδ 1 nj ), where δa denotes the Dirac measure at the point a for any a ∈ [0, 2π]. Clearly µ is a probability measure on [0, 2π] which is continuous. Indeed µ is the distribution of the random variable εj nj , ξ =Xj≥1 where (εj)j≥1 is a sequence of independent Bernoulli random variables taking values 0 and 1 with probabilities p0j = 1 − aj and p1j = aj respectively. Since P aj = +∞, the measure µ is continuous by a result of L´evy (see [13] for a simple proof). It thus remains to prove that µ(nk) → 1 as nk → +∞. Since njnj+1 for each j ≥ 0, (1 − aj(1 − e (1 − aj + aje nk nj )). 2iπ 2iπ nk µ(nk) = Yj≥k+1 nj ) = Yj≥k+1 j=1 xj − 1 ≤ PN nk 2iπ for any N ≥ 1 and any complex numbers xj with j=1 xj − 1. Since for any Recall now the following easy fact: xj ≤ 1 for every j = 1, . . . , N , we have QN j ≥ k + 1, nj ) = 1 − aj + aje 1 − aj(1 − e 2iπ nk nj ≤ 1 − aj + aj = 1, HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 23 we get that nk 2iπ 1 1 aj1 − e nj ≤ 2, as seen nj ≤ 2π ak+1 Xj≥k+1 µ(nk) − 1 ≤ Xj≥k+1 nk nj ≤ 4π ak+1 since the sequence (aj)j≥1 is decreasing and nkPj≥k+1 nj ≤ nk+1Pj≥k+1 in Proposition 3.8 above. Hence µ(nk) → 1, and this proves Proposition 3.9. Remark 3.10. -- Remark that if nk = 2k for instance, the only λ's in T such that λnk tends to 1 are the 2k th roots of 1. More generally, it is not difficult to see that if nknk+1 and sup nk+1/nk is finite, λnk → 1 if and only if there exists a k0 such that λnk0 = 1. Remark 3.11. -- The proof of Proposition 3.9 yields a bit more, namely that given any sequence (ak)k≥0 of positive numbers decreasing to zero and such that the series P ak diverges, there exists a continuous probability measure σ on T such that σ(nk) − 1 ≤ ak for every k ≥ 0. This will turn out to be crucial in the proof of the statement of Example 3.17. In general one cannot obtain such a measure σ withPσ(nk)− 1 < +∞: this would imply that the series Pλnk − 1 converges σ-a.e., so that λnk − 1 → 0 σ-a.e., and we have seen in Remark 3.10 above that this is impossible if nk+1/nk is bounded for instance. The proof of Proposition 3.9 uses in a crucial way the divisibility assumption on the nk's, and it comes as a natural question to ask whether it can be dispensed with: if there exists an a > 1 such that nk+1/nk ≥ a for any k ≥ 0, must (nk)k≥0 be a rigidity sequence? We were not able to settle this question, but it is answered in [7] in the negative: the sequence (nk)k≥0 with nk = 2k + 1 cannot be a rigidity sequence. Indeed we have 2nk = nk+1 + 1, so that if (nk)k≥0 were a rigidity sequence, with ϕ an associated weakly mixing measure-preserving transformation on (X,F, µ), we should have both U 2nk ϕ → I (SOT) and U nk+1 Obviously a rigidity sequence must have density 0 (this is pointed out already in [20]). Some of the simplest examples of non-rigidity sequences (nk)k≥0 satisfy nk+1/nk → 1. Our three Examples 3.12, 3.13 and 3.14 overlap with examples of [7]. ϕ → I (SOT), so that Uϕ = I which is impossible. Example 3.12. -- Let p ∈ Z[X] be a polynomial with nonnegative coefficients, p 6= 0. Then the sequence (nk)k≥0 with nk = p(k) cannot be a rigidity sequence. This follows directly from Weyl's polynomial equidistribution theorem (see for instance for any irrational number θ ∈ [0, 1], the sequence (p(k)θ)k≥0 is uniformly [25, p. 27]): equidistributed. Hence 1 N N Xk=1 e2iπp(k)θ → 0 as N → +∞ for every θ ∈ [0, 1] \ Q. Hence if σ is any continuous probability measure on T, and this forbids σ(nk) to tend to 1. We have proved in fact: 1 N N Xk=1 σ(nk) → 0, 24 TANJA EISNER & SOPHIE GRIVAUX Example 3.13. -- If there exists a countable subset Q of [0, 1] and a δ > 0 such that for any θ ∈ [0, 1] \ Q, N then (nk)k≥0 is not a rigidity sequence. lim inf N→+∞ 1 N e2iπnkθ ≤ 1 − δ, Xk=1 See [1] for some examples of such sequences. Let us point out that (contrary to what happens for Jamison sequences), it is obvious to exhibit non-rigidity sequences (nk)k≥0 with lim inf nk+1/nk = 1 and lim sup nk+1/nk = +∞: take any sequence (n2k)k≥0 such that n2k+2/n2k → +∞, and set n2k+1 = n2k + 1. If (nk)k≥0 were a rigidity sequence, with ϕ an associated weakly mixing measure-preserving transformation on (X,F, µ), we should have U nk ϕ → I (SOT), so that Uϕ = I, a contradiction. A similar type of argument yields Example 3.14. -- If (nk)k≥0 denotes the sequence of prime numbers, then (nk)k≥0 is not a rigidity sequence. ϕ f − U m Proof. -- This follows from a result of Vinogradov that any sufficiently large odd number can be written as a sum of three primes. Suppose by contradiction that (nk)k≥0 is a rigidity sequence with ϕ an associated weakly mixing measure-preserving transformation ϕ → I (SOT). Let f 6= 0 be a function in L2(X,F, µ) withRX f dµ = on (X,F, µ). Then U nk 0. If ε > 0 is any positive number, let k0 be such that for any k ≥ k0, U nk ϕ f − f < ε and every odd integer greater than or equal to k0 can be written as a sum of three primes. Consider the finite set of integers A = {0, nk1 , nk1 +nk2, nk1 +nk2 +nk3 ; 0 ≤ ki ≤ k0 for i = 1, 2, 3}. We claim that for any odd integer 2n + 1 ≥ k0, there exists an m ∈ A such that U 2n+1 ϕ f < 3ε. Indeed, let us write 2n + 1 as 2n + 1 = nk1 + nk2 + nk3 with 0 ≤ k1 ≤ k2 ≤ k3, and consider separately four cases: -- if k1 > k0, then U nk1 +nk2 +nk3 -- if k1 ≤ k0 and k2 > k0, U ϕ nk1 +nk2 +nk3 -- if k2 ≤ k0 and k3 > k0, U ϕ -- if k3 ≤ k0, there is nothing to prove. Now since ϕ is weakly mixing, U 2n+1 f → 0 (WOT) along a set D which is of density 1 in the set of odd integers. Since A is finite, it follows that there exists some m ∈ A such that kU lj ϕ f − U m ϕ fk < 3ε for an increasing sequence (lj)j≥0 ⊂ D. Thus, for every g ∈ L2(X,F, µ) we have nk1 nk2 ϕ f − f +U ϕ f − f +U nk2 nk1 ϕ f − f + U ϕ f ≤ U nk1 +nk2 ϕ f − f ≤ U f − U f − U nk3 ϕ f − f < 3ε; nk3 ϕ f − f < 2ε; nk3 ϕ f − f < ε; nk1 +nk2 +nk3 ϕ f ≤ U ϕ hU m ϕ f, gi ≤ hU m ϕ f − U lj ϕ f, gi + hU lj ϕ f, gi ≤ 3εkgk + hU lj ϕ f, gi. Taking the weak limit as j → ∞ of the above expression implies kU m a contradiction. ϕ fk ≤ 3ε. Thus f = 0, The proof of Example 3.14 actually shows that if there exists an integer r ≥ 2 such that any sufficiently large integer in a set of positive density can be written as a sum of r elements of the set {nk ; k ≥ 0}, then (nk)k≥0 cannot be a rigidity sequence. As pointed out in [7], the statement of Example 3.14 can also be deduced from the fact that (nkx)k≥0 is uniformly distributed for all but a countable set of values of x ∈ [0, 1]. HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 25 We finish this section with some more examples of rigidity sequences. We consider the sequence (qn)n≥1 of quotients of the convergents of some irrational numbers α ∈]0, 1[. Let α be such a number, and let α = a1 + 1 1 a2 + 1 a3 + . . . with the an's positive integers, be its continued fraction expansion. The convergents of α are the rational numbers pn qn defined recursively by the equations (p0 = 0, p1 = 1, pn+1 = anpn + pn−1 for n ≥ 2 q0 = 1, q1 = a1, qn+1 = anqn + qn−1 for n ≥ 2. See for instance [18] for more about continued fraction expansions and approximations of irrational numbers by rationals. We have (8) α − < 1 qnqn+1 1 2qnqn+1 ≤(cid:12)(cid:12)(cid:12)(cid:12) pn qn(cid:12)(cid:12)(cid:12)(cid:12) for any n ≥ 1. It follows that e2iπqnα − 1 → 0 as n → +∞. Hence there exist infinitely many numbers λ ∈ T \{1} such that λqn − 1 → 0 as n → +∞, and the sequence (qn)n≥1 is a possible candidate for a rigidity sequence. We begin by recalling a particular case of a result of Katok and Stepin [22], see also [32]: Example 3.15. -- If, with the notation above, α − pn qn = o( 1 q2 n ), then (qn)n≥0 is a rigidity sequence. This can also be seen as a direct consequence of our Example 3.4: by the lower bound in (8), the assumption is equivalent to qn+1/qn → +∞ (i.e. an → +∞). It is also possible to show that (qn)n≥0 is a rigidity sequence (and even more) for some irrational numbers α with lim inf an < +∞. For instance: Example 3.16. -- Let m ≥ 2 be an integer, and let αm be the Liouville number m−(k+1)!. αm =Xk≥0 If (qn)n≥1 denotes the sequence of denominators of the convergents of αm, then there exists a perfect compact subset of T on which λqn tends uniformly to 1. In particular (qn)n≥1 is a rigidity sequence. Proof. -- The proof relies on a paper of Shallit [35] where the continued fraction ex- pansion of αm is determined: if [a0, a1, . . . , aNv ] is the continued fraction expansion of k=0 m−(k+1)!, v a nonnegative integer, then the continued fraction expansion of the next Pv partial sum Pv+1 k=0 m−(k+1)! is given by [a0, a1, . . . , aNv+1] = [a0, a1, . . . , aNv , mv(v+1)! − 1, 1, aNv − 1, aNv −1, . . . , a2, a1] 26 TANJA EISNER & SOPHIE GRIVAUX as soon as Nv is even. One has Nv+1 = 2Nv + 2 so that Nv+1 is indeed even. This yields that the continued fraction expansion of αm is [0, m−1, m+1, m2−1, 1, m, m−1, m12−1, 1, m−2, m, 1, m2−1, m+1, m−1, m72−1, 1, . . .]. We have aNv+1 = m(v−1)v! − 1. For any v ≥ 0, qNv + 2 qNv + 1 = m(v−1)v! − 1 + qNv qNv + 1 ≥ m(v−1)v! − 1 ≥ 1 2 m(v−1)v! for v ≥ 2. Applying the proof of Proposition 3.5 to the sequence (nv)v≥0 = (qNv+1)v≥0, we get that there exists a perfect compact subset K of T containing the point 1 such that for any λ ∈ K and any v ≥ 1, λqNv +1 − 1 ≤ 10π sup j≥v qNj−1+1 qNj+1 ≤ 10π qNv+1 qNv+1+1 ≤ 10π qNv+2 qNv+1 ≤ 20π m−(v−1)v!. Let now p be an integer such that Nv−1 + 2 ≤ p ≤ Nv for some v ≥ 0, and λ ∈ K. We need to estimate λp − 1. If p = Nv−1 + 2, we have qNv−1+2 = aNv−1+2 qNv−1+1 + qNv−1 ≤ (aNv−1+2 + 1)qNv−1+1. In the same way qNv−1+3 ≤ (aNv−1+2 + 1)(aNv−1+3 + 1)qNv−1+1 etc., and j (aNv−1+i + 1) qNv−1+1 for any 2 ≤ j ≤ Nv − Nv−1 = Nv−1 + 2. qNv−1+j ≤ Yi=2 So for Nv−1 + 2 ≤ p ≤ Nv we have p−Nv−1 so that (aNv−1+i + 1) qNv−1+1 qp ≤ Yi=2 Nv−1+2 λqp − 1 ≤ (aNv−1+i + 1)λqNv−1+1 − 1. Yi=2 It remains to estimate the quantityQNv−1+2 {1, aNv−1 − 1, aNv−1−1, . . . , a2, a1} so that QNv−1+2 us write Rv−1 =QNv−1 v−1m(v−2)(v−1)! ≤ 21+2R4 Rv ≤ 2 R2 above, i.e. i=2 i=2 (aNv−1+i + 1). We have {aNv−1+2, . . . , aNv} = i=2 (ai + 1). Let i=2 (ai + 1). We have Rv ≤ Rv−1 m(v−2)(v−1)! 2 Rv−1 by the inequality (aNv−1+i + 1) ≤ 2QNv−1 v−2m(v−2)(v−1)!+2(v−3)(v−2)! ≤ . . . ≤ 22v+1 mPv−1 k=1 2k−1(v−(k+1))(v−k)!. Now (v − 1)! ≥ 2k−1(v − k)!, so Rv ≤ 22v+1 m(v−1)(v−2)(v−1)! . Hence λqp − 1 ≤ 22v+1 m(v−1)(v−2)(v−1)!λqNv−1+1 − 1 for any Nv−1 + 2 ≤ p ≤ Nv. Now we have λqNv−1+1 − 1 ≤ 20π m−(v−1)v! so that λqp − 1 ≤ 20π 22v+2 m(v−1)(v−1)!(v−2−v) = 20π 22v+2 m−2(v−1)(v−1)! HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 27 for Nv−1 + 2 ≤ p ≤ Nv. Since the quantity 22v+2 infinity, it follows that λqn tends to 1 uniformly on K as n tends to infinity. m−2(v−1)(v−1)! tends to 0 as v tends to A stronger result is proved in [7]: actually if α is any irrational number in ]0, 1[, the sequence (qn)n≥0 of denominators of the convergents of α is always a rigidity sequence. We finish our study of rigidity sequences by giving an example of a rigidity sequence such that nk+1/nk → 1. This answers a question of [7]. Example 3.17. -- There exists a sequence (nk)k≥0 with nk+1/nk → 1 as k → +∞ which is a rigidity sequence. Proof. -- Let (kp)p≥2 be a very quickly increasing sequence of integers with k1 = 1 which will be determined later on in the proof. For p ≥ 0, let Np = 22p , and consider the set 2kp+2−1 ANp = ANp,k [k=kp+1 where ANp,k = {N k For instance, p , N k p + N k−1 p , N k p + 2N k−1 p , N k p + 3N k−1 p , . . . , N k p + ((Np − 1)Np − 1)N k−1 p }. 2k2−1 2k3−1 A2 = {2k, 2k + 2k−1}, A4 = [k=1 p+1 = N 2kp+2 p [k=k2 {4k, 4k + 4k−1, 4k + 2 4k−1, . . . , 4k + 11 4k−1}, etc. − N 2kp+2−2 which is less than the first element of , these sets are successive and disjoint. Let (nj)j≥0 be the strictly p As the last element of ANp is N 2kp+2 ANp+1, N kp+2 p increasing sequence such that A = Sp≥0 ANp = {nj ; j ≥ 0}. Let us first check that nj+1/nj → 1: first of all, if nj and nj+1 belong to the same set ANp,k, ≤ 1 + p + l N k−1 nj+1 nj 1 Np· N k−1 = 1 + = p p N k p + (l − 1) N k−1 N k p p + (l − 1) N k−1 N k p If nj is in some set ANp,k and nj+1 is in ANp,k+1, nj+1 nj = N k+1 p N k+1 p − N k−1 p = 1 1 − 1 N 2 p = Np+1 Np+1 − 1· Lastly, if nj is the last integer of ANp and nj+1 is the first integer of ANp+1, nj+1 nj = N 2kp+2 p N kp+2 p+1 − N 2kp+2−2 p = N 2kp+2 p p N 2kp+2 − N 2kp+2−2 p = Np+1 Np+1 − 1· Thus nj+1/nj → 1. Let now σ be a continuous probability measure on T such that -- for any 0 ≤ k ≤ 2k2 − 1, σ(2k) − 1 ≤ ak -- for any p ≥ 1 and kp+1 ≤ k ≤ 2kp+2 − 1, p ) − 1 ≤ σ(N k a2kp+1−1 akp+1 ak 28 TANJA EISNER & SOPHIE GRIVAUX where a0 = a1 = 1 and ak = 1 3.9 and Remark 3.11. Indeed the successive terms of the sequence (1, 2, 4, . . . , 22k2−1, 4k2 , 4k2+1, . . .) = (mj)j≥0 k log k for k ≥ 2. Such a measure does exist by Proposition divide each other. The sequence (a0, a1, . . . , a2k2−1, a2k2 −1 ak2 ak2, a2k2 −1 ak2 ak2+1, . . .) = (bj)j≥0 is decreasing to zero, and P bj is divergent: if the sequence (kp) grows fast enough, 2k3−1 2k2−1 bj ≥ Xj≥0 Xk=2 ak + a2k2−1 ak2 Xk=k2 ak + . . . and since the series P ak is divergent, it is possible to choose kp+1 so large with respect to kp that a2kp−1 akp 2kp+1−1 Xk=kp ak ≥ 1 for instance for each p. So we have a probability measure σ on T such that σ(mj)−1 ≤ bj for each j ≥ 0. It remains to show that σ(nk) − 1 → 0. For kp+1 ≤ k ≤ 2kp+2 − 1 and 0 ≤ l ≤ (Np − 1)Np − 1, we have σ(N k p + l N k−1 p ) − 1 ≤ σ(N k ≤ σ(N k p ) − 1 + l σ(N k−1 p ) − 1 + ((Np − 1)Np − 1)σ(N k−1 ) − 1 p p ) − 1. If kp+1 + 1 ≤ k ≤ 2kp+2 − 1, this is less than a2kp+1−1 akp+1 (ak + ((Np − 1)Np − 1)ak−1) ≤ (Np − 1)Np a2kp+1−1 akp+1 akp+1 ≤ (Np − 1)Np a2kp+1−1. If kp+1 is large enough compared to Np, this quantity is less than 2−p. N kp+1−1 = N 2kp+1−2 . Hence p p−1 σ(N kp+1 p + l N kp+1−1 p ) − 1 ≤ a2kp+1−1 + ((Np − 1)Np − 1) a2kp−2 ≤ (1 + ((Np − 1)Np − 1) akp a2kp−2 akp )a2kp+1−2 If k = kp+1, a2kp+1−2 and this again can be made less than 2−p provided kp+1 is sufficiently large with respect to Np and kp. Hence σ(nk) → 1 as k → +∞, and this proves that (nk)k≥0 is a rigidity sequence. 4. Topologically and uniformly rigid linear dynamical systems 4.1. Back to rigidity in the linear framework. -- Before moving over to topological versions of rigidity for linear dynamical systems, we have to settle the following natural question: which sequences (nk)k≥0 appear as rigidity sequences (in the measure-theoretic sense) for linear dynamical systems? Here is the answer: Theorem 4.1. -- Let (nk)k≥0 be an increasing sequence of positive integers. The follow- ing assertions are equivalent: HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 29 (1) there exists a continuous probability measure σ on T such that σ(nk) → 1 as k → +∞, i.e., (nk)k≥0 is a rigidity sequence; (2) there exists a bounded linear operator T on a separable complex infinite-dimensional Hilbert space H which admits a non-degenerate Gaussian measure m with respect to which T defines a weakly mixing measure-preserving transformation which is rigid with respect to (nk)k≥0. Proof. -- The implication (2) ⇒ (1) is an immediate consequence of Proposition 3.1, so let us prove that (1) ⇒ (2). Let σ be a continuous probability measure σ on T such that σ(nk) → 1 as nk → +∞, and let L ⊆ T be the support of the measure σ. It is a compact perfect subset of T, and σ(Ω) > 0 for any non-empty open subset Ω of L. Kalish constructed in [21] an example of a bounded operator on a Hilbert space whose point spectrum is equal to L, and, as in [3], we use this example for our purposes: let T0 be the operator defined on L2(T) by T0 = M−J, where M f (λ) = λf (λ) and Jf (λ) =R(1,λ) f (ζ)dζ for any f ∈ L2(T) and λ ∈ T. For λ ∈ T, λ = eiθ, (1, λ) denotes the arc {eiα ; 0 ≤ α ≤ θ}, and (λ, 1) the arc {eiα ; θ ≤ α ≤ 2π}. For every λ, the characteristic function χλ of the arc (λ, 1) is an eigenvector of T0 associated to the eigenvalue λ. Let T be the operator induced by T0 on the space H = sp[χλ ; λ ∈ L]. It is proved in [21] that σ(T ) = σp(T ) = L, and it is not difficult to see that E : λ 7→ χλ is a continuous eigenvector field for T on L which is spanning. Hence it is a perfectly spanning unimodular eigenvector field with respect to the measure σ (see [3] for details), and there exists a non-degenerate Gaussian measure m on H whose covariance operator S is given by S = KK ∗, where K : L2(T, σ) → H is the operator defined by Kϕ = RT ϕ(λ)E(λ)dσ(λ) for ϕ ∈ L2(T, σ), with respect to which T defines a weakly mixing measure-preserving transformation. It remains to prove that T is rigid with respect to (nk)k≥0, i.e. that U nk T f tends weakly to f in L2(H,B, m). Using the same kind of arguments as in [4] or [6, Ch. 5], we see that it suffices to prove that for any elements x, y of H, ZH hx, T nk zihy, zidm(z) −→ZH hx, zihy, zidm(z) as nk → +∞. But this is clear: since T K = KV , where V is the multiplication operator by λ on L2(T, σ), we have ZH hx, T nk zihy, zidm(z) = hKK ∗T ∗nk x, yi = hV ∗nk K ∗x, K ∗yi λ−nkhx, E(λ)ihy, E(λ)idσ(λ). = ZT The function h(λ) = hx, E(λ)ihy, E(λ)i belongs to L1(T, σ), and we have seen in the proof of Proposition 3.1 that RT λnk − 1h(λ)dσ(λ) → 0. Hence ZT λ−nk h(λ)dσ(λ) →ZT h(λ)dσ(λ) =ZH hx, zihy, zidm(z), and this proves our statement. Remark 4.2. -- The Kalish-type operators which are used in the proof of Theorem 4.1 have no reason at all to be power-bounded with respect to (nk), contrary to what happens 30 TANJA EISNER & SOPHIE GRIVAUX when considering topological rigidity. We only know for instance, applying the rigidity assumption to the function f (z) = z, that ZH (T nk − I)z dm(z) → 0 as nk → +∞. Remark 4.3. -- Theorem 4.1 gives another proof of the characterization of rigidity se- quences obtained in Proposition 3.1. 4.2. A characterization of topologically rigid sequences for linear dynamical systems. -- Let us prove Theorem 1.12. First of all, (3) implies (2) since, as recalled in Section 2.1, (3) implies that T has perfectly spanning unimodular eigenvectors. We suppose next that (2) holds and show (1). Let X and T be as in (2). For any λ ∈ σp(T )∩T, let eλ be an associated eigenvector with eλ = 1. Since T nk eλ → eλ, λnk − 1 → 0 for any λ ∈ σp(T ) ∩ T. Moreover, by the uniform boundedness principle, supk≥0 T nk = M is finite. Suppose by contradiction that there exists an ε0 > 0 such that for any λ, µ ∈ σp(T ) ∩ T with λ 6= µ, supk≥0 λnk − µnk ≥ ε0. Then for any λ, µ ∈ σp(T ) ∩ T, λnk − µnk − eλ − eµ ≤ λnk eλ − µnk eµ ≤ M eλ − eµ so that λnk − µnk ≤ (M + 1)eλ− eµ. Hence ε0 ≤ (M + 1)eλ− eµ, and the unimodular eigenvectors of T are ε0/(M + 1)-separated. Since X is separable there can only be countably many such eigenvectors, which contradicts the fact that σp(T )∩T is uncountable. So for any ε > 0 there exist λ, µ in σp(T )∩ T with λ 6= µ such that supk≥0 (λµ)nk − 1 ≤ ε, and (λµ)nk − 1 → 0. So (1) holds true. We state again what we have to prove in order to obtain that (1) implies (3): Theorem 4.4. -- Let (nk)k≥0 be an increasing sequence of integers with n0 = 1 such that for any ε > 0 there exists a λ ∈ T \ {1} with k≥0 λnk − 1 ≤ ε sup and λnk − 1 → 0 as k → ∞. Then there exists a bounded linear operator T on a Hilbert space H such that T has a perfectly spanning set of eigenvectors associated to unimodular eigenvalues and for every x ∈ H, T nk x → x as k → ∞. Before starting the proof of Theorem 4.4, let us point out that the statement is not true anymore if we only suppose that there exists a λ ∈ T\{1} such that λnk−1 → 0: if (qn)n≥0 is the sequence of denominators of the partial quotients in the continued fraction expansion of α = √2 for instance, λ = e2iπα is such that λqn − 1 → 0. But the sequence ( qn+1 )n≥0 qn is bounded (see for instance [18]), so that (qn)n≥0 is not even a Jamison sequence. Proof of Theorem 4.4. -- We take the same kind of operator as in the proof of Theorem 2.1, and show that under the assumptions of Theorem 1.12, such an operator T = D + B is such that T np − Dnp tends to 0 as np tends to infinity. Before starting on this, we take advantage of the assumption of the theorem to construct a particular perfect compact subset of T, in which our coefficients λl will be chosen later in the proof: HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 31 Lemma 4.5. -- Under the assumptions of Theorem 4.4, there exists a perfect compact subset K of T such that (K, d(nk )) is separable and for any λ ∈ K, λnk − 1 → 0 as nk → +∞. Proof of Lemma 4.5. -- The proof proceeds along the same lines as in [2]: let (µn)n≥1 be a sequence of elements of T \ {1} such that d(nk)(µ1, 1) < 4−1 , d(nk)(µn, 1) < 4−nd(nk)(µn−1, µn−1) for any n ≥ 2, d(nk)(µn, µn) decreases with n, and moreover µnk n − 1 → 0 as nk → +∞. If (s1, . . . , sn) is any finite sequence of zeroes and ones, we associate to it an element λ(s1,...,sn) of T in the following way: we start with λ(0) = µ1 and λ(1) = µ1, and we have Then if λ(s1,...,sn−1) has already been defined, we set d(nk)(λ(0), λ(1)) = d(nk)(µ1, µ1) > 0. λ(s1,...,sn−1,0) = λ(s1,...,sn−1)µn and λ(s1,...,sn−1,1) = λ(s1,...,sn−1)µn. We have and d(nk)(λ(s1,...,sn−1), λ(s1,...,sn−1,sn)) < 4−nd(nk)(µn−1, µn−1) d(nk)(λ(s1,...,sn−1,0), λ(s1,...,sn−1,1)) = d(nk)(µn, µn), so that for any infinite sequence s = (s1, s2, . . .) of zeroes and ones, we can define λs ∈ T as λs = limn→+∞ λ(s1,...,sn). It is not difficult to check (see [2] for details) that the map s 7−→ λs from 2ω into T is one-to-one, so that K = {λs ; s ∈ 2ω} is homeomorphic to the Cantor set, hence compact and perfect, and that (K, d(nk )) is separable. It remains to see that for any s ∈ 2ω, λnk s − 1 → 0 as nk → +∞. We have for any p ≥ 1 λs = λ(s1,...,sp)Yj≥p λ(s1,...,sj+1)λ(s1,...,sj), so that for any p ≥ 1, λnk s − 1 = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ λnk s − 1 ≤ λnk λnk ≤ λnk ≤ λnk = λnk Hence nk (s1,...,sj) − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (s1,...,sj)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . (s1,...,sj+1) − λnk λnk λnk λnk (s1,...,sj+1)λ (s1,...,sp)Yj≥p (s1,...,sp) − 1 +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Yj≥p (s1,...,sp) − 1 +Xj≥p (s1,...,sp) − 1 + 2Xj≥p (s1,...,sp) − 1 + 2 d(nk )(µp, µp)Xj≥p (s1,...,sp) − 1 + 4−pd(nk)(µp, µp). 2 3 4−(j+1)d(nk)(µj, µj) 4−(j+1) d(nk)(λ(s1,...,sj+1), λ(s1,...,sj)) 32 TANJA EISNER & SOPHIE GRIVAUX Given any γ > 0, take p such that the second term is less than γ/2. Since µnk as nk → +∞, λnk there exists an integer k0 ≥ 1 such that for any k ≥ k0, λnk any k ≥ k0 and any s ∈ 2ω, we have λnk λnk − 1 → 0 as nk → +∞. Let us now go back to the proof of Theorem 1.12. We have seen that n − 1 → 0 (s1,...,sp) − 1 → 0 as nk → +∞ for any finite sequence (s1, . . . , sp). Hence (s1,...,sp) − 1 ≤ γ/2. Thus for s − 1 < γ. So we have proved that for any λ ∈ K, l−1 T np − Dnp2 ≤Xl≥2 Xk=max(1,l−np) t(np) k,l 2, and that it is possible for each l ≥ 2 to take λl with d(np)(λl, λj(l)) so small that l−1 Xk=max(1,l−np) t(np) k,l 2 ≤ 2−l. So we do the construction in this way with the additional requirement that for each l ≥ 1, λl is such that λnp l − 1 → 0 as np → +∞ (this is possible by Lemma 4.5). Let now ε > 0 and l0 ≥ 2 be such that Pl≥l0+1 2−l < ε T np − Dnp2 ≤ 2 . We have for any p such that np ≥ l0 + 1 t(np) 2 + ε 2· l−1 k,l l0 Xl=2 Xk=1 The proof will be complete if we show that for any k, l with 1 ≤ k ≤ l − 1, t(np) as np → +∞, or, equivalently, that s(np) k,l → 0. Recall that by Lemma 2.7, s(np) λnp+1−(l−k) l written as l−1 k,l k,l → 0 can be − λnp+1−(l−k) j Thus we have to show that sk,l = 0 for any 1 ≤ k ≤ l − 1. This is a consequence of the following lemma: Lemma 4.6. -- For any k, l with 1 ≤ k ≤ l − 1 and any p with 0 ≤ p ≤ l − k − 1, we have l−1 Xj=k λ1−(l−k−p) l − λ1−(l−k−p) j = 0. c(k,l) j λl − λj Proof. -- The proof is done by induction on l ≥ 2. For l = 2, we just have to check that c(1,2) 1 1 2 − λ0 λ0 λ2 − λ1 = 0, c(k,l) j s(np) k,l = Xj=k as soon as np ≥ l − k. Since λnp j → 1 for any j ≥ 1, Xj=k s(np) k,l → sk,l := λ1−(l−k) l λl − λj c(k,l) j l−1 λl − λj − λ1−(l−k) j as np → +∞. HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 33 which is obviously true. Supposing now that the induction assumption is true for some l ≥ 2, consider k with 1 ≤ k ≤ l and p with 0 ≤ p ≤ l − k. Then l is equal to c(k,l+1) j Xj=k l λ−(l−k−p) l+1 − λ−(l−k−p) j λl+1 − λj c(k,l+1) j λ−(l−k−p) j λ(l−k−p) l+1 − λ(l−k−p) j l+1 −λ−(l−k−p) Xj=k = −λ−(l−k−p) l+1 = −λ−1 l+1 l Xj=k Xm=0 l−k−p−1 l Xj=k c(k,l+1) j λ−(l−k−p) j j λl−k−p−1−m λm l+1 l−k−p−1 λl+1 − λj Xm=0 λ−m l+1  c(k,l+1) j λ−(l−k−p−m) j l Xj=k  · c(k,l+1) j λ−(l−k−p−m) j It suffices now to show that each sum is equal to 0. We have seen in the proof of Lemma 2.7 that for 1 ≤ k ≤ l − 1, λlc(k,l) λl − λj for k ≤ j ≤ l − 1 λjc(k,l) λl − λj and c(k,l+1) c(k,l+1) j = − l−1 = j j l Xj=k and that c(l,l+1) l c(k,l+1) j l Xj=k = 1. Thus for 1 ≤ k ≤ l − 1 c(k,l) j λl − λj = − l−1 λ−(l−k−p−m) j λ−(l−k−p−m−1) j + l−1 Xj=k c(k,l) j λl − λj λ−(l−k−p−m−1) l l−1 Xj=k Xj=k = c(k,l) j λ1−(l−k−p−m) l − λ1−(l−k−p−m) j · λl − λj l = 0. c(l) l,l+1 Since p + m ≤ l − k − 1, this quantity vanishes by the induction assumption. For k = l, we only have to consider the case p = 0, and here l+1 − λ0 λ0 λl+1 − λl This finishes the proof of Lemma 4.6. We have shown that t(np) k,l → 0 for each 1 ≤ k ≤ l − 1, and it follows immediately that T np − Dnp → 0 as np → +∞. Now if x ∈ H and ε is any fixed positive number, take l0 such that Pl≥l0+1 xl2 ≤ ε/2. Since Dnpx − x2 ≤ l0 Xl=1 l − 12! x2 + 2 Xl≥l0+1 λnp xl2 34 TANJA EISNER & SOPHIE GRIVAUX l − 1 tends to 0 for each l ≥ 2, it follows that Dnpx − x → 0 as np → +∞ for and λnp any x ∈ H, hence T npx − x → 0 which is the conclusion of Theorem 1.12. 4.3. A characterization of uniformly rigid sequences for linear dynamical sys- tems. -- We now prove Theorem 1.13. Clearly (3) ⇒ (2). The implication (2) ⇒ (1) is obvious: using the notation of Section 4.2 above, T nk eλ − eλ tends to 0 uniformly on σp(T ) ∩ T =: K which is uncountable, i.e. λnk − 1 tends to 0 uniformly on K. The converse implication (1) ⇒ (3) follows from Fact 3.7, the proof of Theorem 4.4 above and Lemma 4.7 below. First replacing K by a compact perfect subset of its closure, and then using Fact 3.7, we can suppose that K is such that λnk − 1 tends to 0 uniformly on K and for any ε > 0 there exists a ν ∈ K \ {1} such that supk≥0 νnk − 1 ≤ ε. Then Lemma 4.7. -- Under the assumption above on K, there exists a perfect compact subset K ′ of T such that (K ′, d(nk )) is separable and λnk tends to 1 uniformly on K ′. Proof of Lemma 4.7. -- The proof runs along the same lines as that of Lemma 4.5: we start from elements µn, n ≥ 1, of K \ {1} having the same properties as in Lemma 4.5 (which we know exist - this is why we had to use Fact 3.7), and we construct the unimodular numbers λs, s ∈ 2ω, as in Lemma 4.5, with s − 1 ≤ λnk λnk (s1,...,sp) − 1 + 4−pd(nk)(µp, µp) 2 3 p − 1 + γ 2 ≤ µnk (s1,...,sp) − 1 + for any k ≥ 0, p ≥ 1, and s ∈ 2ω. Let γ > 0, and take p such that the second term is less than γ/2. Then for any s ∈ 2ω and any k ≥ 0 we have γ γ s − 1 ≤ λnk 1 − 1 + . . . + µnk 2 ≤ pλnk − 1∞,K + λnk 2· s −1 ≤ γ for any s ∈ 2ω Take κ such that for any k ≥ κ, λnk−1∞,K ≤ γ/(2p): we have λnk and k ≥ κ, and this shows that λnk tends to 1 uniformly on the set K ′ = {λs ; s ∈ 2ω}. Since (K ′, d(nk)) is separable, Lemma 4.7 is proved. Now in the construction of the operator T , we choose the coefficients λl in the set K ′ given by Lemma 4.7. We have seen in the proof of Theorem 1.12 above that T nk − Dnk tends to 0 as nk tends to infinity. So it suffices to prove that with the additional uniformity assumption of Theorem 1.13, Dnk − I = supl≥1 λnk l − 1 tends to 0 as nk tends to infinity. But Dnk − I ≤ λnk − 1∞,K ′ which tends to 0, so our claim is proved. Proof of Corollary 1.14. -- If (nk)k≥0 is any sequence with nk+1/nk → +∞, or if nknk+1 for any k ≥ 0 and lim sup nk+1/nk = +∞, we have seen in Propositions 3.5 and 3.8 that Theorem 1.13 applies, proving Corollary 1.14. Theorem 1.13 also applies to the sequences (qn)n≥0 considered in Example 3.16. We thus obtain examples, in the linear framework, of measure-preserving transformations on a Hilbert space which are both weakly mixing in the measure-theoretic sense and uniformly rigid. Remark 4.8. -- If (nk)k≥0 is such that nknk+1 for any k ≥ 0 and lim sup nk+1/nk = +∞, the proof of Proposition 3.8 shows that the set K = {λε ; ε ∈ {0, 1}N} contains a dense subset of numbers λ which are N th roots of 1 for some N ≥ 1. Hence, in all the constructions of operators T = D + B considered here, it is possible to choose the HILBERTIAN JAMISON SEQUENCES AND RIGID DYNAMICAL SYSTEMS 35 numbers λl, l ≥ 1, as being N th roots of 1. In this way the operator T becomes additionally chaotic (i.e. it is topologically transitive and has a dense set of periodic vectors). This gives further examples of chaotic operators which are not topologicaly mixing (the first examples of such operators were given in [1]), and shows in particular that there exist chaotic operators which are uniformly rigid. Acknowledgements. We are deeply grateful to Jean-Pierre Kahane for his kind help with the proof of Proposition 3.9. We are also grateful to Mariusz Lema´nczyk, Herv´e Queff´elec, Martine Queff´elec and Maria Roginskaya for helpful discussions. References [1] C. Badea, S. Grivaux, Unimodular eigenvalues, uniformly distributed sequences and linear dynamics, Adv. Math. 211 (2007), pp 766 -- 793. [2] C. Badea, S. Grivaux, Size of the peripheral point spectrum under power or resolvent growth conditions, J. Funct. Anal. 246 (2007), pp 302 -- 329. [3] F. Bayart, S. Grivaux, Hypercyclicity and unimodular point spectrum, J. Funct. Anal. 226 (2005), pp 281 -- 300. [4] F. Bayart, S. Grivaux, Frequently hypercyclic operators, Trans. Amer. Math. Soc. 358 (2006), pp 5083 -- 5117. [5] F. Bayart, S. Grivaux, Invariant Gaussian measures for operators on Banach spaces and linear dynamics, Proc. Lond. Math. Soc. 94 (2007), pp 181 -- 210. [6] F. Bayart, ´E. Matheron, Dynamics of linear operators, Cambridge University Press 179 (2009). [7] V. Bergelson, A. Del Junco, M. Lema´nczyk, J. Rosenblatt, Rigidity and non- recurrence along sequences, preprint 2011. [8] M. Boshernitzan, E. Glasner, On two recurrence problems, Fund. Math. 206 (2009), pp 113 -- 130. [9] I. P. Cornfeld, S. V. Fomin, Y. G. Sinai, Ergodic theory, Grundlehren der Mathe- matischen Wissenschaften 245, Springer-Verlag, New York, 1982. [10] M. De la Rosa, L. Frerick, S. Grivaux, A. Peris, Frequent hypercyclicity, chaos, and unconditional Schauder decompositions, to appear in Israel J. Math.. [11] E. Flytzanis, Unimodular eigenvalues and linear chaos in Hilbert spaces, Geom. Funct. Anal. 5 (1995), pp 1 -- 13. [12] H. Furstenberg, B. Weiss, The finite multipliers of infinite ergodic transformations, in "The structure of attractors in dynamical systems" (Proc. Conf., North Dakota State Univ., Fargo, N.D., 1977), pp. 127 -- 132, Lecture Notes in Math. 668, Springer, Berlin, 1978. [13] J. Galambos, I. K´atai, A simple proof for the continuity of infinite convolutions of binary random variables, Statist. Probab. Lett. 7 (1989), pp 369 -- 370. [14] S. Glasner, D. Maon, Rigidity in topological dynamics, Erg. Th. Dyn. Syst. 9 (1989), pp 309 -- 320. [15] E. Glasner, Ergodic theory via joinings, Mathematical Surveys and Monographs 101, American Mathematical Society, Providence, RI, 2003. [16] E. Glasner, Classifying dynamical systems by their recurrence properties, Topol. Meth- ods Nonlinear Anal. 24 (2004), pp 21 -- 40. [17] S. Grivaux, A new class of frequently hypercyclic operators, to appear in Indiana Univ. Math. J.. 36 TANJA EISNER & SOPHIE GRIVAUX [18] G, H. Hardy, E. M. Wright, An introduction to the theory of numbers, Oxford University Press, Oxford, 2008. [19] B. Jamison, Eigenvalues of modulus 1, Proc. Amer. Math. Soc. 16 (1965), pp 375 -- 377. [20] J. James, T. Koberda, K. Lindsey, C. E. Silva, P. Speh, On ergodic transformations that are both weakly mixing and uniformly rigid, New York J. Math. 15 (2009), pp 393 -- 403. [21] G. Kalish, On operators on separable Banach spaces with arbitrary prescribed point spectrum, Proc. Amer. Math. Soc. 34 (1972), pp 207 -- 208. [22] A. B. Katok, A. M. Stepin, Approximations in ergodic theory, Russian Math. Surveys 22 (1967), pp 77 -- 102. [23] A. Katok, Approximation and generity in abstract ergodic theory, Notes 1985. [24] T. W. Korner, Recurrence without uniform recurrence, Ergodic Theory Dynam. Sys- tems 7 (1987), pp 559 -- 566. [25] L. Kuipers, H. Niederreiter, Uniform distribution of sequences. Pure and Applied Mathematics, Wiley-Interscience, New York-London-Sydney, 1974. [26] M. Lema´nczyk, C. Mauduit, Ergodicity of a class of cocycles over irrational rotations, J. London Math. Soc. 49 (1994), pp 124 -- 132. [27] M. G. Nadkarni, Spectral Theory of Dynamical Systems, Birkhauser Advanced Texts: Basel Textbooks, Birkhauser Verlag, Basel, 1998. [28] N. Nikolskii, Selected problems of weighted approximation and spectral analysis, Pro- ceedings of the Steklov Institute of Mathematics, 120 (1974), American Mathematical Society, Providence, 1976. [29] V. Peller, Hankel operators and their applications, Springer Monographs in Mathemat- ics, Springer-Verlag, New-York, 2003. [30] K. Petersen, Ergodic Theory, Cambridge Studies in Advanced Mathematics, Cambridge University Press, 1983. [31] K. Petersen, Lectures on ergodic theory, http://www.math.unc.edu/Faculty/petersen. [32] M. Queff´elec, Substitution dynamical systems -- spectral analysis, second edition, Lec- ture Notes in Mathematics 1294, Springer-Verlag, Berlin, 2010. [33] T. Ransford, Eigenvalues and power growth, Israel J. Math. 146 (2005), pp 93 -- 110. [34] T. Ransford, M. Roginskaya, Point spectra of partially power-bounded operators, J. Funct. Anal. 230 (2006), pp 432 -- 445. [35] J. Shallit, Simple continued fractions for some irrational numbers, II, J. Numb. Th. 14 (1982), pp 228 -- 231. [36] A. M. Stepin, Spectral properties of generic dynamical systems, Izv. Akad. Nauk SSSR Ser. Mat. 50 (1986), pp 801 -- 834. [37] B. Sz.-Nagy, C. Foias¸, Sur les contractions de l'espace de Hilbert, IV, Acta Sci. Math. Szeged 21 (1960), pp 251 -- 259. [38] P. Walters, An Introduction to Ergodic Theory, Graduate Texts in Mathematics 79, Springer-Verlag, New-York, Berlin, 1982. Tanja Eisner, Korteweg de Vries Institut voor Wiskunde, Universiteit van Amsterdam, P.O. Box 94248, 1090 GE Amsterdam, The Netherlands • E-mail : [email protected] Sophie Grivaux, Laboratoire Paul Painlev´e, UMR 8524, Universit´e Lille 1, Cit´e Scientifique, 59655 Villeneuve d'Ascq Cedex, France • E-mail : [email protected]
1811.08629
1
1811
2018-11-21T08:32:39
On The Grand Wiener Amalgam Spaces
[ "math.FA" ]
In this article, notations are included in Section 1. In Section 2, we define the grand Wiener amalgam space by using the classical Wiener amalgam space [9, 15, 16, 17] and the generalized grand Lebesgue space [18, 13] . Section 3, concerns the inclusions between these spaces and some applications. In last section Section 4, we prove the Holders inequality for grand Wiener amalgam space. We also find the associate space and dual of this space, and we prove that the grand Wiener amalgam space is not reflexive.
math.FA
math
On The Grand Wiener Amalgam Spaces A.Turan Gurkanlı Abstract. In this article, notations are included in Section 1. In Section 2, we define the grand Wiener amalgam space by using the classical Wiener amalgam space [9, 15, 16, 17] and the generalized grand Lebesgue space [18, 13] . Section 3, concerns the inclusions between these spaces and some applications. In last section Section 4, we prove the Holder's inequality for grand Wiener amalgam space. We also find the associate space and dual of this space, and we prove that the grand Wiener amalgam space is not reflexive. 1. Notations Let Ω be locally compact hausdorff space and let (Ω, B, µ) be finite Borel measure space. The grand Lebesgue space Lp) (µ) was introduced in [18] by the norm kf kp) = sup 0<ε≤p−1(cid:18)εZΩ f p−ε dµ(cid:19) 1 p−ε where 1 < p < ∞. This is a Banach space. For 0 < ε ≤ p − 1, Lp (µ) ⊂ Lp) (µ) ⊂ Lp−ε (µ) hold. For some properties and applications of Lp) (µ) spaces we refer to papers [1] , [4] , [6] , [7] , [12]and [13] . A generalization of the grand Lebesgue spaces are the spaces Lp),θ (µ) , θ ≥ 0, defined by the norm (see [1] , [13]) kf kp),θ,µ = kf kp),θ = sup 0<ε≤p−1 ε θ p−ε (cid:18)ZΩ f p−ε dµ(cid:19) 1 p−ε = sup ε 0<ε≤p−1 θ p−ε kf kp−ε ; when θ = 0 the space Lp),0 (µ) reduces to Lebesgue space Lp (µ) and when θ = 1 the space Lp),1 (µ) reduces to grand Lebesgue space Lp) (µ) . We have for all 1 < p < ∞ and ε > 0 Lp (µ) ⊂ Lp),θ (µ) ⊂ Lp−ε (µ) . Different properties and applications of these spaces were discussed in [1] , [13] and [4] . 2000 Mathematics Subject Classification. Primary 46E30; Secondary 46E35; 46B70. Key words and phrases. Grand Lebesgue space, generalized grand Lebesgue space, grand Wiener Amalgam space. 1 2 A.TURAN G URKANLI Let 1 ≤ p < ∞, θ ≥ 0, and J be the one of the set of N, Z or Zn. We define the grand Lebesgue sequence space ℓp),θ = ℓp),θ (J) by the norm kukℓp),θ (J) = sup 0<ε≤p−1 εθXk∈J 1 p−ε ukp−ε! = sup ε 0<ε≤p−1 θ p−ε kukℓp−ε(J) . ′ Let p = p p−1 , 1 < p < ∞. First consider an auxiliary space namely L(p′,θ (Ω) , θ > 0, defined by kgk(p′ ,θ = inf ∞ g= Xk=1 ( ∞ Xk=1 gk inf 0<ε≤p−1 ε− θ p−ε (cid:18)ZΩ ′ gk(p−ε) dx(cid:19) 1 (p−ε) ′) where the functions gk, k ∈ N, being in M0, the set of all real valued measurable functions, finite a.e. in Ω. After this definition the generalized small Lebesgue spaces have been defined by where ′ Lp) ,θ (Ω) =ng ∈ M0 : kgkp)′ ,θ < +∞o , kgkp)′ ,θ = sup 0≤ψ≤g ψ∈L(p′,θ kψk(p′ ,θ . For θ = 0 it is kf k(p′ ,0 = kf kp′ ,θ , [5] , [11] . tions f : Ω → C such that f χK ∈ Lp), for any compact subset K ⊂ Ω,where χK is the characteristic function of f. It is a topological vector space with the family of Let 1 ≤ p, q ≤ ∞. The space (cid:0)Lp)(cid:1)locconsists of (classes of) measurable func- seminorms f → kf kp) . Since Lp ⊂ Lp), it is easy to show that (Lp)loc ֒→(cid:0)Lp)(cid:1)loc . 2. The Grand Wiener Amalgam Space Definition 1. Let Ω be locally compact hausdorff space and let (Ω, B, µ) be finite Borel measure space. Also assume that 1 < p, q < ∞ and Q ⊂ Ω is a fix compact subset with nonemty interior. The grand Wiener amalgam space W(cid:0)Lp),θ1 , Lq),θ2(cid:1) consists of all functions (classes of ) f ∈(cid:0)Lp),θ1(cid:1)loc such that the control function F p),θ1 f (x) = F p),θ1,Q f lies in Lq),θ2 . The norm of W (cid:0)Lp),θ1 , Lq),θ2(cid:1) defined by (x) =(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ1 (cid:13)(cid:13)(cid:13)q),θ2 F p),θ1 f kf kW (Lp),θ1 ,Lq),θ2) =(cid:13)(cid:13)(cid:13) = sup ε 0<ε≤p−1 θ1 p−ε (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ1(cid:13)(cid:13)(cid:13)q),θ2 . Since generalized grand Lebesgue spaces are not translation invariant, this re- is depend of the choice of Q. We give the following theorem for the independedness. sult reflects to the grand Wiener amalgam spaces. Then the definition of W (cid:0)Lp),θ1 , Lq),θ2(cid:1) W(cid:0)Lp),θ1 , Lq),θ2(cid:1) is independent of the choice of Q, i.e., different choices of Q Theorem 1. if θ2 = 0 then the definition of grand Wiener amalgam space define the same space with equivalent norms. ON THE GRAND WIENER AMALGAM SPACE 3 Proof. We know that when θ2 = 0, the generalized grand Lebesgue space reduces to lebesgue space Lp (Ω) . Then if θ2 = 0, the grand Wiener amalgam space W (cid:0)Lp),θ1, Lq),θ2(cid:1) reduces to W (cid:0)Lp),θ1, Lq(cid:1) . Since Lq (Ω) is solid and strongly translation invariant, by using the proof technic in Proposition 11.3.2 (b) in [15] or Theorem 2 in [9] , the proof is completed. (cid:3) function space (shortly BF-space), namely its norm satisfies the following proper- Theorem 2. Let 1 < p, q < ∞, and θ ≥ 0. Then W (cid:0)Lp) ,θ, Lq),θ(cid:1) is a Banach ties, where f, g and fn are in W(cid:0)Lp),θ, Lq),θ(cid:1) , λ ≥ 0 and E is a measurable subset of Ω : 1. kf kW (Lp),θ,Lq),θ) ≥ 0 2. kf kW (Lp),θ,Lq),θ) = 0 if and only if f = 0 a.e in Ω 3. kλf kW (Lp),θ ,Lq),θ) = λ kf kW (Lp),θ,Lq),θ) 4. kf + gkW (Lp),θ,Lq),θ) ≤ kf kW (Lp),θ,Lq),θ) + kgkW (Lp),θ,Lq),θ) 5. if g ≤ f a.e. in Ω, then kgkW (Lp),θ,Lq),θ) ≤ kf kW (Lp),θ,Lq),θ) 6. if 0 ≤ fn ↑ f a.e. in Ω, then kfnkW (Lp),θ,Lq),θ) ↑ kf kW (Lp),θ ,Lq),θ) 7. kχEkW (Lp),θ ,Lq),θ) < +∞ f dx ≤ C (p, θ, E) kf kW (Lp),θ,Lq),θ) 8.ZE for some 0 < C < ∞. Proof. The first three properties follow from the definition of the norm k.kW (Lp),θ,Lq),θ) . = kf kW (Lp),θ,Lq),θ) + kgkW (Lp),θ,Lq),θ) . (x)(cid:13)(cid:13)(cid:13)q),θ Proof of property 4. Let f, g ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) . Then kf + gkW (Lp),θ,Lq),θ) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(f + g) χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x + gχQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ +(cid:13)(cid:13)gχQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ +(cid:13)(cid:13)(cid:13) (x)(cid:13)(cid:13)(cid:13)q),θ ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p),θ ≤ (cid:13)(cid:13)(cid:13) (x)(cid:13)(cid:13)(cid:13)q),θ Proof of property 5. Let g ≤ f . Then g χQ+x ≤ f χQ+x a.e. in Ω. Since =(cid:13)(cid:13)(cid:13) (x) + F p),θ F p),θ F p),θ g F p),θ f f g Lp),θ (Ω) is a BF-space on Ω, we have F p),θ g (x) =(cid:13)(cid:13)gχQ+x(cid:13)(cid:13)p),θ ≤(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p),θ Thus we obtain kgkW (Lp),θ,Lq),θ) ≤ kf kW (Lp),θ,Lq),θ) . = F p),θ f (x) . Proof of property 6. Since 0 ≤ fn ↑ f a.e. in Ω, then fnχQ+x ↑ f χQ+x a.e. in Ω. Then by proposition 2.1, in [1] . One more applaying this proposition we have F p),θ fn (x) =(cid:13)(cid:13)fnχQ+x(cid:13)(cid:13)p),θ kfnkW (Lp),θ,Lq),θ) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)fnχQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ f ↑(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p),θ ↑(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ = F p),θ (x) = kf kW (Lp),θ,Lq),θ) . 4 A.TURAN G URKANLI Proof of property 7. Since ε θ p−ε is increasing in [0, p − 1] , and µ (Ω) is finite, then (cid:13)(cid:13)χEχQ+x(cid:13)(cid:13)W (Lp),θ ,Lq),θ) (1) θ p−ε is 1 1 . sup sup sup ε ε ε θ θ θ 0<ε≤p−1 0<ε≤p−1 p−ε [µ (Ω)] =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)χE∩Q+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ p−ε(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)q),θ = (cid:13)(cid:13)χE∩Q+x(cid:13)(cid:13)W (Lp),θ,Lq),θ) = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13) p−ε (cid:13)(cid:13)χE∩Q+x(cid:13)(cid:13)p−ε(cid:13)(cid:13)(cid:13)(cid:13)q),θ p−ε dµ(cid:27) p−ε (cid:26)ZΩ(cid:12)(cid:12)χE∩Q+x(cid:12)(cid:12) p−ε(cid:13)(cid:13)(cid:13)(cid:13)q),θ p−ε(cid:13)(cid:13)(cid:13)(cid:13)q),θ p(cid:13)(cid:13)(cid:13)q),θ p for all 0 < ε ≤ p − 1. Since ε (p − 1)θ [µ (Ω)] (p − 1)θ [µ (Ω)] p−ε < µ (Ω) p−ε [µ (Ω)] 0<ε≤p−1 0<ε≤p−1 sup ε = θ 1 1 1 1 1 sup η 0<η≤q−1 θ q−η (cid:13)(cid:13)(cid:13) = (p − 1)θ (q − 1)θ [µ (Ω)] = (p − 1)θ (q − 1)θ [µ (Ω)] 1 p [µ (Ω)] 1 p + 1 q < ∞. 1 p(cid:13)(cid:13)(cid:13)q−η 1 q If µ (Ω) < 1, then µ (Ω) increasing, by (1) we have (cid:13)(cid:13)χEχQ+x(cid:13)(cid:13)W (Lp),θ,Lq),θ) If µ (Ω) > 1, then µ (Ω) 1 p−ε > µ (Ω) 1 p for all 0 < ε ≤ p − 1. Again from (1) , we have (cid:13)(cid:13)χEχQ+x(cid:13)(cid:13)W (Lp),θ,Lq),θ) ≤ (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13) sup ε θ p−ε [µ (Ω)] 0<ε≤p−1 (p − 1)θ [µ (Ω)](cid:13)(cid:13)(cid:13)q),θ ≤ (p − 1)θ (q − 1)θ [µ (Ω)]2 < ∞. 1 p−ε(cid:13)(cid:13)(cid:13)(cid:13)q),θ = sup 0<η≤q−1 η θ q−η (cid:13)(cid:13)(cid:13) (p − 1)θ µ (Ω)(cid:13)(cid:13)(cid:13)q−η Proof of property 8. Fix 0 < ε ≤ p − 1. By the Holder's inequality for Wiener's amalgam space we have ZE f dx =ZΩ f χE dx ≤ kf kW (Lp−ε,Lq−ε) kχEkW (L(p−ε)′ ,L(q−ε)′ ) = C kf kW (Lp−ε,Lq−ε) 1 p−ε + 1 where Thus we obtain (p−ε) ′ = 1 q + 1 (q−ε) ′ = 1, and C = C (p, θ, E) = kχEkW (L(p−ε)′ ,L(q−ε)′ ) . ZE f dx ≤ C sup ε 0<ε≤p−1 θ p−ε kf kW (Lp−ε,Lq−ε) = C kf kW (Lp),Lq)) . (cid:3) Theorem 3. The grand Wiener amalgam space W (cid:0)Lp),θ, Lq),θ(cid:1) is a Banach space. ON THE GRAND WIENER AMALGAM SPACE 5 Proof. To show that W (cid:0)Lp),θ, Lq),θ(cid:1) is a Banach space it is suffices to show that if (fn)n∈N is a sequence in W (cid:0)Lp),θ, Lq),θ(cid:1) with kfnkW (Lp),θ ,Lq),θ) < ∞, Xn∈N then Xn∈N fn converges to an element of W (cid:0)Lp),θ, Lq),θ(cid:1) . The proof of this is mu- tadis and mutandis same as in the proof of completness of Wiener amalgam space W (Lp, Lq) , (see Proposition 11.3.2, in [15] , and [9]). (cid:3) 3. Inclusions and consequences Proposition 1. Let 1 < p, q < ∞ . Then for an arbitrary ε and η, 0 < ε ≤ p − 1, 0 < η ≤ q − 1, we have W (Lp, Lq) ⊂ W (cid:16)Lp),θ, Lq),θ(cid:17) ⊂ W (cid:0)Lp−ε, Lq−η(cid:1) . If q ≤ p, then the inclusion W (Lp, Lq) ⊂ W (cid:0)Lp),θ, Lq),θ(cid:1) is strict. Proof. By the definition of generalized grand Lebesgue space we have (2) F p−ε f Then from (2) , (x) =(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε ≤(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ = F p),θ f (x) . kf kW (Lp−ε,Lq−η ) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε(cid:13)(cid:13)(cid:13)q−η ≤(cid:13)(cid:13)(cid:13) q−η (cid:13)(cid:13)(cid:13) (x)(cid:13)(cid:13)(cid:13)q),θ This implies W (cid:0)Lp),θ, Lq),θ(cid:1) ⊂ W (Lp−ε, Lq−η) . ≤ sup F p),θ 0<η≤q−1 ε f f θ F p),θ (x)(cid:13)(cid:13)(cid:13)q−η = kf kW (Lp),θ,Lq,θ)) . Lp ⊂ Lp),θ and Lq ⊂ Lq),θ. Then there exists D1 > 0 and D2 > 0 such that We now want to show that W (Lp, Lq) ⊂ W (cid:0)Lp),θ, Lq),θ(cid:1) . It is known that kgkp),θ ≤ D1 kgkp , (3) khkq),θ ≤ D2 khkq . Then by (2) and (3) F p),θ f (4) Hence by (3) and (4) , (x) =(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ kf kW (Lp),θ,Lq),θ) =(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)q),θ ≤(cid:13)(cid:13)(cid:13) Thus W (Lp, Lq) ⊆ W (cid:0)Lp),θ, Lq),θ(cid:1) . F p),θ f = D1F p f (x) . ≤ D1(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p ≤ D1D2(cid:13)(cid:13)(cid:13) f(cid:13)(cid:13)(cid:13)q),θ D1F p F p f(cid:13)(cid:13)(cid:13)q = kf kW (Lp,Lq) . 6 A.TURAN G URKANLI (0, 1) and f (x) = x− 1 To see that the inclusion W (Lp, Lq) ⊆ W(cid:0)Lp),θ, Lq),θ(cid:1) is strict, take Ω = p , for p > 1. Then t− 1 F p) f (x) = (cid:13)(cid:13)(cid:13) sup 0<ε≤p−1 t− 1 t− 1 θ 0 ε p χQ+x(cid:13)(cid:13)(cid:13)p),θ ≤(cid:13)(cid:13)(cid:13) p−ε (cid:18)Z 1 a→0(cid:16) p p−ε h lim p−ε (cid:16) p ε(cid:17) p−ε ε ε ε 1 θ θ sup 0<ε≤p−1 sup 0<ε≤p−1 p(cid:13)(cid:13)(cid:13)p),θ p (p−ε)dt(cid:19) t=a(cid:17)i p t=1 t ε = sup 0<ε≤p−1 1 p−ε 1 p−ε = = = = sup ε 0<ε≤p−1 θ p−ε (cid:18)(cid:13)(cid:13)(cid:13) = sup ε 0<ε≤p−1 = sup ε 0<ε≤p−1 θ−1 p−ε .p ε 1 p−ε . θ t− 1 p(cid:13)(cid:13)(cid:13)p−ε(cid:19) p−ε lim a→0(cid:18)Z 1 a→0(cid:16) p p−ε h lim a θ t− 1 p (p−ε)dt(cid:19) p(cid:17)(cid:17)i 1 ε p−ε ε (cid:16)1 − a 1 p−ε! Since θ ≥ 1, (5) Thus from (5) , F p) f (x) = sup 0<ε≤p−1 θ−1 p−ε .p ε 1 p−ε ≤ (p − 1)θ−1 p < pθ. kf kW (Lp),Lq)) = (cid:13)(cid:13)(cid:13) = F p) f (cid:13)(cid:13)(cid:13)q),θ sup η 0<η≤q−1 <(cid:13)(cid:13)pθ(cid:13)(cid:13)q),θ q−η (cid:26)Z 1 θ 0 (cid:0)pθ(cid:1)q−η = (q − 1)θ pθ < ∞. 1 q−η dx(cid:27) = sup 0<η≤q−1 θ q−η pθ ε Then f (x) = x− 1 f (t) = t− 1 t− 1 t /∈ W (Lp, Lq) (0, 1) . So we have f (t) = t− 1 p /∈ Lp (0, 1) . Since W (Lp, Lq) (0, 1) ⊂ Lp (0, 1) for q ≤ p, then f (t) = p ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (0, 1) . In the other hand it is easy to show that p ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) \W (Lp, Lq) . Proposition 2. If 1 < p2 ≤ p1 < ∞ and 1 < q < ∞, then W (cid:0)Lp1),θ, Lq),θ(cid:1) ⊂ Proof. Let f ∈ W (cid:0)Lp1),θ, Lq),θ(cid:1) . Then f χQ+x ∈ Lp1),θ. Since p2 ≤ p1, by Thus if q ≤ p, the inclusion W (Lp, Lq) ⊂ W(cid:0)Lp),θ, Lq),θ(cid:1) is strict. W(cid:0)Lp2),θ, Lq),θ(cid:1) . Theorem 3, in [15] we have Lp1),θ ⊂ Lp2),θ, and (cid:3) for some C > 0. Thus by the solidness of Lq),θ, (cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p2),θ ≤ C(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)p1),θ , x ∈ Ω kf kW (Lp2),θ ,Lq),θ) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p2 ),θ(cid:13)(cid:13)(cid:13)kf kq),θ ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p1),θ(cid:13)(cid:13)(cid:13)q),θ and we have W (cid:0)Lp1),θ, Lq),θ(cid:1) ⊂ W (cid:0)Lp2),θ, Lq),θ(cid:1) . W(cid:0)Lp),θ, Lq2),θ)(cid:1) . Proposition 3. If 1 < q2 ≤ q1 < ∞ and 1 < p < ∞, then W (cid:0)Lp),θ, Lq1),θ(cid:1) ⊂ = kf kW (Lp1),θ ,Lq),θ) , (cid:3) ON THE GRAND WIENER AMALGAM SPACE 7 Proof. Let f ∈ W (cid:0)Lp),θ, Lq1),θ(cid:1) . Then F p) ∞,by Theorem 3 in [14] , there exists C > 0 such that f ∈ Lq1),θ. Since 1 < q2 ≤ q1 < (6) for all θ > 0. Then F p),θ the inclusion f F p),θ f (cid:13)(cid:13)(cid:13) f < ∞ F p),θ ≤ C(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)W (Lp),θ,Lq2 ),θ) (cid:13)(cid:13)(cid:13)W (Lp),θ ,Lq1),θ) ∈ Lq2),θ. This implies f ∈ W (cid:0)Lp),θ, Lq2),θ(cid:1) . Thus we have W (cid:16)Lp),θ, Lq1),θ(cid:17) ⊂ W (cid:16)Lp),θ, Lq2),θ)(cid:17) . (cid:3) By using the Proposition 2 and proposition 3, we easily prove the following corollary. Corollary 1. If 1 < p2 ≤ p1 < ∞ and 1 < q2 ≤ q1 < ∞, then W (cid:0)Lp1),θ, Lq1),θ(cid:1) ⊂ W(cid:0)Lp2),θ, Lq2),θ(cid:1) . Proposition 4. Let 1 < pi, qi < ∞, (i = 1, 2, 3) . If there exist constants C1 > 0, C1 > 0, such that for all u ∈ Lp1),θ, v ∈ Lp2), (7) kuvkLp3),θ ≤ C1 kukLp1),θ kvkLp2),θ and for all u ∈ Lq1),θ, v ∈ Lq2),θ (8) kuvkLq3 ),θ ≤ C2 kukLq1),θ kvkLq2),θ , then there exists a constant C > 0, such that for all f ∈ W (cid:0)Lp1),θ, Lq1),θ(cid:1) and g ∈ W(cid:0)Lp2),θ, Lq2),θ(cid:1) , we have f g ∈ W(cid:0)Lp3),θ, Lq3),θ(cid:1) and kf.gkW (Lp3),θ ,Lq3),θ) ≤ C1C2 kf kW (Lp1),θ ,Lq1),θ) kgkW (Lp2),θ ,Lq2),θ) . χQ+x, from (7) and (8) , Proof. Let f ∈ W (cid:0)Lp1),θ, Lq1),θ(cid:1) and g ∈ W (cid:0)Lp2),θ, Lq2),θ(cid:1) . Since χ2 kf gkW (Lp3),θ ,Lq3),θ) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(f g) χQ+x(cid:13)(cid:13)Lp3),θ(cid:13)(cid:13)(cid:13)Lq3),θ Q+x = = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)f χQ+x(cid:1)(cid:0)f χQ+x(cid:1)(cid:13)(cid:13)Lp3),θ(cid:13)(cid:13)(cid:13)Lq3),θ ≤ C1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)Lp1),θ(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)Lp2),θ(cid:13)(cid:13)(cid:13)Lq3),θ ≤ C1C2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)Lp1),θ(cid:13)(cid:13)(cid:13)Lq1),θ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f χQ+x(cid:13)(cid:13)Lp2),θ(cid:13)(cid:13)(cid:13)Lq2),θ = C1C2 kf kW (Lp1),θ ,Lq1),θ) kgkW (Lp2),θ ,Lq2),θ) . Proposition 5. Let 1 < p ≤ ∞. Then W(cid:0)Lp),θ, Lp),θ(cid:1) (Ω) = Lp),θ (Ω) . Proof. According to the definition of the supremum, for an arbitrary η > 0, there exists 0 < ε0 ≤ p − 1, such that for all f ∈ Lp),θ (Ω) , (9) sup 0<ε≤p−1(cid:16)ε θ p−ε kf kp−ε(cid:17) ≤ ε θ p−ε0 kf kp−ε0 + η. (cid:3) 8 A.TURAN G URKANLI Then we have F p),θ f and so (11) F p),θ f (x) = sup Thus from (10) and (11) we have kf kW (Lp),θ,Lp),θ) = (cid:13)(cid:13)(cid:13) (x) = (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ θ ≤ (cid:18)ε p−ε0 0 = sup 0<ε≤p−1(cid:16)ε θ p−ε (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε(cid:17) p−ε0 F p−ε0 θ + η, f θ θ θ ε p−ε0 0 0<ε≤p−1 F p),θ = sup + η(cid:19) = ε (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0 p−ε kFf kp−ε(cid:17) ≤ ε p−ε (cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)p−ε + η(cid:13)(cid:13)(cid:13)(cid:13)p−ε0 (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0 (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0(cid:13)(cid:13)(cid:13)(cid:13)p−ε0 kf kp−ε0(cid:19) + η.µ (Ω) + η. + ε f θ θ θ f ε F p),θ 0<ε≤p−1(cid:16)ε (cid:13)(cid:13)(cid:13)p),θ (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) (cid:18)ε p−ε0 0 ε ε θ θ θ θ θ p−ε0 ≤ ε 0 p−ε0 0 p−ε0 ≤ ε 0 p−ε0 0 p−ε0 = ε 0 p−ε0 0 p−ε0 = ε 0 kf kW (Lp−ε0 ,Lp−ε0 ) + kηkp−ε0 + η (10) + η (cid:13)(cid:13)(cid:13)p−ε0 kFf kp−ε0 + η. θ ≤ ε p−ε0 0 + η F p),θ f (cid:13)(cid:13)(cid:13) θ p−ε0 kηkp−ε0 + η If η → 0, the right side of (11) approaches to C1 kf kW (Lp),θ,Lp),θ) and we have (12) kf kW (Lp),θ,Lp),θ) ≤ C1 kf kp),θ for some constant C1 > 0. Conversely let f ∈ Lp),θ. For this η > 0,we obtain kf kp),θ = sup ε 0<ε≤p−1 θ θ p−ε0 p−ε kf kp−ε ≤ ε 0 kf kp−ε0 + η θ = ε p−ε0 0 kf kW (Lp−ε0 ,Lp−ε0 ) + η ≤ θ = ε p−ε0 0 ≤ (cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13)(cid:13) = ε ε p−ε0 0 θ θ ε p−ε0 0 −θ p−ε0 0 + η + η (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0(cid:13)(cid:13)(cid:13)p−ε0 + η(cid:13)(cid:13)(cid:13)(cid:13)p−ε0 (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0 (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0(cid:13)(cid:13)(cid:13)(cid:13)p−ε0 (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p−ε0(cid:13)(cid:13)(cid:13)(cid:13)p−ε0! + ε ε kηkp−ε0 (cid:13)(cid:13)(cid:13)(cid:13) p−ε0 0 p−ε0 0 p−ε0 0 + ε ε θ θ θ + η θ p−ε0 0 kηkp−ε0 + η If η → 0, then the right side of (12) approaches to C2 kf kW (Lp),θ ,Lp),θ) and we have (13) kf kp),θ ≤ C2 kf kW (Lp),θ ,Lp),θ) for some constant C2 > 0. Combining (12) and (13) , we obtain W (cid:0)Lp),θ, Lp),θ(cid:1) (Ω) = Lp),θ (Ω) . (cid:3) ON THE GRAND WIENER AMALGAM SPACE 9 4. Holder's inequality, Duality and reflexivity in grand Wiener amalgam spaces It is known by Theorem 2, that k.kW (Lp),θ ,Lq),θ) is a Banach function norm and W(cid:0)Lp),θ, Lq),θ(cid:1) (Ω) is a Banach function space. Definition 2. The associate space of W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) determined by the W (Lp),θ ,Lq),θ) associate norm k.k ′ is where ′ W (cid:16)Lp),θ, Lq),θ(cid:17) ′ = sup  W (Lp),θ ,Lq),θ) ′ < ∞(cid:27) , (Ω) =(cid:26)g ∈ M0 : kgk f g dµ : f ∈ M0 (Ω) , kf kW (Lp),θ,Lq),θ) ≤ 1  ZΩ . Theorem 4. (Holder's inequality) If f ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) and g ∈ (Ω) , then f g is integrable and ′ kgk W (Lp),θ,Lq),θ) W(cid:0)Lp),θ, Lq),θ(cid:1) ZΩ (14) f (x) .g (x) dx ≤ kf kW (Lp),Lq)) . kgk W (Lp),θ ,Lq),θ) ′ . Proof. Since W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) is a Banach function space by Theorem 2, the proof is completed by Theorem 2.4., in [3] . (cid:3) Proposition 6. The closure C∞ 0 W (Lp),θ,Lq),θ)of the set C∞ 0 in the space W(cid:0)Lp),θ, Lq),θ(cid:1) (Ω) consists of functions f ∈ W(cid:0)Lp),θ, Lq),θ(cid:1) (Ω) such that θ p−ε kf kW (Lp−ε,Lq−ε) = 0, (15) lim ε→0 ε where C∞ 0 compact support. is the space of infinitely differentiable complex valued functions with Proof. First we will show that (15) holds for C∞ 0 (Ω) 0 (Ω) is dense in W (Lp, Lq) by Theorem 1, in [9] . Let 0 W (Lp),θ ,Lq),θ) . Then there exists a sequence (fn) ⊂ W (Lp, Lq) such that is dense Lp (Ω) , then C∞ f ∈ C∞ 0 W (Lp),θ ,Lq),θ) . Since C∞ kfn − f kW (Lp,Lq) → 0. Thus for given η > 0, there exists n0 ∈ N such that η 2 kfn0 − f kW (Lp,Lq) < (16) . Since ε θ p−ε kfn0 kW (Lp−ε,Lq−ε) = ε and fn0χQ+x ∈ Lp, by the Holder's inequality (17) ε θ p−ε (cid:13)(cid:13)fn0 χQ+x(cid:13)(cid:13)p−ε = ε θ p−ε   , 1 p−ε θ p−ε (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)fn0 χQ+x(cid:13)(cid:13)p−ε(cid:13)(cid:13)(cid:13)q−ε dt ZΩ (cid:12)(cid:12)fn0 (t) χQ+x(cid:12)(cid:12)  p−ε 10 A.TURAN G URKANLI ≤ ε = ε Thus from (17) , ε (cid:13)(cid:13)(cid:13) θ p−ε (cid:13)(cid:13)fn0χQ+x(cid:13)(cid:13)p−ε(cid:13)(cid:13)(cid:13)q−ε θ p−ε (cid:13)(cid:13)fn0 χQ+x(cid:13)(cid:13)p p−ε µ (Ω) θ ε 1 p−ε − 1 p µ (Ω) p(p−ε) (cid:13)(cid:13)fn0 χQ+x(cid:13)(cid:13)p . ≤ (cid:13)(cid:13)(cid:13) = ε θ ε p−ε µ (Ω) θ p−ε µ (Ω) ≤ ε θ p−ε µ (Ω) = ε θ p−ε µ (Ω) ε ε p(p−ε) (cid:13)(cid:13)fn0χQ+x(cid:13)(cid:13)p(cid:13)(cid:13)(cid:13)q−ε p(p−ε) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)fn0χQ+x(cid:13)(cid:13)p(cid:13)(cid:13)(cid:13)q−ε p(p−ε) µ (Ω) q−ε − 1 1 ε q (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)fn0χQ+x(cid:13)(cid:13)p(cid:13)(cid:13)(cid:13)q q(q−ε) kfn0kW (Lp,Lq) → 0, ε p(p−ε) + ε as ε → 0. Hence there exists ε0 such that when ε < ε0, (18) ε p−ε kfn0k θ W (Lp−ε,Lq−ε) Then by (16) and (18) , ε θ p−ε kf kW (Lp−ε,Lq−ε) ≤ ε θ p−ε kfn0 − f k ε =(cid:13)(cid:13)(cid:13) θ p−ε (cid:13)(cid:13)fn0 χQ+x(cid:13)(cid:13)p−ε(cid:13)(cid:13)(cid:13)q−ε ≤ η 2 . ≤ kfn0 − f kW (Lp),θ ,Lq),θ) + < W (Lp−ε ,Lq−ε) η 2 + ε θ p−ε kfn0k η 2 η 2 + = η W (Lp−ε,Lq−ε) when ε < ε0. This completes the proof. (cid:3) Proposition 7. If q ≤ p, and θ ≥ 1, then the set C∞ 0 (Ω) is not dense in W(cid:0)Lp),θ, Lq),θ(cid:1) (Ω) . Proof. It is enough to give the proof for 1 p , for p > 1. We showed in Proposition 2, that t− 1 f (t) = t In the other hand since C∞ 0 then t− 1 This ends the proof. p /∈ C∞ 0 W (Lp),θ ,Lq),θ) . Thus t− 1 W (Lp),θ,Lq),θ)⊂ C∞ 0 Lp),θ and t− 1 p ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (0, 1) . p /∈ C∞ 0 (0, 1) Lp),θ , W (Lp),θ,Lq),θ) . (cid:3) p ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) r C∞ 0 the special case Ω = (0, 1) . Let Definition 3. Let (X, k.kX ) be a Banach function space and let f ∈ X be any arbitrary function. We say that f has absolutely continuous norm in X if for every sequence {En}n∈N satisfying En → φ.We will denote the subspace of the functions in X with absolutely continuous norm by Xa. If X = Xa, then X it self is said to have absolute continuous norm. lim n→∞(cid:13)(cid:13)f χEn(cid:13)(cid:13)X = 0 We need the following known two theorems to find the dual of the grand Wiener amalgam spaces.. Theorem 5. ( See [3] , Corollary 4.3). The dual space X ∗of a Banach function if and only if X has absolutely ′ space X is isometric to the associate space X continuous norm. Theorem 6. ( See [3] , Corollary 4.4). A Banach function space is reflexive if and only if both X and its associate space X ′ have absolute continuous norm. ON THE GRAND WIENER AMALGAM SPACE 11 Theorem 7. The grand Wiener amalgam spaces W (cid:0)Lp),θ, Lq),θ(cid:1) (c, d) has not absolute continuous norm. Proof. We will give the proof of this theorem for the interval (c, d) = (0, 1) . One can prove similarly this theorem for any interval (c, d) . Take the function f (t) = t− 1 p . We will show that f is not absolute continuous. Let Q = (0, a) ⊂ (0, 1) . Then (19) = = = (cid:13)(cid:13)(cid:13)(cid:16)t− 1 sup 0<ε≤p−1 sup 0<ε≤p−1 sup 0<ε≤p−1 θ ε t− 1 p χQ(cid:17) χQ+x(cid:13)(cid:13)(cid:13)p),θ =(cid:13)(cid:13)(cid:13) p−ε (cid:13)(cid:13)(cid:13) p χ(x,a)(cid:13)(cid:13)(cid:13)p−ε p−ε (cid:18)Z a p (cid:19) p−ε (cid:16) p ε ha pi(cid:17) t− p−ε p − x ε ε x 1 θ θ ε ε p−ε t− 1 p χQ∩Q+x(cid:13)(cid:13)(cid:13)p),θ = sup ε 0<ε≤p−1 = sup 0<ε≤p−1 1 p−ε = sup 0<ε≤p−1 θ t− 1 t− 1 p χ(x,a)(cid:13)(cid:13)(cid:13)p),θ =(cid:13)(cid:13)(cid:13) p−ε (cid:18)Z a p−ε(cid:19) p(cid:12)(cid:12)(cid:12) x (cid:12)(cid:12)(cid:12) p−ε (cid:18)Z a p(cid:19) p−ε ha t−1+ ε θ−1 p−ε p p − x ε x 1 θ ε ε 1 p−ε 1 p−ε 1 p−ε . ε pi θ−1 p−ε and p 1 p−ε are increasing functions of ε. Since 0 < a < 1, It is easy to see that ε from (19) we have (20) (cid:13)(cid:13)(cid:13)(cid:16)t− 1 p χQ(cid:17) χQ+x(cid:13)(cid:13)(cid:13)p),θ p χ(0,a)(cid:13)(cid:13)(cid:13)W (Lp),θ,Lq),θ) t− 1 lim a→0(cid:13)(cid:13)(cid:13) Also since x tends to 0 as a → 0, from (20) we obtain p−1 t− 1 = lim p i . = (p − 1)θ−1 ph1 − x p χ(x,a)(cid:13)(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)(cid:13)q),θ a→0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) a→0(cid:13)(cid:13)(cid:13) (p − 1)θ−1 ph1 − x = (cid:13)(cid:13)(cid:13) (p − 1)θ−1 ph1 − x = (p − 1)θ−1 p(cid:13)(cid:13)(cid:13) x→0(cid:16)1 − x = (p − 1)θ−1 p k1kq),θ 6= 0. = lim lim a→0 lim p−1 p−1 p i(cid:13)(cid:13)(cid:13)q),θ p i(cid:13)(cid:13)(cid:13)q),θ p (cid:17)(cid:13)(cid:13)(cid:13)q),θ p−1 Thus f (t) = t− 1 lute continuous norm. p ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (0, 1) , but it has not absolute continuous norm in W (cid:0)Lp),θ, Lq),θ(cid:1) (0, 1) . Then by definition 3, W(cid:0)Lp),θ, Lq),θ(cid:1) (0, 1) has not abso- Corollary 2. The grand Wiener amalgam spaces W (cid:0)Lp),θ, Lq),θ(cid:1) (c, d) is not reflexive. (cid:3) Proof. The proof is clear from Theorem 9 and Theorem 10. (cid:3) Corollary 3. The dual space (cid:0)W (cid:0)Lp),θ, Lq),θ(cid:1) (c, d)(cid:1)∗ amalgam space W (cid:0)Lp),θ, Lq),θ(cid:1) (c, d) is not isometric to the associate space(cid:0)W (cid:0)Lp),θ, Lq),θ(cid:1) (c, d)(cid:1) this grand Wiener amalgam space W(cid:0)Lp),θ, Lq),θ(cid:1) (c, d). Proof. The proof of this theorem is easy from Theorem 8 and Theorem 10. (cid:3) of the grand Wiener ′ of 12 A.TURAN G URKANLI Theorem 8. Let ′ p p−1 = p , ′ q q−1 = q and 1 < p, q < ∞. Then and the norms kgk kgk W (Lp),θ ,Lq),θ) are equivalent. ′ ,Lq) W(cid:16)Lp) (cid:16)W (cid:16)Lp),θ, Lq),θ(cid:17) (Ω)(cid:17) ′ =  ′ (cid:17) and supZΩ ′ = W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) , f (x) g (x) dx : f ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) , kf kW (Lp),Lq)) ≤ 1   Proof. If f ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) and g ∈ W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) , then kf kW (Lp),θ,Lq),θ) . kgkW (Lp)′,θ ,Lq)′ ,θ) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)p)′ ,θ(cid:13)(cid:13)(cid:13)q)′ ,θ ∈ Lq),θ and (cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)Lp)′,θ ∈ Lq)′,θ, by the Holder inequality Since (cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ for the generalized grand Lebesgue space(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)Lp)′,θ ∈ L1 (Ω) and (21) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)p)′ ,θ(cid:13)(cid:13)(cid:13)1 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)(cid:13)q),θ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)p)′ ,θ(cid:13)(cid:13)(cid:13)q)′ ,θ Also since f.χQ+x ∈ Lp),θ and g.χQ+x ∈ Lp)′,θ, one more applying the Holder inequality for the generalized Lebesgue space we have . . (22) kf gk1 = kf gkW (L1,L1) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)f.χQ+x(cid:1)(cid:0)g.χQ+x(cid:1)(cid:13)(cid:13)1(cid:13)(cid:13)(cid:13)1 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)f.χQ+x(cid:13)(cid:13)p),θ(cid:13)(cid:13)g.χQ+x(cid:13)(cid:13)Lp)′,θ(cid:13)(cid:13)(cid:13)1 . Combining (21) and (22) we obtain ZΩ f (x) .g (x) dx = kf.gk1 ≤ kf kW (Lp),Lq)) . kgk ′ W(cid:16)Lp) ,Lq) ′ (cid:17) . From this inequality we have ≤ kgk ′ W(cid:16)Lp) ,Lq) ′ (cid:17) .   kgk W (Lp),θ,Lq),θ) This implies Thus we have supZΩ f (x) g (x) dx : f ∈ W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω) , kf kW (Lp),Lq)) ≤ 1 ′ =  W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) ⊂(cid:16)W(cid:16)Lp),θ, Lq),θ(cid:17) (Ω)(cid:17) W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) ⊃(cid:16)W (cid:16)Lp),θ, Lq),θ(cid:17) (Ω)(cid:17) W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) =(cid:16)W (cid:16)Lp),θ, Lq),θ(cid:17) (Ω)(cid:17) . . ′ ′ ′ If one uses the same technic in ([13] , Theorem 11.7.1 (c)) , obtains ON THE GRAND WIENER AMALGAM SPACE 13 Since the spaces W (cid:16)Lp)′,θ, Lq)′,θ(cid:17) (Ω) and (cid:0)W (cid:0)Lp),θ, Lq),θ(cid:1) (Ω)(cid:1) ′ (cid:17) and they are Banach spaces with respect to the norms k.k ,Lq) ′ W(cid:16)Lp) ′ are equal and kgk W (Lp),θ,Lq),θ) supZΩ ′ =  f (x) g (x) dx : f ∈ W (cid:16)Lp),θ, Lq),θ(cid:17) (Ω) , kf kW (Lp),Lq)) ≤ 1  , respectively, then these norms are equivalent by Two norm theorem (see Theorem 7.3.3, [19]) (cid:3) References [1] Anatriello G, Chil R and Fiorenza A. Identification of fully Measurable Grand Lebesgue Spaces. Journal of Function Spaces, Vol. 2017. [2] Anatriello G. Iterated grand and small Lebesgue spaces. Collect. Math. 2014; 64 : 273 − 284. [3] Bennett C and Sharpley, R. Interpolation of Operators, Academic Press, INC, Orlando, Florida,1988. [4] Capone C, Formica MR, Giova R. Grand Lebesgue spaces with respect to measurable func- tions. Nonlinear Analysis 2013; 85 : 125 − 131. [5] Capone C, and Fiorenza A. On small Lebesgue spaces. Journal of function spaces and appli- cations 2005; 3 (1) : 73 − 89. [6] Castillo R. E, Raferio H. Inequalities with conjugate exponents in grand Lebesgue spaces. Hacettepe Journal of Mathematics and Statistics 2015; 44 (1) : 33 − 39. [7] Castillo R. E, Raferio H. An Introductory Course in Lebesgue Spaces. Springer International Publishing Switzerland 2016. [8] Danelia N, Kokilashvili V. On the approximation of periodic functions within the frame of grand Lebesgue spaces. Bulletin of the Georgian national academy of sciences 2012; 6(2) : 11 − 16. [9] Feichtinger HG, Banach convolution algebras of Wiener's type, Proc. Conf. " Functions, Series, Operators", Budapest, 1980, Colloquia Math. Soc. J. Bolyai, North Holland Publ. Co., Amsterdam- Oxford- New York 1983 : 509 − 524. [10] Feichtinger HG and Grochenig KH, Banach spaces related to integrable group representations and their atomic decompositions I, J. Funct. Anal., 86(1989), 307 -- 340. [11] Fiorenza A, and Karadzhov GE. Grand and small Lebesgue spaces and their analogs, Journal for Analysis and its Applications 2004; 23 (4) : 657 − 681. [12] Fiorenza A. Duality and reflexity in grand Lebesgue spaces, Collect. Math. 2000; 51 (2) : 131 − 148. [13] Greco L, Iwaniec T, Sbordone C. Inverting the p-harmonic operator, Manuscripta Math.1997; 92 : 259 − 272. [14] Gurkanlı AT, Inclusion and the approximate identities of the generalized grand Lebesgue spaces. Turk J Math Doi:10.3906/mat-1803-89, (Accepted for publication) [15] Heil C, An Introduction to Weighted Wiener Amalgams, In: Wavelets and their Applications (Chennai, 2002), Allied Publishers, New Delhi, 2003 : 183 − 216. [16] Holland F. Square -summable positive-definite functions on real line, Linear operators Ap- prox. II, Ser. Numer. Math. 25, Birkhauser, Basel, 1974, 247 − 257. [17] Holland F. Harmonic analysis on amalgams of Lp and ℓq, London Math. Soc. 10 − 2, (1975) , 295 − 305. [18] Iwaniec T, Sbordone C. On the integrability of the Jacobian under minimal hypotheses, Arc. Rational Mech. Anal. 1992; 119 : 129 − 143.. [19] Larsen R, Functional Analysis an Introduction, Marcel Dekker, INC. New York, 1973. Istanbul Arel University Faculty of Science and Letters Department of Mathe- matics and Computer Sciences E-mail address: [email protected]
1111.2643
1
1111
2011-11-11T01:40:42
The asymptotic expansion of the heat kernel on a compact Lie group
[ "math.FA", "math.DG", "math.MG", "math.SP" ]
Let $G$ be a compact connected Lie group equipped with a bi-invariant metric. We calculate the asymptotic expansion of the heat kernel of the laplacian on $G$ and the heat trace using Lie algebra methods. The Duflo isomorphism plays a key role.
math.FA
math
THE ASYMPTOTIC EXPANSION OF THE HEAT KERNEL ON A COMPACT LIE GROUP SEUNGHUN HONG ABSTRACT. Let G be a compact connected Lie group equipped with a bi-invariant metric. We calculate the asymptotic expansion of the heat kernel of the laplacian on G and the heat trace using Lie algebra methods. The Duflo isomorphism plays a key role. CONTENTS 1. Introduction 2. Preliminaries 3. The Asymptotic Expansion of the Heat Kernel of the Laplacian on a Compact Lie Group References 1 2 5 7 1. INTRODUCTION (1.1) The laplacian 4 on a compact riemannian manifold M depends only on the metric on M. Conversely, one can deduce the metric from the laplacian. Hence, it is reasonable to expect the spectrum Sp(4) of the laplacian to be constrained by the geometry and vice versa. The first result in this vein is Weyl's law [11, 12] which states that the number N(λ) of the eigenvalues of the laplacian less than λ satisfies the asymptotic equality vol(Ω) (4π)n/2Γ ( n as λ→∞, where Ω is a bounded open subset of R2 or R3 on which the laplacian is defined. Gårding N(λ) λn/2 + O(1/λ) 2 + 1) = [5] proved the higher-dimensional case (for generic elliptic operators). On closed riemannian manifolds, the same law was proved for the laplacian by Duistermaat and Guillemin [4], and for generic elliptic operators by Minakshisundaram and Pleijel [9]. Weyl's law can be reformulated as an asymptotic behavior of the function ∞(cid:88) k=1 Z(t) = e−tλk e−tλ1 e−tλ2 et4 = e−tλ3 . ... where −λk denotes the eigenvalue of the kth eigenfunction of the laplacian. This funcion resembles the "partition function" in physics -- a function that is often invariant under the symmetry of the physical system it describes. It is the trace of the heat diffusion operator Ñ é 2010 Mathematics Subject Classification. Primary 58J05, 58J35, 58J37, 58J50, 58J60; Secondary 35K08. Key words and phrases. compact Lie groups, Laplace-Beltrami operator, heat kernel expansion, heat trace expansion, Duflo isomorphism. The author wishes to express his heartfelt gratitude to his advisor, Professor N. Higson, for the kind guidance and advice. 1 The relation between Z(t) and the number of eigenfunctions becomes evident as we consider the 2 limit t→ 0+; in that limit, the partial sum SEUNGHUN HONG K(cid:88) e−tλk converges to K, the number of the eigenvalues from λ1 to λK. Weyl's law can be shown to be equivalent to the asymptotic law k=1 (1.2) tn/2Z(t) ∼ vol(Ω) (4π)n/2 + O(t) expansion as t→ 0+. Minakshisundaram and Pleijel [9] showed that the partition function has an asymptotic as t→ 0+. McKean and Singer [8] calculated a0, a1, a2 and, in particular, showed that a0 is the (a0 + a1t + a2t2 + ··· ) (cid:17)dim M/2 (cid:16) 1 (cid:82) 4πt Z(t) ∼ riemannian volume of M and a1 is 1 6 coefficients are extremely hard to calculate in general. M S where S is the scalar curvature of M. The higher order Our goal is to consider the asymptotic expansion of Z(t) in the case where M is a compact connected Lie group G, equipped with a metric that is invariant under the left and right-translations. We follow the heat kernel method, that is, we calculate the asymptotic expansion of the heat kernel kt(x, y) of the laplacian. The asymptotic expansion for Z(t) can then be obtained from the relation (cid:90) Z(t) = kt(x, x) vol(x). G Here vol(x) is the riemannian volume form. Our motivation comes from the expectation that the high degree of symmetry will substantially simplify the calculations. Our strategy is to utilize the tight connection between G and its Lie algebra g. The key ingredient is the Duflo isomorphism Duf : S(g)g→ Z(g). The space S(g)g can be identified with the constant coefficient differential operators on g, and Z(g) can be identified with the bi-invariant differential operators on G. Owing to the bi-invariance of the metric on G, the laplacian 4G on G is a bi-invariant differential operator; and it corresponds to the Casimir element in Z(g). The operator on g corresponding to the preimage of the Casimir under the Duflo isomorphism is (not surprisingly) the laplacian 4g on the euclidean space g with an extra constant term. The heat kernel of the flat laplacian is simply the gaussian kernel. Based on these relations, we can deduce, without much effort, the asymptotic expansion of the heat kernel of the laplacian on G. algebra, namely, the tangent space TeG at the identity e ∈ G. We denote by (cid:101)X the left-invariant (1.3) Notations Throughout this article G denotes a compact connected Lie group, and g its Lie vector field on G generated by X ∈ g. We denote by h , i the (selected) bi-invariant metric on G. Such a metric is equivalent to an Ad(G)-invariant inner product on g. (2.1) We review some basic analytic and algebraic notions related to the laplacian on a compact Lie group. 2. PRELIMINARIES Analytic Aspects. (2.2) Proofs for most of the statements made in this subsection can be found in [2, Ch.2]. ASYMPTOTIC EXPANSION OF THE HEAT KERNEL ON A COMPACT LIE GROUP (2.3) The Laplacian Let C the space of smooth vector fields on G. Define the gradient operator grad : C (G) denote the space of smooth functions on G, and let X(G) denote divergence operator div : X(G)→ C ∞ (G) by 3 ∞ (G)→ X(G) and the ∞ hgrad f, Xiκ = Xf, (div X) vol = LX vol, 4Gf = div(grad f). ∞ for X ∈ X(G), where vol is the riemannian volume form and LX is the Lie derivative with respect to X. Then the laplacian (or the Laplace-Beltrami operator) 4G : C (G) is defined by ∞ (G)→ C The definition of 4G is independent of local coordinates and depends only on the metric. Because our metric h , i is bi-invariant, so is the laplacian. An expression for 4G in local coordinates (x1, . . . , xn) can be given as follows. Let g be the matrix defined by gij = h∂i, ∂ji where ∂i = ∂/∂xi. Let gij denote the (i, j)-entry of g−1. Then, (cid:88) i,j 1(cid:112)det g (cid:112)det ggij∂jf). ∂i( (2.4) 4Gf = (cid:80) ∞ (2.5) The Spectrum of the Laplacian So far the laplacian is an unbounded operator whose domain (cid:80) (G) is a dense subspace of L2(G). The domain can be extended to the Sobolev space H2(G), that C is, the space of measurable functions u on G such that the norm kukH2 = kukL2 + i k∂iukL2 + i,j k∂i∂jukL2 is finite. The Sobolev embedding theorem tells us that H2(G) is a subspace of L2(G) and that the inclusion map is compact. The extension 4G : H2(G)→ L2(G) is the unique self-adjoint extension of 4G which was originally defined on C of the theory of unbounded operators, the laplacian is essentially self-adjoint on C (G). In the language ∞ It turns out that (1 − 4G), where 1 denotes the identity operator, admits an inverse that is com- i=1 of (1 − 4G)−1 form an pact. Then the spectral theorem implies that the eigenfunctions { uk } ∞. We can also conclude from the spectral theorem that the eigenfunctions uk orthonormal basis for L2(G). Owing to the regularity of elliptic differential operators, the eigen- functions are of C can be ordered in such a way that the corresponding eigenvalues −λk of 4G give a nonincreasing unbounded sequence of negative real numbers, (G). ∞ ∞ 0 > −λ1 (cid:62) −λ2 (cid:62) −λ3 (cid:62) ··· . The heat diffusion operator of 4G is then defined by the matrix Ñ é et4G = e−tλ1 e−tλ2 e−tλ3 ... with respect to the basis consisting of eigenfunctions of the laplacian. (2.6) The Heat Kernel The heat diffusion operator is, in fact, an integral operator on L2(G) with a C (M × M), such that ∞-kernel. That means, there is some Kt ∈ C ∞ (cid:90) (et4G f)(x) = (2.7) for any f ∈ L2(G). The kernel Kt is called the heat kernel of 4G. Owing to the equivariance of 4G, we have Kt(x, y) = Kt(e, x−1y), where e is the identity of G any x, y are arbitrary points in G. So the heat kernel is completely determined by the function Kt(x, y)f(y) vol(y) G We will call this the heat convolution kernel for 4G. With it, equation (2.7) can be rephrased as (2.8) (et4G f)(x) = kt(x−1y)f(y) vol(y). kt(x) := Kt(e, x). (cid:90) G 4 SEUNGHUN HONG The trace of the heat diffusion operator, Z(t) = tr(et4G ), is called the partition function or the heat-trace of 4G. It can be calculated in terms of the heat kernel as follows: (2.9) Kt(x, x) vol(x) = kt(e) vol(G). Z(t) = (cid:90) G (2.10) The convolution kernel kt admits an asymptotic expansion as t→ 0+, where ht is the gaussian kernel and ai are smooth functions on M. The gaussian kernel kt ∼ ht(a0 + a1t + a2t2 + ··· ) ht, under the exponential chart near e ∈ G, takes the form e−kXk2/4t (4πt)dim M/2 ht(X) = . The asymptotic expansion (2.10) means that, for each nonnegative integers r, N, and n, there is a constant C such that (cid:13)(cid:13)(cid:13)kt − ht N(cid:88) i=0 aiti(cid:13)(cid:13)(cid:13)Cr (cid:54) Ctn (cid:80)∞ i=0 aiti is a formal solution to the differential equation for sufficiently small t. Here k · kCr denotes the usual norm on Cr(M). The asymptotic series st := ht (2.11) under the condition st(e) = 1. This gives a family of differential equations that can be solved inductively: (∂t + 4G)st = 0 k(cid:88) (∂t + 4G)ht aiti = httk4Gak. Algebraic Aspects. i=0 (2.12) The Universal Enveloping Algebra The universal enveloping algebra U(g) of g is con- structed by first taking the tensor algebra T (g) of g and then taking the quotient by the ideal I(g) generated by the elements of the form X ⊗ Y − Y ⊗ X − [X, Y]: The adjoint action of X ∈ g on g extends to U(g) as a derivation. The invariant subspace of U(g) under all such inner derivations is the center of the universal enveloping algebra; U(g) = T (g)/I(g). Z(g) = U(g)g. Suppose we have a simple tensor X1 ··· Xn in U(g). It generates the left-invariant differential operator X1 ··· Xn, where Xi denotes the left-invariant vector field on G generated by Xi ∈ g. This gives an algebra isomorphism between U(g) and the space D(G) of left-invariant differential operators. Under this bijection, the center Z(g) of the universal enveloping algebra corresponds to the subalgebra of bi-invariant differential operators on G. (2.13) The Casimir Element We pointed out earlier that the laplacian 4G is a bi-invariant opera- tor. If { Xi }n i=1 (n = dim g) is an orthonormal basis for g, then we claim that (2.14) In other words, the element in U(g) that corresponds to 4G is the Casimir element: i=1 Xi Xi. 4G = n(cid:88) n(cid:88) Cas = XiXi. i=1 Since the differential operators on both sides of (2.14) are left-invaraint, it is enough to check their equality at e ∈ G. To that end, take the exponential coordinate system (x1, . . . , xn) centered at e. In other words, the coordinates (x1, . . . , xn) correspond to the point exp(cid:0)(cid:80) ASYMPTOTIC EXPANSION OF THE HEAT KERNEL ON A COMPACT LIE GROUP (cid:1) in G. In this 5 n i=1 xiXi coordinate system, we have n(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)e n(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)0 . ∂2f ∂x2 i Xi Xif = i=1 i=1 ∂ ∂xi n i=1 (cid:80) (see [6, Ch.1, §6] for details). This is, in general, different from the Lie-theoretic exponential map exp which has nothing to do with the metric. But the two exponential maps do agree if the metric is bi-invariant, which can be seen as follows. Let ∇ be the riemannian connection so that it satisfies (2.15) We need to show that the right-hand side is equal to 4Gf at the identity; in other words, we need to verify that the expression of 4G at the identity under the exponential chart is . To (cid:12)(cid:12)(cid:12)0 see that this is the case, recall the riemannian exponential map Exp : g→ G arising from the metric h(cid:101)Y,(cid:101)Zi = h∇(cid:101)X(cid:101)Y,(cid:101)Zi + h(cid:101)Y,∇(cid:101)X(cid:101)Zi. Using the identity 2h∇(cid:101)X(cid:101)Y,(cid:101)Zi = (cid:101)Xh(cid:101)Y,(cid:101)Zi +(cid:101)Y h(cid:101)Z,(cid:101)Xi −(cid:101)Zh(cid:101)X,(cid:101)Yi + ∇(cid:101)X h[(cid:101)X,(cid:101)Y],(cid:101)Zi − h[(cid:101)Y,(cid:101)Z],(cid:101)Xi + h[(cid:101)Z,(cid:101)X],(cid:101)Yi and the skew-symmetricity of the ad(g)-action, one can check holds for all X, Y in g. In particular, ∇(cid:101)X(cid:101)X = 0. It follows [6, Ch.2, Prop.1.4] that the geodesic γX(t), X(0) = X, is a group homomorphism R → G. By the uniqueness of 1- such that γX(0) = e and γ0 parameter subgroups, we have γX(t) = exp(tX). This implies that the riemannian exponential map is identical to the Lie-theoretic exponential map. As a consequence, the matrix [gij] of the metric under the exponential chart satisfies gij(e) = δij (Kronecker delta) and ∂kgij(e) = 0. Therefore, by equation (2.4), we have ∇(cid:101)X(cid:101)Y = [(cid:101)X,(cid:101)Y] (2.16) that 1 2 (2.17) Hence 4G agrees with n are equal everywhere on G. 4Gf (cid:80) i=1(cid:101)Xi(cid:101)Xi at the identity, and this proves that these two invariant operators ∂2f ∂x2 i i=1 = . n(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)e (cid:12)(cid:12)(cid:12)(cid:12)0 (2.18) The Duflo Isomorphism Let S(g) be the symmetric algebra of g. The adjoint action of X ∈ g on g extends, as an inner derivation, to S(g). Denote by S(g)g the subalgebra of S(g) that is invariant under all such inner derivations. We identify S(g)g with the constant coefficient differential operators on g. Duflo [3] showed that there is an algebra isomorphism Duf : S(g)g→ Z(g). If we view, for the moment, U(g) as the convolution algebra of distributions on G supported at e ∈ G and S(g) as the convolution algebra of distributions on g supported at 0 ∈ g, then Duf is j · exp∗, that is, the push-forward along the exponential map followed by the multiplication by the function j(X) = det1/2( sinh adX /2 adX /2 ). We note that the push-forward map exp∗ alone gives the Poincaré-Birkhoff-Witt isomorphism S(g)g→ Z(g), which is only a vector space isomorphism. (cid:80) 3. THE ASYMPTOTIC EXPANSION OF THE HEAT KERNEL OF THE LAPLACIAN ON A COMPACT LIE GROUP (3.1) Let 4g be the laplacian of the euclidean space g; if { Xi }n then 4g = coefficient differential operators on g. The image of 4g under the Duflo isomorphism is i=1 is an orthonormal basis for g, i=1 XiXi. This is an element of S(g)g, which we identify as the space of constant n (3.2) sentation ad : g→ End(g) to the universal enveloping algebra U(g). Equation (3.2) can be proved where trg denotes the trace for the linear operators on g obtained by extending the adjoint repre- in more than one way. For a Lie-algebraic proof, we refer to the work of Alekseev and Meinrenken Duf(4g) = Cas + 1 24 trg(Cas), 6 SEUNGHUN HONG [1]. Recall that Cas, under the identification of Z(g) with the space of bi-invariant differential op- erators on G, corresponds to the laplacian 4G on G. And, by a result of Kostant [10, Eq.1.85], we have trg Cas = −hρ, ρi 1 24 (3.3) where h , i is the inner product -- induced from the metric -- on the dual space t∗ of a maximal abelian subalgebra t of g, and ρ ∈ t∗ is the half the sum of the positive roots of G. Hence, we have (3.4) (3.5) Lemma Let Duf(4g)exp be the differential operator defined near a neighborhood of 0 ∈ g by expressing the differential operator Duf(4g) on G under the exponential chart near the identity e ∈ G. We have Duf(4g) = 4G − hρ, ρi . Duf(4g)exp = j−1 ◦ 4g ◦ j, adX /2 ) and its (j−1 ◦ 4g ◦ j)fexp)(cid:12)(cid:12)0. Thus, Duf(4g)exp agrees with j−1◦4g◦ j at 0 ∈ g. This implies that they agree where j and j−1 above indicates the multiplication by the function j(X) = det1/2( sinh adX /2 reciprocal, respectively. Proof. Recall that the Duflo isomorphism is given by j·exp∗ under the identification of U(g) with the convolution algebra of distributions on G supported at e ∈ G and S(g) with the convolution algebra of distributions on g supported at 0 ∈ g. Thus, back in the language of differential operators, Duf(4g)f 0 = on a neighborhood of 0 ∈ g, provided that the operators Duf(4g)exp and j−1 ◦ 4g ◦ j both define invariant differential operators near e ∈ G. The invariance of Duf(4g)exp is clear from equation (3.4). For the invariance of j−1 ◦ 4g ◦ j, we refer to [7, Ch.II, Eq.71, p.273]. (cid:3) := pt ◦ exp has the asymptotic (3.6) Lemma Let pt be the convolution kernel of et Duf(4g). Then pexp expansion 0 where fexp := f ◦ exp. Since j(0) = 1, we have Duf(4g)expfexp e = 4g(jfexp) t t→ 0+, pexp t ∼ htj−1, valid in some neighborhood of 0 ∈ g, where ht is the gaussian kernel on g. (cid:80)∞ Proof. Let st := ht i=0 aiti be the asymptotic expansion for pexp t . It is the formal solution to (3.7) where Duf(4g)exp is the differential operator Duf(4g) expressed in the exponential chart near the identity e ∈ G. By Lemma (3.5), the differential equation (3.7) is equivalent to (∂t + Duf(4g)exp)st = 0, (∂t − j−1 ◦ 4g ◦ j)st = 0. We need to show that ht/j satisfies this differential equation. This is easily done by invoking the fact that ht satisfies the heat equation (∂t − 4g)ht = 0; indeed, (∂t − j−1 ◦ 4g ◦ j)ht/j = j−1∂tht − j−14ght = j−1(∂t − 4g)ht = 0. (cid:3) (3.8) Lemma Let G be a compact connected Lie group equipped with a bi-invariant metric. The scalar curvature S of G is equal to − 1 Proof. We pointed out in (2.16) that the riemannian connection ∇ on G satisfies ∇(cid:101)X(cid:101)Y = 1 routine calculation shows that the Riemann curvature tensor Rm satisfies 2 [(cid:101)X,(cid:101)Y]. A 4 trg(Cas). Rm((cid:101)X,(cid:101)Y,(cid:101)Z,(cid:102)W) = − 1 4 h[[X, Y], Z], Wi = − h[X, Y], [Z, W]i , 1 4 ASYMPTOTIC EXPANSION OF THE HEAT KERNEL ON A COMPACT LIE GROUP 7 where, for the last equality, we have used the fact that ad(g)-action is skew-symmetric. Let { Xi }dim g be an orthonormal basis for g. For the scalar curvature S, we have i=1 dim g(cid:88) dim g(cid:88) Rm((cid:101)Xi,(cid:101)Xj,(cid:101)Xj,(cid:101)Xi) = − dim g(cid:88) i,j=1 1 4 had(Cas)Xi, Xii = − S = = − 1 4 i,j=1 h[[Xi, Xj], Xj], Xii i,j=1 1 4 trg(Cas). (cid:3) (3.9) Theorem Let G be a compact connected Lie group equipped with a bi-invariant metric. Let kt be t = kt ◦ exp has the asymptotic expansion the heat convolution kernel for the laplacian on G. Then kexp kexp t ∼ ht j as t → 0+, valid in a neighborhood of 0 ∈ g, where S is the scalar curvature, j is the function etS/6 defined by the power series j(X) = det1/2( sinh adX /2 ht(X) = e−kXk/4t/(4πt)dim g/2. Proof. Owing to equations (3.3), (3.4), and Lemma (3.8), we have et4G = etS/6et Duf(4g). This implies kt = etS/6pt where pt is the convolution kernel for et Duf(4g). The theorem now follows (cid:3) from Lemma (3.6). adX /2 ) and ht is the gaussian kernel on g, that is, (3.10) Corollary Let G be a compact connected Lie group equipped with a bi-invariant metric. Let S be the scalar curvature. The heat-trace Z(t) = tr(et4G ) of the laplacian 4G on G has the asymptotic expansion Z(t) ∼ vol(G) etS/6 as t→ 0+. Proof. This follows from Theorem (3.9) and equation (2.9). (cid:3) REFERENCES [1] A. Alekseev and E. Meinrenken, Lie theory and the Chern-Weil homomorphism, Ann. Sci. École Norm. Sup. (4) 38 (2005), no. 2, 303 -- 338, DOI 10.1016/j.ansens.2004.11.004 (English, with English and French summaries). MR2144989 (2006d:53020) [2] N. Berline, E. Getzler, and M. Vergne, Heat kernels and Dirac operators, Grundlehren Text Editions, Springer-Verlag, Berlin, 2004. Corrected reprint of the 1992 original. MR2273508 (2007m:58033) [3] M. Duflo, Opérateurs différentiels bi-invariants sur un groupe de Lie, Ann. Sci. École Norm. Sup. (4) 10 (1977), no. 2, [4] [5] 265 -- 288 (French, with English summary). MR0444841 (56 #3188) J. J. Duistermaat and V. W. Guillemin, The spectrum of positive elliptic operators and periodic bicharacteristics, Invent. Math. 29 (1975), no. 1, 39 -- 79. MR0405514 (53 #9307) L. Gårding, Dirichlet's problem for linear elliptic partial differential equations, Math. Scand. 1 (1953), 55 -- 72. MR0064979 (16,366a) [6] S. Helgason, Differential geometry, Lie groups, and symmetric spaces, Pure and Applied Mathematics, vol. 80, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1978. MR514561 (80k:53081) [7] , Groups and geometric analysis, Pure and Applied Mathematics, vol. 113, Academic Press Inc., Orlando, FL, 1984. Integral geometry, invariant differential operators, and spherical functions. MR754767 (86c:22017) [8] H. P. McKean Jr. and I. M. Singer, Curvature and the eigenvalues of the Laplacian, J. Differential Geometry 1 (1967), no. 1, 43 -- 69. MR0217739 (36 #828) [9] S. Minakshisundaram and Å. Pleijel, Some properties of the eigenfunctions of the Laplace-operator on Riemannian mani- folds, Canadian J. Math. 1 (1949), 242 -- 256. MR0031145 (11,108b) [10] B. Kostant, A cubic Dirac operator and the emergence of Euler number multiplets of representations for equal rank subgroups, Duke Math. J. 100 (1999), no. 3, 447 -- 501, DOI 10.1215/S0012-7094-99-10016-0. MR1719734 (2001k:22032) [11] H. Weyl, Das asymptotische Verteilungsgesetz der Eigenwerte linearer partieller Differentialgleichungen (mit einer An- wendung auf die Theorie der Hohlraumstrahlung), Math. Ann. 71 (1912), no. 4, 441 -- 479, DOI 10.1007/BF01456804 (German). MR1511670 [12] , Über die Abhängigkeit der Eigenschwingungen einer Membran und deren Begrenzung, J. Reine Angew. Math. 141 (1912), 1 -- 11, DOI 10.1515/CRLL.1912.141.1 (German). 8 SEUNGHUN HONG DEPARTMENT OF MATHEMATICS, THE PENNSYLVANIA STATE UNIVERSITY, UNIVERSITY PARK, PA 16802 U.S.A. E-mail address: [email protected] URL: http://www.math.psu.edu/hong
1607.04988
2
1607
2016-11-13T10:23:27
Hankel and Toeplitz operators: continuous and discrete representations
[ "math.FA", "math.SP" ]
We find a relation guaranteeing that Hankel operators realized in the space of sequences $\ell^2 ({\Bbb Z}_{+}) $ and in the space of functions $L^2 ({\Bbb R}_{+}) $ are unitarily equivalent. This allows us to obtain exhaustive spectral results for two classes of unbounded Hankel operators in the space $\ell^2 ({\Bbb Z}_{+}) $ generalizing in different directions the classical Hilbert matrix. We also discuss a link between representations of Toeplitz operators in the spaces $\ell^2 ({\Bbb Z}_{+}) $ and $L^2 ({\Bbb R}_{+}) $.
math.FA
math
HANKEL AND TOEPLITZ OPERATORS: CONTINUOUS AND DISCRETE REPRESENTATIONS D. R. YAFAEV Abstract. We find a relation guaranteeing that Hankel operators realized in the space of sequences ℓ2(Z+) and in the space of functions L2(R+) are unitarily equiva- lent. This allows us to obtain exhaustive spectral results for two classes of unbounded Hankel operators in the space ℓ2(Z+) generalizing in different directions the classical Hilbert matrix. We also discuss a link between representations of Toeplitz operators in the spaces ℓ2(Z+) and L2(R+). . A F h t a m [ 2 v 8 8 9 4 0 . 7 0 6 1 : v i X r a 1. Introduction 1.1. This paper is based on the talk given by the author at the conference "Spectral Theory and Applications" held in May 2015 in Krakow. So, it is somewhat eclectic. Our aim is to discuss various properties Hankel and Toeplitz (known also as Wiener- Hopf) operators. We refer to the books [1, 5, 11, 12, 14] for basic information on these classes of operators. Our main goal is to describe a relation between discrete and continuous representa- tions of Hankel and Toeplitz operators in a sufficiently consistent way and to draw spec- tral consequences from this relation. We do not suppose that operators are bounded, and so we are naturally led to work with quadratic forms and distributional integral kernels. As is well known, the discrete (in the space ℓ2(Z+)) and continuous (in the space L2(R+)) representations are linked by the Laguerre transform. For bounded op- erators, this yields the unitary equivalence of the corresponding operators in ℓ2(Z+) and L2(R+). However in singular cases their equivalence may be lost because the nat- ural domains of the quadratic forms in discrete and continuous representations are not linked by the Laguerre transform. As show simple examples, the continuous represen- tation seems to be more general. Passing to the Fourier transforms, one can also realize discrete Hankel and Toeplitz operators in the Hardy space H2 +(T) of functions analytic in the unit circle T and continuous operators in the Hardy space H2 +(R) of functions analytic in the upper half-plane; see, e.g., the book [6], for the precise definition of these spaces. Section 2 is of a preliminary nature. We first consider the discrete A and the con- tinuous A convolution operators in the spaces ℓ2(Z) and L2(R; dx), respectively. Of 2000 Mathematics Subject Classification. 47B25, 47B35. Key words and phrases. Unbounded Hankel and Toeplitz operators, various representations, mo- ment problems, generalized Hilbert matrices. 1 2 D. R. YAFAEV course the Fourier transform allows one to reduce these operators to the multiplications B and B in the spaces L2(T) and L2(R; dλ). The operators B and B are obviously related by a change of variables. This yields a link between the operators A and A which is given by the Laguerre transform. Our main goal in this section is to discuss explicit formulas relating matrix elements of A and the integral kernel of A. Then, we apply these results to Toeplitz operators realized in the spaces ℓ2(Z+) and L2(R+). We are aiming at a systematic presentation of known results but insist upon the case of unbounded operators. Moreover, some formulas, for example, (2.20) and (2.21), are perhaps new. 1.2. In Section 3, we pass to the main subject of this paper, to Hankel operators. We recall that Hankel operators H are defined in the space ℓ2(Z+) by the formula (Hf )n = Xm∈Z+ an+mfm, f = {fn}n∈Z+. Similarly, Hankel operators H act in the space L2(R+) by the formula (Hf)(t) =ZR+ a(t + s)f(s)ds. Note that spectral properties of the operators H are determined by the behavior of their matrix elements an as n → ∞ while, as far as the operators H are concerned, both the behavior of integral kernels a(t) as t → ∞ and t → 0 as well as their local singularities at finite points t 6= 0 are essential. Following the scheme of Section 2, we first find a link between the discrete (1.1) and continuous (1.2) realizations of Hankel operators. Then we consider the case where the matrix elements and the kernels of Hankel operators admit the integral representations (1.1) (1.2) (1.3) (1.4) and an =Zclos D a(t) =Zclos C+ zndM(z), n = 0, 1, . . . , e−ζtdΣ(ζ), t > 0, with some complex measures dM(z) and dΣ(ζ). Here D = {z ∈ C : z < 1} is the unit disc, C+ = {ζ ∈ C : Re ζ > 0} is the right half-plane, and clos D, clos C+ are the closures of these sets. Formulas (1.3) and (1.4) unify different types of integral repre- sentations of an and a(t), for example, representations in terms of Carleson measures or in terms of symbols of the corresponding bounded Hankel operators. The central result of Section 3, Theorem 3.3, formally means that the "operators" H and H are unitarily equivalent provided the measures dM(z) and dΣ(ζ) in (1.3) and (1.4) are linked by the equality dM(z) = 2α(ζ + α)−2dΣ(ζ), z = ζ − α ζ + α , (1.5) HANKEL AND TOEPLITZ OPERATORS 3 for some value of the parameter α > 0. Although quite simple, Theorem 3.3 is very useful because it relates the discrete and continuous representations directly avoiding the general construction of Section 2. More important, it allows one to translate spectral results obtained for the operator H into the results for the operator H, and vice versa. Such examples are discussed in Section 4. 1.3. Section 4 is devoted to Hankel operators generalizing in different directions two classical examples: the Hilbert matrix and the Carleman operator. To put our results into the right context, let us briefly recall basic spectral properties of these operators. The Hilbert matrix is the Hankel operator H defined by formula (1.1) where an = (n + 1)−1 for all n ≥ 0. As shown in the papers [8, 13], the spectrum of H is absolutely continuous, it is simple and coincides with the interval [0, π]. The Carleman operator is defined by formula (1.2) where a(t) = t−1. Using the Mellin transform, it is easy to show that the spectrum of the operator H is absolutely continuous, has multiplicity 2, and it also coincides with the interval [0, π]. So both these operators are bounded but not compact. It can be deduced from the results on the Hilbert matrix that Hankel operators H are bounded if an = O(n−1) and they are compact if an = o(n−1) as n → ∞. Similarly, the results on the Carleman operator imply that Hankel operators H with integral kernels a ∈ L∞loc(R) are bounded if a(t) = O(t−1) and they are compact if a(t) = o(t−1) as t → 0 and t → ∞. In Section 4 we study Hankel operators H with matrix elements an such that ann → ∞. These operators are unbounded. We exhibit two quite different cases where the spectral analysis of Hankel operators H can be carried out sufficiently explicitly. Our approach relies on Theorem 3.3 and the results on Hankel operators H with singular integral kernels a(t) obtained earlier in [18, 21]. We emphasize that some properties of Hankel operators are more transparent in the discrete representation while other properties -- in the continuous representation. Such examples are given in Section 4. So when studying Hankel operators, it is very useful to keep in mind their various representations. 2. Various representations of convolutions and Toeplitz operators 2.1. First, we recall standard relations between various spaces we consider. Let us introduce the following diagrams: f = {fn}n∈Z −−−→ f(x) = (Φu)(x) y x ℓ2(Z) yF ∗ L2(T) L−−−→ L2(R; dx) U−−−→ L2(R; dλ) xΦ (2.1) Here the unitary mapping F : L2(T) → ℓ2(Z) corresponds to expanding a function in the Fourier series: u(µ) = (F∗f )(µ) −−−→ u(λ) = (Uu)(λ) fn = (F u)n :=ZT u(µ)µ−ndm0(µ) 4 where D. R. YAFAEV dm0(µ) = (2πiµ)−1dµ is the normalized Lebesgue measure on the unit circle T. The adjoint operator F∗ : ℓ2(Z) → L2(T) acts by the formula u(µ) = (F∗f )(µ) =Xn∈Z f(x) = (Φu)(x) := (2π)−1/2Z ∞ −∞ fnµn. (2.2) e−ixλu(λ)dλ. Similarly, Φ is the Fourier transform, Of course the operator Φ : L2(R; dλ) → L2(R; dx) is unitary. The unitary operator U = Uα : L2(T) → L2(R; dλ) is defined by the equality (Uu)(λ) =q α π (λ + iα)−1u(cid:0) λ−iα λ+iα(cid:1) (2.3) where a positive parameter α can be fixed in an arbitrary way. Let us set L = ΦUF∗. Then L : ℓ2(Z) → ℓ2(R; dx) is the unitary operator, and it can be expressed in terms of the Laguerre polynomials. Recall that the Laguerre polynomials (see the book [2], Chapter 10.12) are defined by the formula Lp n(t) = n!−1ett−pdn(e−ttn+p)/dtn, n = 0, 1, . . . , n(t) has degree n; in particular, Lp Of course the polynomial Lp mials are orthogonal with respect to the measure tpe−tdt and t ≥ 0, 0(t) = 1. These polyno- p > −1. Z ∞ 0 Lp n(t)Lp m(t)tpe−tdt = Γ(n + p + 1) n! δn,m (2.4) where Γ(·) is the gamma function and δn,m is the Kronecker symbol. The parameter p > −1 is arbitrary, but we need the cases p = 0 and p = 1 only. Let us use the identity (see formula (10.12.32) in [2]) for Ln := L0 n: 1 ζ + 1/2(cid:16)2ζ − 1 2ζ + 1(cid:17)n , Re ζ > −1/2. 0 Ln(t)e−(1/2+ζ)tdt = Z ∞ 1+(x) = i(2π)−1Z ∞ Putting here ζ = −iλ and making the inverse Fourier transform, we find that Ln(2αx)e−αx dλ, n = 0, 1, . . . , e−ixλ(λ + iα)−1(cid:0)λ − iα λ + iα(cid:1)n −∞ where 1+(x) is the characteristic function of R+. Recall that the Hardy space H2 +(T) (resp. H2 (T)) consists of functions u ∈ L2(T) whose Fourier coefficients (F u)n = 0 for − n < 0 (resp., for n ≥ 0). Since the functions µn, n = 0, 1, . . ., form an orthonormal basis in the space H2 +(T), it follows from relations (2.3) and (2.6) that the functions −i(2α)1/2Ln(2αt)e−αt is an orthonormal basis in the space L2(R+). We also see that i(2α)1/2Ln(−2αt)eαt, n = 0, 1, 2, . . ., is an orthonormal basis in the space L2(R−). (2.5) (2.6) HANKEL AND TOEPLITZ OPERATORS 5 Moreover, relations (2.3) and (2.6) imply that the operator L : ℓ2(Z) → L2(R; dx) acts by the formula (Lf )(x) = −i(2α)1/2 ∞Xn=0 fnLn(2αx)e−αx 1+(x) + i(2α)1/2 ∞Xn=0 f−n−1Ln(−2αx)eαx 1+(−x), f = {fn}n∈Z. (2.7) To be precise, the unitary operator L is first defined on the dense set D ⊂ ℓ2(Z) consisting of elements f with only a finite number of non-zero components fn, and then it is extended by the continuity onto the whole space ℓ2(Z). 2.2. Next, we discuss representations of the convolution/multiplication operators in all these spaces. We start with the space ℓ2(Z) where the operator of the discrete convolution acts by the formula an−mfm, f = {fn}n∈Z. (2.8) (Af )n = Xm∈Z a[f, f ] = Xn,m∈Z Without some assumptions on the sequence a = {an}n∈Z, in general Af 6∈ ℓ2(Z) even for f ∈ D. Therefore instead of the operator A, we consider its quadratic form an−mfmfn, f ∈ D, (2.9) that consists of a finite number of terms for an arbitrary sequence a. Similarly, the convolution operator A acts in the space L2(R; dx) by the formula (2.10) (2.11) and its quadratic form is given by the equality a[f, f] =ZR (Af)(x) =ZR a(x)(cid:16)ZR F(x) =ZR a(x − y)f(y)dy, f(y)f(x + y)dy(cid:17)dx f(y)f(x + y)dy for test functions f ∈ C∞0 (R). Since, for such functions f, the function also belongs to C∞0 (R), the form (2.11) is correctly defined for a distribution a in the space C∞0 (R)′ dual to C∞0 (R). Of course the Fourier transform allows one to realize convolutions as multiplication operators. Let P = F∗D be the set of all quasi-polynomials (2.2). For a distribution b ∈ P′ (the space dual to P), we formally define the operator B in the space L2(T) by the equality (2.12) (Bu)(µ) = b(µ)u(µ), 6 D. R. YAFAEV or, in precise terms, we introduce its quadratic form Then B = F∗AF if b[u, u] =ZT b(µ) = (F∗a)(µ) =Xn∈Z b(µ)u(µ)2dm0(µ), u ∈ P. anµn, a = {an}n∈Z. (2.13) (2.14) Strictly speaking, we have a relation between the quadratic forms b[u, u] = a[f, f ] if f = F u. Similarly, we put Z := Φ∗C∞0 (R). Recall that the set Z consists of analytic functions satisfying a certain estimate at infinity (see, e.g., [4] for details). For a distribution b ∈ Z′ (the space dual to Z), we formally define the multiplication operator B in the space L2(R; dλ) by the equality (Bu)(λ) = b(λ)u(λ), or, in precise terms, we introduce its quadratic form b[u, u] =ZR b(λ)u(λ)2dλ. (2.15) (2.16) Then B = Φ∗AΦ if b = (2π)1/2Φ∗a where a ∈ C∞0 (R)′. Strictly speaking, we have a relation between the quadratic forms b[u, u] = a[f, f] if f = Φu. Obviously, the operators A, A, B and B are formally symmetric if a−n = an, a(−x) = a(x), b(µ) = b(µ) and b(λ) = b(λ). To be more precise, this means that the corresponding quadratic forms are real. We emphasize that the bases un(µ) = µn, n ∈ Z, and F un are the canonical bases in the spaces L2(T) and ℓ2(Z), respectively. On the contrary, the bases un = Uun and, especially, Φun in the spaces L2(R; dλ) and L2(R; dx) do not apparently play any distinguished role. So, it seems more natural to consider the form defined by (2.11) on functions f ∈ C∞0 (R) for a distribution a ∈ C∞0 (R)′. Similarly, we consider the form (2.16) for u ∈ Z and b ∈ Z′. 2.3. Let us find a link between the discrete and continuous representations. It is formally quite simple for the operators B and B. By definitions (2.3) and (2.12), the operator B = UBU∗ acts in the space L2(R; dλ) as the multiplication by the function (distribution) (2.17) b(λ) = b(cid:0) λ−iα λ+iα(cid:1). Its quadratic form is given by the formula (2.16) where u(λ) = (Uu)(λ) =q α π (λ + iα)−1Xn∈Z fn(cid:0) λ−iα λ+iα(cid:1)n HANKEL AND TOEPLITZ OPERATORS 7 belongs to the set PPP = UP and b belongs to the dual space PPP′. We emphasize however that this link is only formal because the sets Z and PPP of test functions u(λ) are different. It remains to directly link the representations of convolution operators in the spaces ℓ2(Z) and L2(R; dx). Recall that the unitary operator L = ΦUF∗ : ℓ2(Z) → L2(R; dx) is given by equality (2.7). Let the operator A be defined in the space ℓ2(Z) by formula (2.8) and B = UF∗AFU∗. Then A = ΦBΦ∗ = LAL∗ is the convolution in the space L2(R; dx) acting by the formula (2.10) where a = (2π)−1/2Φb and b is defined by (2.14), (2.17). The quadratic form of the operator A is defined by formula (2.11), but we have the same problem as for the multiplication operators: the domains C∞0 (R) and LD = ΦPPP of quadratic forms of the operators A and LAL∗ are different. the relation (see formula (10.12.15) in the book [2]) Our goal is to find an expression for a(x), x ∈ R, in terms of a = {an}n∈Z. Recall d dt Ln(t) = −L1 n−1(t), t > 0, ∀n ≥ 1, (2.18) for the Laguerre polynomials. Let S′ be the space dual to the Schwartz space S = S(R) of rapidly decaying C∞ functions. It follows from (2.6) that Z ∞ e−ixλ(cid:0) λ − iα d dx λ + iα(cid:1)ndλ = i( + α)(cid:0)Ln(2αx)e−αx d dx −∞ = 2π( + α)Z ∞ 1+(x)(cid:1) = −4παL1 e−ixλ(λ + iα)−1(cid:0) λ − iα n−1(2αx)e−αx −∞ λ + iα(cid:1)ndλ 1+(x) + 2πδ(x) (2.19) where δ(x) is the Dirac delta-function and the Fourier transform is understood in the sense of S′. Passing here to the complex conjugation and making the change of the variables x 7→ −x, we also see that Z ∞ −∞ e−ixλ(cid:0)λ − iα λ + iα(cid:1)−ndλ = −4παL1 n−1(−2αx)eαx 1+(−x) + 2πδ(x). Therefore it formally follows from equalities (2.14) and (2.17) that the distribution a = (2π)−1/2Φb satisfies the relation a(x) =Xn∈Z anδ(x) − 2α ∞Xn=1 L1 n−1(2αx)e−αx(cid:0)an1+(x) + a−n1+(−x)(cid:1). (2.20) An expression for this distribution can also be given in a somewhat different form. Let ϕ ∈ C∞0 (R) (or ϕ ∈ S). Then using (2.18) and integrating by parts, we see that Z ∞ a(x)ϕ(x)dx = a0ϕ(0) Z ∞ 0 ∞Xn=1 Ln(2αx)e−αx(cid:0)an(αϕ(x) − ϕ′(x)) + a−n(αϕ(−x) + ϕ′(−x))(cid:1)dx. (2.21) −∞ + 8 D. R. YAFAEV Equation (2.20) formally shows that a(x) = κδ(x) + k(x) where an = κ (2.22) Xn∈Z and k(x) is the second term in the right-hand side of (2.20). Using the identities (2.4), we can solve equation (2.20) for the coefficients an: a±n = −2αn−1Z ∞ 0 k(±x)L1 n−1(2αx)e−αxxdx, n = 1, 2, . . . ; then (2.22) yields a0. The relations between various representations of convolution/multiplication opera- tors can be summarized by the following diagrams complementing (2.1): an −−−→ a(x) −−−→ A = ΦBΦ∗ y x A y x b(µ) −−−→ b(λ) B = F∗AF −−−→ B = UBU∗ 2.4. Let us now consider Wiener-Hopf (Toeplitz) operators. We start with the space ℓ2(Z+) where Wiener-Hopf operators W act again by formula (2.8) but now n, m ∈ Z+: (2.24) an−mfm, f = {fn}n∈Z+. (W f )n = Xm∈Z+ Quite similarly to (2.9), their quadratic forms are defined by the formula (2.23) w[f, f ] = Xn,m∈Z+ an−mfmfn (2.25) where f ∈ D+ := D ∩ ℓ2(Z+) and the sequence a = {an}n∈Z is again arbitrary. A Wiener-Hopf operator W acts in the space L2(R+) by the formula a(x − y)f(y)dy, and its quadratic form is again given by the equality (Wf)(x) =ZR+ a(x)(cid:16)ZR where a ∈ C∞0 (R)′ but now f ∈ C∞0 (R+). Next, we pass to the representation of Toeplitz operators in the Hardy spaces. Denote by P+ the orthogonal projection in L2(T) onto the Hardy space H2 +(T), and let, as before, the operator B be formally defined by equality (2.12) where the distribution b ∈ P′. Then the Toeplitz operator T : H2 f(y)f(x + y)dy(cid:17)dx w[f, f] =ZR +(T) is defined by the relation (2.26) +(T) → H2 T u = P+Bu ∞Xn=0 xΦ ℓ2(Z+) yF ∗ +(T) H2 L−−−→ L2(R+) −−−→ W = ΦTΦ∗ W y x (2.28) HANKEL AND TOEPLITZ OPERATORS 9 Finally, we discuss the representation in the Hardy space H2 on elements u ∈ P+ := F∗D+; obviously, the set P+ consists of all polynomials u(µ) = Pn∈Z+ fnµn. The quadratic form t[u, u] of the operator T is given by the right-hand side of (2.13) where u ∈ P+ and b ∈ P′ are arbitrary. +(R) consisting of func- tions u ∈ L2(R) whose Fourier transforms (Φu)(x) = 0 for x < 0. Let P+ be the orthogonal projection in L2(R) onto H2 +(R), and let B be the operator (2.15). Then the Toeplitz operator T : H2 +(R) → H2 +(T) is defined by the relation Tu = P+Bu. Its quadratic form t[u, u] is given by the the right-hand side of (2.16) where u ∈ Φ∗C∞0 (R+) =: Z+ and b ∈ Z′. The relation (2.17) between the functions (distribu- tions) b(µ) and b(λ) remains of course true. Note that the Laguerre operator L : ℓ2(Z±) → L2(R±) (here Z− = Z \ Z+) and (Lf )(t) = −i(2α)1/2 f = {fn}n∈Z+, but formula (2.20) for kernel of the operator W remains true. fnLn(2αt)e−αt, For Toeplitz operators, instead of (2.1), (2.23) we have the diagrams t > 0, (2.27) U−−−→ H2 +(R) T = F∗WF −−−→ T = UTU∗ Of course all the remarks above concerning a certain difference between the discrete and continuous representations of convolution operators apply also to Wiener-Hopf operators. The case of semibounded Wiener-Hopf operators is specially discussed in [23] (the discrete representation) and in [24] (the continuous representation). 2.5. Let us say a few words about bounded operators. For Wiener-Hopf operators W defined via the quadratic form (2.25), it is the classical Toeplitz result that the operator W is bounded, that is, w[f, f ] ≤ Ckfk2 for some C > 0, if and only if a = F b where b ∈ L∞(T). The corresponding result for integral Wiener-Hopf operators W is stated explicitly in [24]. In the assertion below the Fourier transform is understood in the sense of the Schwartz space S′. Proposition 2.1. [24, Theorem 1.1] Let the form w[f, f] be defined by the relation (2.26) where f ∈ C∞0 (R+) and the distribution a ∈ C∞0 (R)′. Then w[f, f] ≤ Ckfk2, (so that the corresponding operator W is bounded ) if and only if a = (2π)−1/2Φb where b ∈ L∞(R). Moreover, kWk = kbkL∞(R). 10 D. R. YAFAEV Observe that Proposition 2.1 is not a direct consequence of the Toeplitz criterion for the boundedness of the operators W in the space ℓ2(Z+). The difference is that the domains D and C∞0 (R+) of the corresponding quadratic forms are not linked by the Laguerre transform L. 3. Hankel operators 3.1. Let us pass to Hankel operators. We start with the space ℓ2(Z+) where Hankel operators H act according to the formula (cf. (2.24)) (Hf )n = Xm∈Z+ an+mfm, f = {fn}n∈Z+, and their quadratic forms (cf. (2.25)) are defined by the formula an+mfmfn (3.1) where f ∈ D+ = D ∩ ℓ2(Z+) and the sequence a = {an}n∈Z+ is arbitrary. A Hankel operator H acts in the space L2(R+) by the formula h[f, f ] = Xn,m∈Z+ h[f, f] =ZR+ (Hf)(t) =ZR+ a(t)(cid:16)Z t F(t) =Z t 0 a(t + s)f(s)ds, f(s)f(t − s)ds(cid:17)dt. and its quadratic form is given by the equality Since, for all test functions f ∈ C∞0 (R+), the function f(s)f(t − s)ds 0 also belongs to C∞0 (R+), the form (3.2) is correctly defined for all distributions a ∈ C∞0 (R+)′ in the space dual to C∞0 (R+). +(T) +(T) is formally defined by the relation Let us now pass to the representation of Hankel operators in the Hardy spaces H2 and H2 +(R). A Hankel operator G in the space H2 Gu = P+BJu where (Ju)(µ) = ¯µu(¯µ) so that the involution J : H2 ∓(T). As before B is the multiplication operator (2.12) by the function b(µ) in the space L2(T). To be precise, we define G via its quadratic form ±(T) → H2 g[u, u] =ZT ω(µ)u(¯µ)u(µ)dm0(µ) where ω(µ) = ¯µb(µ), u ∈ P+ = F∗D+ and ω ∈ P′+. Obviously, we have g[u, u] = h[F u,F u] if ω = F∗a. (3.2) (3.3) (3.4) (3.5) HANKEL AND TOEPLITZ OPERATORS 11 Hankel operators in the Hardy space H2 +(R) ⊂ L2(R) of functions analytic in the upper half-plane are defined quite similarly. Let, as before, P+ be the orthogonal projection in L2(R) onto H2 +(R), and let B be the operator (2.15). A Hankel operator G is formally defined in the space H2 +(R) by the relation Gu = P+BJu where (Ju)(λ) = u(−λ). To be precise, we have to pass to the Fourier transform in (3.2) which yields the representation g[u, u] =ZR Ω(λ)u(−λ)u(λ)dλ, u ∈ Z+, for the quadratic form of the operator G. Obviously, we have g[u, u] = h[Φu, Φu] if Ω = (2π)1/2Φ∗a. The functions ω(µ) and Ω(λ) are known as symbols of the discrete and continuous Hankel operators. Making the change of the variables µ = (λ − iα)(λ + iα)−1 in (3.4) we see that (3.6) (3.7) (3.8) if g[Uu,Uu] = g[u, u] λ+iα(cid:1)ω(cid:0) λ−iα λ+iα(cid:1). Ω(λ) = −(cid:0) λ−iα For Hankel operators G and G, this equality plays the role of (2.17). Since a = (2π)−1/2ΦΩ and ω = F∗a, relation (2.19) where x > 0 yields the representation (cf. (2.20)) a(t) = 2α ∞Xn=0 anL1 n(2αt)e−αt, t > 0. (3.9) Let us now use that (n+ 1)−1/22αt1/2L1 in the space L2(R+). Therefore it follows from (3.9) that n(2αt)e−αt, n = 0, 1, . . ., is the orthonormal basis an = 2α(n + 1)−1ZR+ a(t)L1 n(2αt)e−αttdt. (3.10) Similarly to Toeplitz operators (cf. (2.28)), the relations between various represen- tations of Hankel operators can be summarized by the following diagram: −−−→ H = ΦHΦ∗ x G = F∗HF −−−→ G = UGU∗ H y It is easy to see that Hankel operators H, H, G and G are formally symmetric if an = ¯an, a(t) = a(t), ω(¯µ) = ω(µ) and Ω(−λ) = Ω(λ). To be more precise, this means that the corresponding quadratic forms are real. 12 D. R. YAFAEV Similarly to the case of convolutions and Toeplitz operators, the relations (3.9) and (3.10) are only formal. We also note that the domains D+ and C∞0 (R+) of the forms (3.1) and (3.2) are not related by the Laguerre transform (2.27); likewise, the domains P+ and Z+ of the forms (3.4) and(3.6) are not related by the change of variables (2.3). We emphasize however that this discrepancy is essential in singular cases only. 3.2. As far as the conditions of boundedness are concerned, the results on Hankel operators are formally similar to those on Toeplitz operators. For Hankel operators H defined via the quadratic form (3.1), it is the classical Nehari result that the operator H is bounded if and only if there exists ω ∈ L∞(T) such that a = F ω. An important difference compared to Toeplitz operators is that now the equation a = F ω does not determine the function ω uniquely and may also be satisfied with unbounded ω. The corresponding result for integral Hankel operators H is stated explicitly in [17]. Proposition 3.1. [17, Theorem 3.4] Let the form h[f, f] be defined by the relation (3.2) where the distribution a ∈ C∞0 (R+)′. Then the corresponding operator H is bounded if and only if there exists a function Ω ∈ L∞(R) such that a = (2π)−1/2ΦΩ. In this case kHk = kΩkL∞(R). By the same reasons as for Toeplitz operators (see the remark after Proposition 2.1), this result is not a direct consequence of the Nehari theorem. Of course, the operators F , Φ, U and L establishing the equivalence of various rep- resentations of Hankel operators are not unique. For example, the operator U = Uα defined by equality (2.3) depends on the parameter α > 0, and there is no distin- guished choice of this parameter. To state the problem precisely, for each of the spaces H = ℓ2(Z+), L2(R+), H2 +(R), let us introduce the group G(H) of all unitary automorphisms of the set A(H) of bounded Hankel operators in the space H. By def- inition, a unitary operator U ∈ G(H) if and only if UHU∗ ∈ A(H) for all operators H ∈ A(H). For different H, the groups G(H) are related by the transformations F , Φ, U and L. So it suffices to describe this group for one of the choices of H. +(T), H2 +(R)) admits a very explicit description. Let Dρ, It turns out that the group G(H2 (Dρu)(λ) = ρ1/2u(ρλ), ρ > 0, be the dilation operator in H2 the equation +(R), and let the involution J be defined in this space by (Ju)(λ) = iλ−1u(−λ−1). +(R)). Surprisingly, the group G(H2 Obviously, DρGD∗ρ and JGJ∗ are Hankel operators for all Hankel operators G ∈ A(H2 +(R)) is exhausted by these transformations. Let us state the precise result. Theorem 3.2. [19, Theorem A.1] A unitary operator U ∈ G(H2 has one of the two forms: U = µDρ or U = µDρJ for some µ ∈ T and ρ > 0. +(R)) if and only if it HANKEL AND TOEPLITZ OPERATORS 13 3.3. Now we consider the case when the matrix elements an of a Hankel operator H and the integral kernel a(t) of a Hankel operator H are given by formulas (1.3) and (1.4), respectively. Here dM(z) is a finite complex measure on clos D and dΣ(ζ) is a locally finite complex measure on clos C+. We suppose that the measures dM(z) and dΣ(ζ) are linked by equation (1.5); in this case M({1}) = 0. We denote by dM(z) and dΣ(ζ) the variations of these measures and assume that M(clos D) = 2αZclos C+ ζ + α−2dΣ(ζ) < ∞. (3.11) Relation (1.3) only implies that the sequence an is bounded as n → ∞, and hence the operator H is not defined in ℓ2(Z+) even on the set D. Similarly, the operator H is not defined in L2(R+) even on the set C∞0 (R+). So as usual, instead of operators we have to work with the corresponding quadratic forms (3.1) and (3.2). Formula (1.4) defines a(t) as a distribution on a set of test functions, denoted X , that can be chosen as follows. A function F ∈ X if and only if F ∈ C∞(R+), there exist limits F(k)(+0) for k = 0, 1, 2, F(+0) = 0 and F(k) ∈ L1(R+) for k = 0, 1, 2. Integrating twice by parts, we find that, for such F, the estimate holds. Therefore the form 0 (cid:12)(cid:12)Z ∞ e−ζtF(t)dt(cid:12)(cid:12) ≤ Cζ + 1−2, ha, Fi :=Zclos C+(cid:0)Z ∞ e−ζtF(t)dt(cid:1)dΣ(ζ) 0 ζ ∈ clos C+, (3.12) is well defined for all F ∈ X . Let us also introduce a set of test functions f ∈ C∞(R+), denoted Y, such that there exist limits f (k)(+0) and f (k) ∈ L1(R+) for k = 0, 1, 2. By definition (3.3), we have F ∈ X if f ∈ Y. Therefore if h ∈ X ′, then the quadratic form h[f, f] is well defined for all f ∈ Y. Note that D+ := LD+ ⊂ Y. Our main result in this subsection formally means that the "operators" H and H are unitarily equivalent provided the corresponding measures are linked by the equality (1.5). Let us state this result precisely. Theorem 3.3. Let the matrix elements an and the integral kernel a(t) be given by equalities (1.3) and (1.4), and let h[f, f ] and h[f, f] be the corresponding quadratic forms (3.1) and (3.2) defined for f ∈ D+ and f ∈ Y. Suppose that the measures dM(z) and dΣ(ζ) are linked by the relation (1.5) for some α > 0 and satisfy the condition (3.11). Then for all f ∈ D+ the identity h[f, f ] = h[Lf,Lf ] (3.13) holds. 14 D. R. YAFAEV Proof. Let f ∈ D+ and f = Lf ∈ Y. It follows from formulas (3.3) and (3.12) that in this case (3.14) Moreover, in view of the identity (2.5) and definition (2.27), we have h[f, f] =Zclos C+ 0 dΣ(ζ)(cid:0)Z ∞ e−ζtf(t)dt = −i√2α e−ζtf(t)dt(cid:1)(cid:0)Z ∞ fn(cid:16) ζ − α ζ + α(cid:17)n ζ + α ∞Xn=0 e−ζtf(s)ds(cid:1). 0 Z ∞ 0 where the sum consists of a finite number of terms. Substituting this expression into relation (3.14), we find that h[f, f] = 2α ∞Xn,m=0 fmfnZclos C+(cid:16)ζ − α ζ + α(cid:17)n+m 1 (ζ + α)2 dΣ(ζ). After the change of the variables z = ζ−α ζ+α, we see that this expression equals h[f, f] = ∞Xn,m=0 fmfnZclos D zn+mdM(z) where the measure dM(z) is defined by relation (1.5). According to (1.3) this expression coincides with (3.1). This concludes the proof of the identity (3.13). (cid:3) Corollary 3.4. Suppose that the expression (3.13) is estimated by Ckfk2 with some C > 0. Then there exist bounded operators H and H corresponding to these quadratic forms, and they are unitarily equivalent: HL = LH. Obviously, coefficients (1.3) and function (1.4) are real if the measures dM(z) and dΣ(ζ) are invariant with respect to the complex conjugation: M(X) = M(X) and Σ(Y ) = Σ(Y ) (3.15) for all X ⊂ clos D and Y ⊂ clos C+. Corollary 3.5. Under the assumptions of Theorem 3.3, let condition (3.15) be satisfied. Suppose that an operator H is defined on D+ and (Hf, f ) = h[f, f ] for f ∈ D+, or equivalently an operator H is defined on D+ and (Hf, f) = h[f, f] for f ∈ D+. If H is essentially self-adjoint on D+ (or equivalently H is essentially self-adjoint on D+), then their closures are unitarily equivalent: (clos H) L = L(clos H). In particular, the measures dM(z) and dΣ(ζ) may be supported by the intervals [−1, 1] and [0,∞), respectively; as before we suppose also that M({1}) = 0. Then Theorem 3.3 reads as follows. Theorem 3.6. Let the matrix elements an and the kernel a(t) be given by the equalities an =Z 1 −1 νndM(ν), n = 0, 1, . . . , (3.16) HANKEL AND TOEPLITZ OPERATORS 15 and a(t) =Z ∞ 0 e−λtdΣ(λ), t > 0. Suppose that the measures dM(ν) and dΣ(λ) are linked by the relation dM(ν) = 2α(λ + α)−2dΣ(λ), ν = λ − α λ + α , (3.17) (3.18) for some α > 0 and satisfy the condition M([−1, 1)) = 2αZ ∞ Then for all f ∈ D+ the identity (3.13) holds. 0 λ + α−2dΣ(λ) < ∞. The following result is a combination of Theorem 3.6, of Theorems 1.2 and 3.4 in [22] on the discrete case and the preceding results (Theorem 3.10) of [18] on the continuous case. Theorem 3.7. Under the assumptions of Theorem 3.6 suppose that the measure (3.18) is non-negative and that M({−1}) = Σ({0}) = 0. Then: (i) The form h[f, f ] defined on D+ is closable, and it is closed on the set of elements (ii) The form h[f, f] defined on D+ is closable, and it is closed on the set of functions f = (f0, f1, . . .) ∈ ℓ2(Z+) such that h[f, f ] =Z 1 −1(cid:12)(cid:12)(cid:12) ∞Xn=0 (cid:12)(cid:12)(cid:12)Z ∞ f ∈ L2(R+) such that h[f, f] =Z ∞ 0 0 2 fnνn(cid:12)(cid:12)(cid:12) e−λtf(t)dt(cid:12)(cid:12)(cid:12) 2 dM(ν) < ∞. dΣ(λ) < ∞. (iii) The non-negative operators H and H corresponding to these quadratic forms are unitarily equivalent: HL = LH. Sometimes the representations (1.3) and (1.4) are too restrictive. For example, if a Hankel operator H has kernel a(t) = tke−αt, Re α > 0 (such H has a finite rank), then representation (1.4) is formally satisfied with dΣ(ζ) = δ(k)(ζ − α)dM0(ζ) (dM0(ζ) is the planar Lebesgue measure and δ(ζ) is the delta-function) which is a measure for k = 0 only. A very general situation where a ∈ C∞0 (R+)′ is an arbitrary distribution was considered in [20]. Then the role of the measure dΣ(ζ) is played also by a distribution which was called the sigma-function of the Hankel operator H. 3.4. Although quite simple, Theorem 3.3 is very useful for relating the discrete and continuous representations. Let us now discuss the case of bounded Hankel operators. 16 D. R. YAFAEV Let D(z0, r) = {z − z0 < r} be the disc in the complex plane C. Recall that dM(z) is called the Carleson measure on the unit disc D if sup µ∈T,r∈(0,r0) r−1M(D(µ, r) ∩ D) < ∞ (3.19) where r0 is some fixed small number. Similarly, dΣ(ζ) is called the Carleson measure on the right half-plane C+ if sup λ∈R,R>0 R−1Σ(D(iλ, R) ∩ C+) < ∞. (3.20) Theorem 3.8. Let the measures dM(z) and dΣ(ζ) be linked by the equation (1.5). Then dM(z) is a Carleson measure on the unit disc D if and and only if dΣ(ζ) is a Carleson measure on the half-plane C+. This assertion is checked in the Appendix by straightforward but rather tedious cal- culations. Theorem 3.8 can also be indirectly deduced from general results on analytic functions (see, e.g., Section E in Chapter VIII of [7]). Theorem 3.9. A Hankel operator H in the space ℓ2(Z+) is bounded if and only if its matrix elements admit the representation (1.3) where dM(z) is some Carleson measure on the unit disc D. This result is stated as Theorem 7.4 in Chapter 1 of the book [12] and can be easily deduced from Theorem A2.12 of [12] where the representation of the symbol of a bounded Hankel operator as the Poisson balayage of some Carleson measure is given. For the proof of the latter result, we refer to Section G in Chapter X of the book [7] or pages 271, 272 of the book [3]. Putting together Theorems 3.3, 3.9 and Lemma 3.8, we get the following result. Theorem 3.10. A Hankel operator H in the space L2(R+) is bounded if and only if its kernel admits the representation (1.4) with some Carleson measure dΣ(ζ) on the right half-plane C+. It also follows from Corollary 3.4 that if dM(z) and dΣ(ζ) in the equation (1.5) are Carleson measures, then the corresponding operators H and H are unitarily equivalent. Of course the representations (1.3) and (1.4) are highly non-unique. Let us give a simple Example 3.11. Let dM0(z) be the Lebesgue measure on D, and let the measure dN(z) be supported by the point 0 so that N({0}) = 1 and N(D \ {0}) = 0. Put Then ZD zndM(z) =ZD dM(z) = dM0(z) − πdN(z). zndM0(z) =Z 1 drrn+1Z 2π 0 0 einθdθ = 0 if n ≥ 1, HANKEL AND TOEPLITZ OPERATORS 17 and ZD dM(z) =ZD dM0(z) − π = 0. 3.5. In particular, the measures dM(z) and dΣ(ζ) may be carried by the intervals (−1, 1) and (0,∞), respectively. Then conditions (3.19) and (3.20) mean that M((1 − ε, 1)) = O(ε) and M((−1,−1 + ε)) = O(ε) as ε → 0 and sup R>0 R−1Σ(0, R) < ∞. (3.21) (3.22) For non-negative Hankel operators, the results of the previous subsection can be stated in a simpler and more definite form. Recall that the Hankel form h[f, f ] was defined by relation (3.1). The following assertion is the classical result of H. Widom. Theorem 3.12. [15, Theorem 3.1] Suppose that h[f, f ] ≥ 0 for all f ∈ D+. Then the following conditions are equivalent: (i) The operator H is bounded. (ii) The representation (3.16) holds with a non-negative measure dM(ν) satisfying condition (3.21). (iii) an = O(n−1) as n → ∞. Let us now state a continuous analogue of Theorem 3.12. Theorem 3.13. Let a ∈ C∞0 (R+)′. Suppose that h[f, f] ≥ 0 for all f ∈ C∞0 (R+). Then the following conditions are equivalent: (i) The operator H is bounded. (ii) The representation (3.17) holds with a non-negative measure dΣ(λ) satisfying condition (3.22). (iii) a(t) = O(t−1) as t → ∞ and t → 0. Proof. According to Theorem 5.1 in [20] the condition h[f, f] ≥ 0 implies that, for some non-negative measure dΣ(λ) on R, a(t) =Z ∞ −∞ e−λtdΣ(λ) (3.23) where the integral converges for all t > 0. If Σ(R \ R+) > 0, then a(t) ≥ c > 0 so that the operator H cannot be bounded. Thus representation (3.23) reduces to (3.17). Let the measures dΣ(λ) and dM(ν) be linked by equation (3.18). By Lemma 3.8 (which is quite easy in this particular case) the conditions (3.21) and (3.22) are equivalent and hence, by Theorem 3.3, the operator H is bounded if and only if the operator H with matrix elements (3.16) is bounded. So the conditions (i) and (ii) are equivalent according to Theorem 3.12. 18 D. R. YAFAEV It follows from (3.23) that a ∈ C∞(R+). So under assumption (iii) the operator H is bounded because the Carleman operator (it has integral kernel a(t) = t−1) is bounded. Conversely, integrating in (3.17) by parts we see that a(t) = tZ ∞ 0 e−λtΣ(λ)dλ. Therefore condition (3.22) implies (iii). (cid:3) Of course under the assumptions of Theorems 3.12 and 3.13 the measures dM(ν) and dΣ(λ) are unique. 3.6. By their definitions, Carleson measures are carried by the open sets D or C+. Let us now consider the opposite case when the measure dM(z) is supported on the unit circle T and the corresponding measure dΣ(ζ) is supported on the line Re ζ = 0. Then relations (1.3), (1.4) and (1.5) read, respectively, as an =ZT a(t) =ZR µndM(µ), n = 0, 1, . . . , e−iλtdΣ(iλ), t > 0, (3.24) (3.25) . (3.26) (4.1) (4.2) and dM(µ) = −2α(λ − iα)−2dΣ(iλ), µ = λ + iα λ − iα In particular, if the measures dM(µ) on T and dΣ(iλ) on R are absolutely continuous, that is, dM(µ) = ω(¯µ)dm0(µ) and dΣ(iλ) = (2π)−1Ω(λ)dλ, then (3.24), (3.25) give the representations of the matrix elements an of H and the integral kernel a(t) of H in terms of their symbols ω(µ) and Ω(λ) (see formulas (3.5) and (3.7)). In this case (3.26) yields the standard relation (3.8) between these symbols. We recall that by the Nehari theorem [10], the operators H or H are bounded if and only if the symbols ω or Ω can be chosen as bounded functions, but the construction above does not require this condition. 4.1. In this section, we consider signed real measures dΣ(λ) and dM(ν) satisfying the assumptions of Theorem 3.6, but we do not assume that they satisfy the Carleson conditions (3.21) or (3.22). We suppose that these measures are absolutely continuous. Then (3.16) and (3.17) read as 4. Singular case a(t) =Z ∞ 0 e−λtσ(λ)dλ and an =Z 1 −1 νnη(ν)dν, n = 0, 1, . . . , HANKEL AND TOEPLITZ OPERATORS 19 with some real functions σ such that Z ∞ 0 (λ + 1)−2σ(λ)dλ < ∞ (4.3) and η ∈ L1(−1, 1). The relation (3.18) yields η(ν) = σ(cid:0)α 1 + ν 1 − ν(cid:1). It turns out that even this particular case leads to interesting examples. Our plan is to use the results obtained in [18, 21] for integral Hankel operators H, and then with the help of Theorem 3.6 to state new results for matrix Hankel operators H. Let us first illustrate our approach on the Carleman operator and the Hilbert matrix already discussed in Subsection 1.3. If σ(λ) = 1, then according to (4.1) we have a(t) = t−1 which yields the Carleman operator H. It follows from (4.3) that η(ν) = 1 and hence, by (4.2), the corresponding "discrete" Hankel operator H = L∗HL has matrix elements an = (1 + (−1)n)(n + 1)−1. Let 1X be the characteristic function of a set X ⊂ R. Suppose that η(ν) = 1(0,1)(ν). Then according to (4.2) we have an = (n + 1)−1 which yields the Hilbert matrix H. It follows from (4.3) that σ(λ) = 1(α,∞)(λ) so that the corresponding "continuous" Hankel operator H = LHL∗ has integral hernel a(t) = t−1e−αt. Our goal is to study the case when the sigma-function η(ν) is singular at the points ν = ±1 and the Hankel operator H with matrix elements (4.2) is unbounded. We will consider two families of such Hankel operators. The first family admits an explicit spectral analysis (Theorem 4.3). Spectral information about the second family is more limited (Theorem 4.6). 4.2. First we consider functions η(ν) with arbitrary logarithmic singularities at the points ν = 1 and ν = −1. To be precise, we assume that η(ν) = pXl=0 γl lnl(cid:0)α 1 + ν 1 − ν(cid:1), p ≥ 1, (4.4) where γl, l = 0, 1, . . . , p, are any real numbers. Without loss of generality, we set γp = 1. According to (4.2) the matrix elements of the corresponding Hankel operator (1.1) are given by the formulas an = pXl=0 γlZ 1 −1 νn lnl(cid:0)α 1 + ν 1 − ν(cid:1)dν. Note (see Proposition 4.5, below) that an = (1 + (−1)n+p) lnp n n (cid:0)1 + O( 1 ln n )(cid:1) (4.5) (4.6) 20 D. R. YAFAEV as n → ∞. It can be expected that such Hankel operator H is unbounded because η(ν)dν is not the Carleson measure and ann → ∞ as n → ∞. Nevertheless, since an = ¯an and {an}n∈Z+ ∈ ℓ2(Z+), the operator H is well defined and symmetric on the dense set D+ ⊂ ℓ2(Z+) of sequences with only a finite number of non-zero components. Our study of the Hankel operator H with matrix elements (4.5) relies on a combi- nation of Theorem 3.6 with the results of [16], [21] on integral Hankel operators H in the space L2(R+). In view of (4.3) the sigma-function of H = LHL∗ equals σ(λ) = γl lnl λ, p ≥ 1, pXl=0 and its integral kernel a(t) is given by formula (4.1). Let us define a differential operator L = v pXl=0 γlDlv, D = id/dξ, (4.7) of order p in the space L2(R) where v is the operator of multiplication by the universal function √π v(ξ) = pcosh(πξ) Γ(1/2 + iξ)Z ∞ (F f)(ξ) = (2π)−1/2 Γ(1/2 + iξ) . 0 t−1/2−iξf(t)dt Let us introduce also a unitary transformation F : L2(R+) → L2(R) by the formula (so it is the Mellin transform, up to an insignificant phase factor). Note that F f ∈ S if f ∈ D+ = LD+. Putting together Theorem 3.6 and Theorem 3.2 of [16], we can state the following result. Proposition 4.1. Let matrix elements of the Hankel operator H be defined by formula (4.5) where the function η(ν)is given by formula (4.4). Then for all f ∈ D+, the identity holds. (Hf, f ) = (LFLf, FLf ). It is shown in Theorem 3.13 of [16] that the operator L defined on the set C p 0 (R+) is essentially self-adjoint. The same arguments work if L is defined on the set F D+. Therefore Proposition 4.1 allows us to obtain a similar result for the operator H. Proposition 4.2. The Hankel operator H with matrix elements (4.5) is essentially self-adjoint on the set D+. Our goal is to obtain rather a complete information about the spectral structure of the closure of H which will also be denoted H. HANKEL AND TOEPLITZ OPERATORS 21 Theorem 4.3. Let matrix elements of the Hankel operator H be defined by formula (4.5) where γp = 1 and γℓ for ℓ ≤ p − 1 are arbitrary real numbers. Then (i) The spectrum of the operator H is absolutely continuous except eigenvalues that may accumulate to zero and infinity only. (ii) The absolutely continuous spectrum of the operator H covers R and is simple for p odd. It coincides with [0,∞) and has multiplicity 2 for p even. (iii) The point 0 is not an eigenvalue of the operator H. If p is odd, then the multi- plicities of eigenvalues of the operator H are bounded by (p − 1)/2. If p is even, then the multiplicities of positive eigenvalues are bounded by p/2 − 1, and the multiplicities of negative eigenvalues are bounded by p/2. Given Proposition 4.1, this result is a consequence (except the assertion about the point 0 which is proven in Theorem 4.7 of [16]) of the corresponding statement, The- orem 4.8 in [21], for the differential operator L. We emphasize that the method of [21] yields a sufficiently explicit spectral analysis of the operator L and, in particular, information about its eigenfunctions of the continuous spectrum. In view of the uni- tary equivalence, this yields the corresponding results for the Hankel operator H (and H = LHL∗), but we will not dwell upon them. It remains to justify asymptotic relation (4.6). Lemma 4.4. The asymptotic relation Z 1 0 νn lnl(1 − ν)dν = (− ln n)l n (cid:0)1 + O( 1 ln n )(cid:1). as n → ∞ holds. Proof. Differentiating the formula Z 1 0 νn(1 − ν)ǫdν = Γ(n + 1)Γ(ǫ + 1) Γ(ǫ + n + 2) l times in ǫ and then putting ǫ = 0, we see that (4.8) (4.9) . (4.10) Z 1 0 νn lnl(1 − ν)dν = ∂l ∂ǫl(cid:16) Γ(n + 1)Γ(ǫ + 1) Γ(ǫ + n + 2) (cid:17)(cid:12)(cid:12)(cid:12)ǫ=0 According to formula (1.18.4) in [2] we have Γ(n + 1) Γ(ǫ + n + 2) = n−ǫ−1(cid:0)1 + O(n−1)(cid:1), (4.11) and this asymptotic formula can be infinitely differentiated in ǫ. In view of (4.10), this yields (4.8). (cid:3) 22 D. R. YAFAEV Integrating separately over the intervals (−1, 0), (0, 1) and making the change of variables ν 7→ −ν in the integral over (−1, 0), we see that Z 1 −1 νn lnl(cid:0)α 1 + ν 1 − ν(cid:1)dν = (−1)lZ 1 +(−1)nZ 1 0 0 νn(cid:0) ln(1 − ν) − ln(1 + ν) − ln α(cid:1)l νn(cid:0) ln(1 − ν) − ln(1 + ν) + ln α(cid:1)ldν. dν The asymptotic behavior of the integrals on the right is determined by a neighborhood of the point ν = 1 where the function ln(1 − ν) is singular. The terms with ln(1 + ν) and ln α do not give a contribution to the leading term of the asymptotics. Therefore putting together formulas (4.5) (where γp = 1) and (4.8) we obtain the following result. Proposition 4.5. Matrix elements (4.5) of the Hankel operator H satisfy the asymp- totic relation (4.6). We finally note that for p = 1, the differential operator (4.7) reduces by an ex- plicit unitary transformation to the operator id/dξ. Therefore the same is true for the corresponding Hankel operators H and H, see [16] for details. 4.3. Now we consider even more singular compared to (4.4) case when the function η(ν) has power singularities at the points ν = 1 or ν = −1. Let q 6= 0, q ∈ (−1, 1), η(ν) =(cid:16) 1 + ν 1 − ν(cid:17)q , and let the sequence an be defined by formula (4.2). Then according to Theorem 3.7 the non-negative Hankel operator H with such matrix elements is correctly defined via its quadratic form. Our goal here is to describe its spectral structure. Theorem 4.6. The spectrum of the Hankel operator H = H(q) with the matrix ele- ments (4.12) an = an(q) =Z 1 νn(cid:16) 1 + ν 1 − ν(cid:17)q −1 dν, q ∈ (−1, 1), q 6= 0, (4.13) is absolutely continuous, coincides with the half-axis [0,∞) and has constant multiplic- ity. In view of Theorem 3.7 this result can be deduced from the corresponding assertion for the Hankel operator H = H(q) in the space L2(R+). Indeed, it follows from formula (4.3) that the sigma-function of the operator H = LHL∗ is σ(λ) = 2qλq, and hence according to relation (4.1) its integral kernel equals a(t) = aq(t) = 2qZ ∞ 0 e−tλλqdλ = 2qΓ(q + 1)t−q−1. (4.14) Hankel operators H with such kernels were studied in [18]. It was shown in Theorem 1.2 that for all q > −1, q 6= 0, the spectra of the operators H are absolutely continuous, HANKEL AND TOEPLITZ OPERATORS 23 coincide with the half-axis [0,∞) and have constant multiplicity. q < 1, this result yields Theorem 4.6. In particular, for Remark 4.7. 1. The proof of Theorem 1.2 in [18] relies only on the invariance of the Hankel operator with kernel (4.14) with respect to the group of dilations. So, spectral information about the Hankel operators with matrix elements (4.13) is more limited than under the assumptions of Theorem 4.3. Even the spectral multiplicity of H is unknown. We recall that according to the fundamental results of [9] the spectral multiplicity of a positive bounded Hankel operator does not exceed 2, but, strictly speaking, this result is not applicable because the operator H in Theorem 4.6 is unbounded. Actually, we expect that the spectrum of H is simple since the kernel (4.14) has only one singular point. This point is t = 0 if q > 0 and t = ∞ if q < 0. 2. It is by no means obvious how to prove Theorem 4.6 directly in ℓ2(Z+) because the realization in this space of the group of dilations in the space L2(R+) is not transparent. 3. It follows from definition (4.13) that an(q) = (−1)nan(−q) (4.15) and hence H(−q) = V ∗H(q)V where the unitary operator V is defined on sequences f = (f0, f1, . . .) by the formula (V f )n = (−1)nfn. Thus the Hankel operators H(−q) and H(q) with kernels (4.14) in the space L2(R+) are also unitarily equivalent. This fact does not look obvious in the continuous representation. 4. For q ≥ 1, equality (4.13) makes no sense. In this case there is no reasonable interpretation of the Hankel operator H with sigma-function (4.12) in the space ℓ2(Z+) although the Hankel operator H with integral kernel (4.14) is well defined in the space L2(R+). Finally, we find the asymptotics of the matrix elements (4.13) as n → ∞. In view of formula (4.15) we may suppose that q ∈ (0, 1). Then the asymptotics of the integral (4.13) is determined by a neighborhood of the point ν = 1. So, we write formula (4.13) as an = 2qZ 1 0 νn(1 − ν)−qdν +Z 1 0 νn(1 − ν)−q(cid:0)(1 + ν)q − 2q(cid:1)dν + (−1)nZ 1 0 νn(cid:16) 1 − ν 1 + ν(cid:17)q dν. The first integral on the right coincides with expression (4.9) for ǫ = −q, and its asymptotics as n → ∞ is given by formula (4.11). The second and third integrals are O(n−2+q) and O(n−1−q), respectively. This yields the following result. Proposition 4.8. The sequence (4.13) satisfies the asymptotic relation an = (sgn q)n2qΓ(1 − q)nq−1(cid:0)1 + O(n−ε)(cid:1), n → ∞, where ε = min{1, 2q}. 24 D. R. YAFAEV As could be expected, an → 0 as n → ∞ but much slower than sequence (4.6). Appendix A. Proof of Theorem 3.8 A.1. Suppose, for definiteness, that the equations (1.5) are satisfied with α = 1. Then ζ = 1 + z 1 − z =: ω(z) and z = ω−1(ζ) = ζ − 1 ζ + 1 . A standard calculation shows that if r < z0 − 1, then ω(D(z0, r)) = D(ζ0, R) (A.1) (A.2) (A.3) (A.5) (A.6) (A.7) (A.8) where ζ0 = (1 + z0)(1 − ¯z0) + r2 1 − z02 − r2 , R = 2r 1 − z02 − r2 . Conversely, supposing that R < ζ0 + 1, we see that relation (A.1) holds true with z0 = (ζ0 − 1)(¯ζ0 + 1) − R2 ζ0 + 12 − R2 , r = 2R ζ0 + 12 − R2 . In particular, if z0 = eiθ, θ ∈ [0, 2π), then z0 − 1 = 2 sin(θ/2) and (A.2) reads as ζ0 = 2i sin θ + r2 4 sin2(θ/2) − r2 2r 4 sin2(θ/2) − r2 , R = . (A.4) Similarly, if ζ0 = iλ, λ ∈ R, then ζ0 + 1 = √λ2 + 1 and (A.3) reads as z0 = λ2 + 2iλ − 1 − R2 λ2 + 1 − R2 , r = 2R λ2 + 1 − R2 . A.2. By the proof of Theorem 3.8, we may suppose that the measures dM(z) and dΣ(ζ) are non-negative. It is also convenient to extend these measures onto the whole complex plane C setting M(C \ D) = 0 and Σ(C \ C+) = 0. Below C (sometimes with indices) are different positive constants whose precise values are of no importance. Let us define a measure dfM(ζ) on C by the relation fM (Y ) = M(ω−1(Y )) for an arbitrary set Y ⊂ C. In particular, according to (A.1) we have It follows from (1.5) that Lemma A.1. M(D(z0, r)) = fM(D(ζ0, R)). dfM (ζ) = 2ζ + 1−2dΣ(ζ). • Condition (3.19) is satisfied if sup sup θ∈[0,2π) r≤2γ0 sin(θ/2) r−1M(D(eiθ, r)) < ∞ It is convenient to state the conditions (3.19) and (3.20) in an equivalent way. HANKEL AND TOEPLITZ OPERATORS 25 for some γ0 ∈ (0, 1) and, for some r0 > 0, r−1M(D(1, r)) < ∞. sup r∈(0,r0) • Condition (3.20) is satisfied if sup R≤δ0√λ2+1 sup λ∈R R−1Σ(D(iλ, R)) < ∞ for some δ0 ∈ (0, 1) and, for some R0 > 0, sup R≥R0 R−1Σ(D(0, R)) < ∞. (A.9) (A.10) (A.11) Proof. Indeed, if r ≥ 2γ0 sin(θ/2), then D(eiθ, r) ⊂ D(1, γr) for γ = 1 + γ−1 (3.19) for such θ and r follows from (A.9). Similarly, if R ≥ δ0√λ2 + 1, then R ≥ δ0 and D(iλ, R) ⊂ D(0, δR) for δ = 1 + δ−1 Therefore (3.20) for such λ and R follows from (A.11). 0 . Therefore 0 . (cid:3) First, we consider the conditions (A.8) and (A.10). Lemma A.2. Condition (3.20) on dΣ implies condition (A.8) on dM . Proof. Let r ≤ 2γ0 sin(θ/2), λ = cot(θ/2), and let ζ0 and R be given by formulas (A.4). Since iλ ∈ D(ζ0, R), we have D(ζ0, R) ⊂ D(iλ, 2R), and hence according to (3.20) (A.12) Put z0 = eiθ, then 1 − z0 = 2 sin(θ/2). If ζ ∈ D(ζ0, R), then z = ω−1(ζ) ∈ D(z0, r) and hence z0 − z ≤ r ≤ 2γ0 sin(θ/2). So, we have Σ(D(ζ0, R)) ≤ CR. 21 + ζ−1 = 1 − z ≤ 1 − z0 + z0 − z ≤ 2(1 + γ0) sin(θ/2). In view of (A.7) it now follows from (A.12) that M(D(z0, r)) = fM (D(ζ0, R)) ≤ 2(1 + γ0)2 sin2(θ/2)Σ(D(ζ0, R)) ≤ CR sin2(θ/2). (A.13) Since, according to the second formula (A.4), r2 R sin2(θ/2) = 2−1r(cid:16)1 − 4 sin2(θ/2)(cid:17)−1 estimate (A.13) yields (A.8). Next, we prove the converse assertion. ≤ 2−1(cid:0)1 − γ2 0(cid:1)−1 r, (cid:3) Lemma A.3. Condition (3.19) on dM implies condition (A.10) on dΣ. relations (A.6) and (A.7), we have Proof. Consider the disc D(iλ, R) where R ≤ δ0√λ2 + 1 with δ0 < 1. Σ(D(iλ, R)) ≤ C(λ2 + 1)fM(D(iλ, R)) = C(λ2 + 1)M(D(z0, r)) In view of (A.14) 26 D. R. YAFAEV where z0 and r are given by formulas (A.5). Since eiθ = z0z0−1 ∈ D(z0, r) we see that D(z0, r) ⊂ D(eiθ, 2r). Thus, it follows from condition (3.19) that the right-hand side of (A.14) is bounded by C(λ2 + 1)r = 2CR(cid:16)1 − R2 λ2 + 1(cid:17)−1 Therefore (A.14) implies condition (A.10). ≤ 2(cid:0)1 − δ2 0(cid:1)−1 CR. A.3. It remains to compare the conditions (A.9) and (A.11). Lemma A.4. Conditions (A.9) and (A.11) are equivalent. Proof. Let (A.11) be satisfied. Observe that (cid:3) (A.15) ω(D(1, r)) = C \ D(−1, 2r−1) and set φ(R) = Σ(D(−1, R)). In view of (A.7) we have fM (D(−1, R′) \ D(−1, R)) =Z R′ R dfM(D(−1, ρ)) = 2Z R′ R Integrating here by parts and then passing to the limit R′ → ∞, we see that ρ−2dφ(ρ). 2−1fM (C \ D(−1, R)) = −R−2φ(R) + 2Z ∞ R ρ−3φ(ρ)dρ. Under assumption (A.11) the right- and hence the left-hand sides here are O(R−1) as R → ∞. Thus, according to (A.6) and (A.15), we have r → 0. Conversely, let (A.9) be satisfied. Again in view of (A.7) we have M(D(1, r)) = fM (C \ D(−1, 2r−1)) = O(r), 2Σ(D(−1, R)) = 2Z R dΣ(D(−1, ρ)) =Z R 0 0 ρ2dfM (D(−1, ρ)). Since, by (A.15), it follows that fM (C \ D(−1, R)) = M(D(1, 2R−1)) =: ψ(R), ρ2dψ(ρ) = −R2ψ(R) + 2Z R 2Σ(D(−1, R)) = −Z R 0 0 ρψ(ρ)dρ. Under assumption (A.9), ψ(R) = O(R−1) as R → ∞. So, the right- and hence the left-hand sides here are O(R) as R → ∞ which proves (A.11). Now it easy to conclude the proof of Theorem 3.8. Let condition (3.20) on dΣ be true. Then, by Lemma A.2, dM satisfies (A.8) and, by Lemma A.4, it satisfies (A.9) . So, Lemma A.1 implies that condition (3.19) is fulfilled. Similarly, let condition (3.19) on dM be true. Then, by Lemma A.3, dΣ satisfies (A.10) and, by Lemma A.4, it satisfies (A.11). So, Lemma A.1 implies that condition (3.20) is fulfilled. (cid:3) (cid:3) HANKEL AND TOEPLITZ OPERATORS 27 References [1] A. Bottcher, B. Silbermann, Analysis of Toeplitz operators, Springer-Verlag, 2006. [2] A. Erd´elyi, W. Magnus, F. Oberhettinger, F. G. Tricomi, Higher transcendental functions, Vol. 1, 2, McGraw-Hill, New York-Toronto-London, 1953. [3] J. B. Garnett, Bounded analytic functions, Academic Press, New York, 1981. [4] I. M. Gel'fand, G. E. Shilov, Generalized functions. Vol. 1, Academic Press, New York and London, 1964. [5] I. Gohberg, S. Goldberg, M. Kaashoek, Classes of linear operators, vol. 1, Birkhauser, 1990. [6] K. Hoffman, Banach spaces of analytic functions, Prentice-Hall, Inc., Englewood Cliffs, New York, 1962. [7] P. Koosis, Introduction to H [8] W. Magnus, On the spectrum of Hilbert's matrix, Amer. J. Math. 72 (1950), 405-412. [9] A. V. Megretskii, V. V. Peller, S. R. Treil, The inverse spectral problem for self-adjoint Hankel p spaces, LMS, Lecture Note Series 40, Cambridge Univ. Press, 1980. operators, Acta Math. 174 (1995), 241-309. [10] Z. Nehari, On bounded bilinear forms, Ann. Math. 65 (1957), 153-162. [11] N. K. Nikolski, Operators, functions, and systems: an easy reading, vol. I: Hardy, Hankel, and Toeplitz, Math. Surveys and Monographs vol. 92, Amer. Math. Soc., Providence, Rhode Island, 2002. [12] V. V. Peller, Hankel operators and their applications, Springer Verlag, 2002. [13] M. Rosenblum, On the Hilbert matrix, I, II, Proc. Amer. Math. Soc. 9 (1958), 137-140, 581-585. [14] M. Rosenblum, J. Rovnyak, Hardy classes and operator theory, Oxford Univ. Press, 1985. [15] H. Widom, Hankel matrices, Trans. Amer. Math. Soc. 121 (1966), 1-35. [16] D. R. Yafaev, Diagonalizations of two classes of unbounded Hankel operators, Bulletin Math. Sciences 4 (2014), 175-198. [17] D. R. Yafaev, Criteria for Hankel operators to be sign-definite, Analysis & PDE 8 (2015), 183-221. [18] D. R. Yafaev, Quasi-Carleman operators and their spectral properties, Integral Eq. Op. Th. 81 (2015), 499-534. [19] D. R. Yafaev, On finite rank Hankel operators, J. Funct. Anal. 268 (2015), 1808-1839. [20] D. R. Yafaev, Quasi-diagonalization of Hankel operators, preprint, arXiv:1403.3941(2014); to appear in J. d'Analyse Math´ematique. [21] D. R. Yafaev, Spectral and scattering theory for differential and Hankel operator, arXiv: 1511.04683 (2015). [22] D. R. Yafaev, Unbounded Hankel operator and moment problems, Integral Eq. Oper. Theory, 85 (2016), 289-300. [23] D. R. Yafaev, On semibounded Toeplitz operators, arXiv: 1603.06229 (2016), to appear in J. Oper. Theory. [24] D. R. Yafaev, On semibounded Wiener-Hopf operators, arXiv: 1606.01361 (2016). IRMAR, Universit´e de Rennes I, Campus de Beaulieu, 35042 Rennes Cedex, FRANCE E-mail address: [email protected]
1301.2837
1
1301
2013-01-14T01:16:29
A Note on $\Gamma_n$-isometries
[ "math.FA" ]
In this note we characterize the distinguished boundary of the symmetrized polydisc and thereby develop a model theory for $\Gamma_n$-isometries along the lines of \cite{AY}. We further prove that for invariant subspaces of $\Gamma_n$-isometries, similar to the case $n=2$ \cite{S}, Beurling-Lax-Halmos type representation holds.
math.FA
math
A NOTE ON Γn-ISOMETRIES SHIBANANDA BISWAS AND SUBRATA SHYAM ROY Abstract. In this note we characterize the distinguished boundary of the symmetrized polydisc and thereby develop a model theory for Γn-isometries along the lines of [1]. We further prove that for invariant subspaces of Γn-isometries, similar to the case n = 2 [9], Beurling-Lax-Halmos type representation holds. 1. Introduction 3 1 0 2 n a J 4 1 ] . A F h t a m [ 1 v 7 3 8 2 . 1 0 3 1 : v i X r a We denote by D and D the open and closed unit discs in the complex plane C. Let si, i ≥ 0, be the elementary symmetric function in n variables of degree i, that is, si is the sum of all products of i distinct variables zi so that s0 = 1 and si(z1, . . . , zn) = zk1 · · · zki. X1≤k1<k2<...<ki≤n For n ≥ 1, let s : Cn −→ Cn be the function of symmetrization given by the formula s(z1, . . . , zn) =(cid:0)s1(z1, . . . , zn), . . . , sn(z1, . . . , zn)(cid:1). n ) under the map s of the unit n-polydisc is known as the symmetrized n-disc. The image Γn := s(D The map s is a proper holomorphic map [8]. Following [1], any commuting n-tuple of operators having Γn as a spectral set will be called a Γn- contraction. Many of the fundamental results in the theory of contractions have close parallels for Γ2-contractions as shown in [1]. In this paper we investigate properties of Γn-contarctions and we give a model for Γn-isometries. As an application, we prove a Beurling-Lax-Halmos type theorem characterizing joint invariant subspaces of a pure Γn-isometry. We also indicate how to construct a large class of examples of Γn-contractions. Although a Γ2-contraction can be obtained by symmetrizing any pair of commuting contractions [1], it is no longer true that the symmetrization of any n-tuple of commuting contractions will necessarily give rise to a Γn-contraction, if n > 2 (see Remark 2.12). In fact, the symmetrization of an n-tuple of commuting contractions (T1, . . . , Tn) is a Γn-contraction if and only if (T1, . . . , Tn) satisfies the analogue of von Neumann's inequality for all symmetric polynomials in n variables (see Proposition 2.13). However, it is shown in [1, Examples 1.7 and 2.3] that not all Γ2-contractions are obtained in this way. We usually denote a typical point of Γn by (s1, . . . , sn). We shall also use the notation (S1, . . . , Sn) for an n-tuple of commuting operators associated in some way with Γn. In this paper an operator will always be a bounded linear operator on a Hilbert space. The polynomial ring in n variables over the 2010 Mathematics Subject Classification. 47A13, 47A15, 47A25, 47A45. Key words and phrases. Symmetrized polydisc, spectral set, Beurling-Lax-Halmos theorem, von Neumann inequality. The work of S. Biswas was supported by Inspire Faculty Fellowship funded by DST at Indian Statistical Institute, Kolkata. 1 2 BISWAS AND SHYAM ROY field of complex numbers is denoted by C[z1, . . . , zn]. Consider a commuting n-tuple (S1, . . . , Sn) of operators. We say that Γn is a spectral set for (S1, . . . , Sn), or that (S1, . . . , Sn) is a Γn-contraction, if, for every polynomial p ∈ C[z1, . . . , zn], (1.1) kp(S1, . . . , Sn)k ≤ sup z∈Γn p(z) = kpk∞,Γn . Furthermore, Γn is said to be a complete spectral set for (S1, . . . , Sn), or (S1, . . . , Sn) to be a complete Γn-contraction, if, for every matricial polynomial p in n variables, kp(S1, . . . , Sn)k ≤ sup z∈Γn kp(z)k. Here, if S1, . . . , Sn act on a Hilbert space H and the matricial polynomial p is given by p = [pij] of order m × ℓ, where each pij is a scalar polynomial, then p(S1, . . . , Sn) denotes the operator from Hℓ to Hm with block matrix [pij(S1, . . . , Sn)]. It is a deep result [1, Theorem 1.5] that a Γn-contraction is always a complete Γn-contraction and vice versa, for n = 2. It is not clear whether a similar result is true for n > 2. This will be considered in a future work. We denote the unit circle by T. The distinguished boundary of Γn, denoted by bΓn, defined to be the Silov boundary of the algebra of functions which are continuous on Γn and analytic on the interior of Γn, is s(Tn) [3, Lemma 8]. We shall use some spaces of vector-valued and operator-valued functions. We recall them following [1]. Let E be a separable Hilbert space. We denote by L(E) the space of operators on E, with the operator norm. Let H 2(E) denote the usual Hardy space of analytic E-valued functions on D and L2(E) the Hilbert space of square integrable E-valued functions on T, with their natural inner products. Let H ∞L(E) denote the space of bounded analytic L(E)-valued functions on D and L∞L(E) the space of bounded measurable L(E)-valued functions on T, each with appropriate version of the supremum norm. For ϕ ∈ L∞L(E) we denote by Tϕ the Toeplitz operator with symbol ϕ, given by Tϕf = P+(ϕf ), f ∈ H 2(E), where P+ : L2(E) −→ H 2(E) is the orthogonal projection. In particular Tz is the unilateral shift operator on H 2(E) (the identity function on T will be denoted by z) and T¯z is the backward shift on H 2(E). n The symmetrization map s is a proper holomorphic map, D )) and D ) = Γn is polynomially convex by [10, Theorem 1.6.24]. is polynomially convex. Therefore, s(D Although we have defined Γn-contractions by requiring that the inequality (1.1) holds for all poly- nomials p in C[z1, . . . , zn], this is equivalent to the definition of Γn-contractions by requiring (1.1) to hold for all functions p analytic in a neighbourhood of Γn due to polynomial convexity of Γn as explained in [1]. As discussed in [1], the subtleties surrounding the various notions of joint spectrum and functional calculus for commuting tuples of operators are not relevant to this paper, simply because of the polynomial convexity of Γn. n = s−1(Γn) = s−1(s(D n n 2. Γn and Γn-contractions Note that (s1, . . . , sn) ∈ Γn if and only if all the zeros of the polynomial Pn i=0(−1)n−isn−izi lie in D. This realization of points of Γn will be used repeatedly. We state two theorems about location of zeros of polynomials which will be useful in the sequel. For a polynomial p ∈ C[z], the derivative of p with respect to z will be denoted by p′. Γn-ISOMETRIES 3 Theorem 2.1. (Gauss-Lucas, [6, page. 22]) The zeros of the derivative of a polynomial p lie in the convex hull of the zeros of p. We recall a definition. Definition 2.2. A polynomial p ∈ C[z] of degree d is called self-inversive if zdp( 1 some constant ω ∈ C with ω = 1. ¯z ) = ωp(z) for Theorem 2.3. (Cohn, [6, page, 206]) A necessary and sufficient condition for all the zeros of a polynomial p to lie on the unit circle T is that p is self-inversive and all the zeros of p′ lie in the closed unit disc D. We shall need characterizations of the distinghuished boundary of Γn. Theorem 2.4. Let si ∈ C, i = 1, . . . , n. The following are equivalent: (i) (s1, . . . , sn) is in the distinguished boundary of Γn; (ii) sn = 1, ¯snsi = ¯sn−i and (γ1s1, . . . , γn−1sn−1) ∈ Γn−1, where γi = n−i (iii) (s1, . . . , sn) ∈ Γn and sn = 1. n for i = 1, . . . , n − 1; Proof. Throughout this proof, we put s0 = 1. Let (s1, . . . , sn) be in the distinguished boundary of Γn. By definition, there are λi ∈ T such that (2.1) si = X1≤k1<...<ki≤n λk1 . . . λki for i = 1, . . . , n. This is equivalent to the fact that the polynomial p, given by (2.2) p(z) = nXi=0 (−1)n−isn−izi has all its zeros on T. Moreover, we clearly have sn = 1, ¯snsi = ¯sn−i for i = 1, . . . , n − 1. It follows from Theorem 2.3 that the polynomial (2.3) p′(z) = nXi=1 (−1)n−iisn−izi−1 has all its roots in D, which is equivalent to the fact that (γ1s1, . . . , γn−1sn−1) ∈ Γn−1, where γi = n−i n for i = 1, . . . , n − 1. Therefore (i) implies (ii). Conversely, considering the polynomial in Equation (2.2), we observe that znp(cid:18) 1 ¯z(cid:19) = nXi=0 (−1)i¯sizi and (−1)nsnznp(cid:18) 1 ¯z(cid:19) = p(z) by the first part of (ii). Therefore p is a self-inversive polynomial. Note that all the roots of p′ lies in D as (γ1s1, . . . , γn−1sn−1) ∈ Γn−1, where γi = n−i for i = 1, . . . , n − 1. Thus, it follows n from Theorem 2.3 that p has all its roots on T. This is same as saying that (s1, . . . , sn) is in the distinguished boundary of Γn. Clearly, (i) implies (iii). To see the converse, we note that (iii) implies that there exist λi ∈ D, i = 1, . . . , n, such that (2.1) holds and sn = λ1 . . . λn = 1. So λi ∈ T for all i = 1, . . . , n. It follows from the definition and (2.1) that (s1, . . . , sn) is in the distinguished boundary of Γn. (cid:3) 4 BISWAS AND SHYAM ROY Remark 2.5. From the Lemma above, similar to the part (4) of [2, Theorem 1.3], one can give ex- pressions for sj's, j = 1, . . . , n. Clearly, sn = λ1 . . . λn = eiθ for some θ in R. So λn = eiθ ¯λ1 . . . ¯λn−1. Since λk1 · · · λkj with λj = 1 sj = X1≤k1<k2<...<kj ≤n X1≤k1<k2<...<kj ≤n for j = 1, . . . , n, we have sj = where µj = Since λn = eiθ ¯λ1 . . . ¯λn−1 we obtain λk1 · · · λkj = µj + ¯µj−1λn X 1≤k1<k2<...<kj ≤n−1 λk1 · · · λkj . sj = µj + ¯µj−1eiθ ¯λ1 . . . ¯λn−1 = µj + µj−1 ¯µn−1eiθ = µj + µn−jeiθ, since (µ1, . . . , µn−1) ∈ bΓn−1, ¯µn−1µj−1 = ¯µn−j, j = 1, . . . , n − 1. Lemma 2.6. If (s1, . . . , sn) ∈ Γn, then (γ1s1, . . . , γn−1sn−1) ∈ Γn−1, where γi = n−i n 1, . . . , n − 1. for i = i=0(−1)n−isn−izi, has all its roots in the closed unit disc D. It follows from Theorem 2.1 that the polynomial p′(z) = i=1(−1)n−iisn−izi−1 has all its roots in the closed unit disc as well. Hence we have the desired (cid:3) Proof. The hypothesis is equivalent to the fact that the polynomial p(z) =Pn Pn conclusion. Remark 2.7. Considering the map π : Cn −→ Cn−1 defined by π(z1, . . . , zn) = (γ1z1, . . . , γn−1zn−1) the above Lemma can be restated as π(Γn) ⊆ Γn−1. Lemma 2.8. If (S1, . . . , Sn) is a Γn-contraction, then (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction, where γi = n−i n for i = 1, . . . , n − 1. Proof. For p ∈ C[z1, . . . zn−1], we note that p ◦ π ∈ C[z1, . . . , zn] and by hypothesis kp(γ1S1, . . . , γn−1Sn−1)k = kp ◦ π(S1, . . . , Sn)k ≤ kp ◦ πk∞,Γn = kpk∞,π(Γn) ≤ kpk∞,Γn−1 . This completes the proof. (cid:3) Lemma 2.9. If (s1, . . . , sn) ∈ Γn, then (α + s1, αs1 + s2, . . . , αsn−1 + sn, αsn) ∈ Γn+1 for all α in D. Proof. If (s1, . . . , sn) ∈ Γn, then it follows from the definition of Γn that there are λk ∈ D, k = 1, . . . , n such that λk1 · · · λki , i = 1, . . . , n. If α(:= λn+1) ∈ D, then (s1, . . . , sn+1) ∈ Γn+1, where si = λk1 · · · λki, i = 1, . . . , n + 1. si = X1≤k1<k2<...<ki≤n X 1≤k1<k2<...<ki≤n+1 Γn-ISOMETRIES 5 Putting s0 = 1 and sn+1 = 0, we note that si = αsi−1 + si, i = 1, . . . , n + 1. Therefore we have the desired conclusion. (cid:3) Remark 2.10. For any α ∈ C, considering the one-to-one map πα : Cn −→ Cn+1 defined by πα(z1, . . . , zn) = (α + z1, αz1 + z2, . . . , αzn−1 + zn, αzn), the above Lemma can be restated as πα(Γn) ⊆ Γn+1 for all α ∈ D. Lemma 2.11. Let (S1, . . . , Sn) be a Γn-contraction, then (α + S1, αS1 + S2, . . . , αSn−1 + Sn, αSn) is a Γn+1-contraction for all α in D. Proof. For p ∈ C[z1, . . . , zn+1], we observe that p ◦ πα ∈ C[z1, . . . , zn], so by hypothesis we have kp(α + S1, αS1 + S2, . . . , αSn−1 + Sn, αSn)k = kp ◦ πα(S1, . . . , Sn)k ≤ kp ◦ παk∞,Γn = kpk∞,πα(Γn) ≤ kpk∞,Γn+1. Hence the proof. (cid:3) For an n-tuple T = (T1, . . . , Tn) of commuting operators on a Hilbert space H, let s(T) := (s1(T), . . . , sn(T)), we call s(T), the symmetrization of T. A polynomial p ∈ C[z1, . . . , zn] is called symmetric if p(zσ) := p(zσ(1), . . . , zσ(n)) = p(z) for all z ∈ Cn and σ ∈ Σn, where Σn is the symmetric group on n symbols. If p ∈ C[z1, . . . , zn] is symmetric, then there is a unique q ∈ C[z1, . . . , zn] such that p = q ◦ s, where s is the symmetrization map [4, Theorem 3.3.1]. Remark 2.12. One may be tempted to conjecture that Lemma above is true when α is replaced by a contraction operator T which commutes with all the Si's, i = 1, . . . , n. However, this is no longer true. We take n = 2 and give an example of a Γ2-contraction (S1, S2) and a contraction T such that (T + S1, T S1 + S2, T S2) is not a Γ3-contraction. Let (A1, A2, A3) be a tuple of commuting contractions as in Kaijser-Varopoulos [7, Example 5.7]. We take S1 = A1 + A2, S2 = A1A2 and T = A3. Clearly, (S1, S2) = (A1 + A2, A1A2) is a Γ2-contraction due to Ando's inequality. But (T + S1, T S1 + S2, T S2) = s(A1, A2, A3) is not a Γ3-contraction due to the failure of von Neumann's inequality for more than two commuting contractions. Consider the symmetric polynomial p(z1, z2, z3) = z2 3 − 2z1z2 − 2z2z3 − 2z3z1 1 + z2 2 + z2 in [7, Example 5.7]. Taking q to be the polynomial in 3-variables such that q ◦ s = p, where s is the symmetrization map, one observes that Γ3 cannot be a spectral set for s(A1, A2, A3). Proposition 2.13. The symmetrization of an n-tuple of commuting contractions (T1, . . . , Tn) is a Γn-contraction if and only if (T1, . . . , Tn) satisfies the analogue of von Neumann's inequality for all symmetric polynomials in C[z1, . . . , zn]. Proof. Let T = (T1, . . . , Tn) be a commuting n-tuple of contractions satisfying the analogue of von Neumann's inequality for all symmetric polynomials C[z1, . . . , zn]. We show that s(T) = (s1(T), . . . , sn(T)) is a Γn-contraction. For a polynomial p ∈ C[z1, . . . , zn], we note that kp(s1(T), . . . , sn(T))k = kp ◦ s(T)k ≤ kp ◦ sk∞,D n = kpk∞,s(D n ), 6 BISWAS AND SHYAM ROY since p ◦ s is a symmetric polynomial, the above inequality holds by hypothesis, and it shows that Γn = s(D ) is a spectral set for (s1(T), . . . , sn(T)). n Conversely, let T = (T1, . . . , Tn) be a tuple of commuting contractions such that s(T) is a Γn- contraction and q ∈ C[z1, . . . , zn] be symmetric. So there is a p ∈ C[z1, . . . , zn] such that p ◦ s = q. By hypothesis, we have kq(T)k = kp(s(T))k ≤ kpk∞,Γn = kp ◦ sk∞,D n = kqk∞,D n . (cid:3) Next, we give a straightforward generalization of the Lemma 3.2 in [2]. Proposition 2.14. Let T = (T1, . . . , Tn) be a commuting n-tuple of contractions on a Hilbert space H satisfying the analogue of von Neumann's inequality for all symmetric polynomials in C[z1, . . . , zn]. Let M be an invariant subspace for si(T), i = 1, . . . , n. Then (s1(T)M, . . . , sn(T)M) is a Γn-contraction on the Hilbert space M. Proof. For a polynomial p ∈ C[z1, . . . , zn], we note that kp(s1(T)M, . . . , sn(T)M)k = kp(s1(T), . . . , sn(T))Mk ≤ kp(s1(T), . . . , sn(T))k ≤ kpk∞,Γn, since s(T) = (s1(T), . . . , sn(T)) is a Γn-contraction on H, the last inequality holds by the previous Proposition. Hence we have the desired conclusion. (cid:3) Remark 2.15. 1. It is clear from the proof of the Proposition above that if (S1, . . . , Sn) is a Γn-contraction on a Hilbert space H and M is a common invariant subspace for Si, i = 1, . . . , n, then (S1M, . . . , SnM) is a Γn-contraction on M. 2. As an immediate consequence of Proposition 2.13, we observe that the symmetrizations of the classes of n-tuples of commuting contractions discussed in [5] give rise to a large class of Γn-contractions. 3. It is shown in [2] that applying the Lemma 3.2, a large class of Γ2-contractions can be constructed. In an analogous way, examples of Γn-contractions can be constructed for any integer n > 2 applying Proposition 2.14. 4. From Theorem 3.2 in [1], it is clear that all Γ2-contractions are obtained by applying the Lemma 3.2 in [2] as described in the same paper . However, for n > 3, it is not clear whether all Γn-contractions have similar realizations. We start with the following obvious generalizations of definitions in [1]. 3. Γn-unitary and Γn-isometries Definition 3.1. Let S1, . . . , Sn be commuting operators on a Hilbert space H. We say that (S1, . . . , Sn) is (i) a Γn-unitary if Si, i = 1, . . . , n are normal operators and the joint spectrum σ(S1, . . . , Sn) of (S1, . . . , Sn) is contained in the distinguished boundary of Γn. (ii) a Γn-isometry if there exists a Hilbert space K containing H and a Γn-unitary (eS1, . . . ,eSn) on K such that H is invariant for eSi, i = 1, . . . , n and Si = eSiH, i = 1, . . . , n. Γn-ISOMETRIES 7 (iii) a Γn-co-isometry if (S∗ (iv) a pure Γn-isometry if (S1, . . . , Sn) is a Γn-isometry and Sn is a pure isometry. n) is a Γ-isometry. 1 , . . . , S∗ The proof of the following theorem works along the lines of Agler and Young[1]. Theorem 3.2. Let Si, i = 1, . . . , n, be commuting operators on a Hilbert space H. The following are equivalent: (i) (S1, . . . , Sn) is a Γn-unitary; (ii) S∗ nSn = I = SnS∗ γi = n−i n for i = 1, . . . , n − 1; n, S∗ nSi = S∗ n−i and (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction, where (iii) there exist commuting unitary operators Ui for i = 1, . . . , n on H such that Si = X1≤k1<...<ki≤n Uk1 . . . Uki for i = 1, . . . , n. Proof. Suppose (i) holds. Let (S1, . . . , Sn) be a Γn-unitary. By the spectral theorem for commuting normal operators, there exists a spectral measure M (·) on σ(S1, . . . , Sn) such that where s1, ..., sn are the co-ordinate functions on Cn. Let τ be a measurable right inverse of the restriction of s to Tn, so that τ maps distinguished boundary of Γn to Tn. Let τ = (τ1, . . . , τn) and Si =Zσ(S1,...,Sn) Ui =Zσ(S1,...,Sn) si(z)M (dz), i = 1, . . . , n, τi(z)M (dz), i = 1, . . . , n. X1≤k1<...<ki≤nZσ(S1,...,Sn) = Zσ(S1,...,Sn) si(z)M (dz) = Si, Clearly U1, . . . , Un are commuting unitaries on H and X1≤k1<...<ki≤n Uk1 . . . Uki = τk1(z) . . . τki(z)M (dz) for i = 1, . . . , n. Hence (i) implies (iii). Suppose (iii) holds. Then S∗ n−i, i = 1, . . . , n follow immediately. Moreover, since (S1, . . . , Sn) is a Γn-contraction, we have from Lemma 2.8 that (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction, where γi = n−i nSn = I = SnS∗ nSi = S∗ n and S∗ n for i = 1, . . . , n − 1. Hence (iii) implies (ii). nSiS∗ nSiSn−i = S∗ n−iSn−i = S∗ Suppose (ii) holds. Since Sn is normal, by Fuglede's theorem S∗ i . Now as Si's commute, S∗ i Si and we have each of Si, i = 1, . . . , n is normal. So the unital C ∗-algebra C ∗(S1, . . . , Sn) generated by S1, . . . , Sn is commutative and by Gelfand-Naimark's theorem is ∗-isometrically isomorphic to C(σ(S1, . . . , Sn)). Let S1, . . . , Sn be the images of S1, . . . , Sn under the Gelfand map. By definition, for an arbitrary point z = (s1, , . . . , sn) in σ(S1, . . . , Sn), Si(z) = si for i = 1, . . . , n. By properties of the Gelfand map and hypothesis we have, n−iSn−i = S∗ nSn−iSi = S∗ i Sn = SiS∗ Sn(z) Sn(z) = 1 = Sn(z) Sn(z) and Sn(z) Si(z) = Sn−i(z) for z in σ(S1, . . . , Sn). Thus we obtain sn = 1, ¯snsi = ¯sn−i. Now (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction implies kp(γ1S1, . . . , γn−1Sn−1)k ≤ kpk∞,Γn−1 8 BISWAS AND SHYAM ROY which is equivalent to kpk2 plying Gelfand transform we have ∞,Γn−1 − p(γ1S1, . . . , γn−1Sn−1)∗p(γ1S1, . . . , γn−1Sn−1) is positive. Ap- kpk2 ∞,Γn−1 − p(γ1 S1(z), . . . , γn−1 Sn−1(z))∗p(γ1 S1(z), . . . , γn−1 Sn−1(z)) ≥ 0 for z in σ(S1, . . . , Sn). This shows (γ1s1, . . . , γn−1sn−1) is in the polynomially convex hull of Γn−1. Since Γn−1 is polynomially convex, (γ1s1, . . . , γn−1sn−1) is in Γn−1. Therefore by Theorem 2.4 (s1, . . . , sn) is in the distinguished boundary of Γn and hence σ(S1, . . . , Sn) ⊂ bΓn. This proves (ii) implies (i). (cid:3) It is not a priori clear whether unitarity of Sn would imply that (S1, . . . , Sn) to be Γn-unitary. However, the following is true. sn(T)(=Qn Lemma 3.3. Let T = (T1, . . . , Tn) be a tuple of commuting contractions on a Hilbert space H. If i=1 Ti) is a unitary, then s(T) is a Γn-unitary. Proof. Note that each of the Ti, i = 1, . . . , n is invertible as they are commuting with each other and their product is unitary. Since each Ti is a contraction, we have kT −1 i k ≥ 1. First we show that kT −1 k k > 1. So, there exist a y in H such that kT −1 i k = 1 for all i = 1, . . . , n. If not, then for some k in {1, . . . , n} we have kT −1 i6=k Ti)y = x. Thus k yk > kyk. Let (Qn ksn(T)−1xk = kT −1 k yk kyk · kxk ksn(T)−1k ≥ kyk i6=k Ti)yk > 1 k(Qn i Ti = I, i = 1, . . . , n. Suppose for some ℓ in {1, . . . , n}, T ∗ is a contradiction as by hypothesis sn(T)−1 is also a unitary. Next we show that each Ti is isometry, that is, T ∗ ℓ Tℓ 6= I. There ℓ Tℓ)y 6= 0. Since Tℓ is a contraction, kyk2 − kTℓyk2 = exists nonzero y in H such that (I − T ∗ h(I − T ∗ ℓ k > 1 which is a contradiction. Thus 1 , . . . , T ∗ i Ti = I, i = 1, . . . , n. Applying the same trick to T∗ = (T ∗ T ∗ n ), we see that each Ti is a unitary. Hence by part (iii) of the Theorem above, s(T) is a Γn-unitary. (cid:3) 2 yk2 > 0. But this shows, kT −1 ℓ Tℓ)y, yi = k(I − T ∗ 1 ℓ Tℓ) The following Lemma will be useful in characterizing pure Γn-isometry. Lemma 3.4. Let Φ1, . . . , Φn be functions in L∞L(E) and MΦi , i = 1 . . . , n, denotes the corre- sponding multiplication operator on L2(E). Then (MΦ1 , . . . , MΦn) is a Γn contraction if and only if (Φ1(z), . . . , Φn(z)) is a Γn contraction for all z in T. Proof. Note that for kMΨk = kΨk∞ for Ψ ∈ L∞L(E). By definition, (MΦ1 , . . . , MΦn) is a Γn contraction if and only if (3.1) kp(MΦ1 , . . . , MΦn )k ≤ kpk∞,Γn, for all polynomials p ∈ C[z1, . . . , zn]. Since p(MΦ1 , . . . , MΦn) = Mp(Φ1,...,Φn) and kMΨk = kΨk∞ := supz∈T kΨ(z)k for Ψ ∈ L∞L(E), we have (3.1) is true if and only if kp(Φ1(z), . . . , Φn(z))k ≤ kpk∞,Γn, which completes the proof. (cid:3) Remark 3.5. An interesting case in the Lemma above, is when dim E = 1. In this case, the n-tuple (Φ1(z), . . . , Φn(z)) is a Γn-contraction means that (Φ1(z), . . . , Φn(z)) is in Γn which is true as Γn is polynomially convex. Γn-ISOMETRIES 9 Theorem 3.6. Let Si, i = 1, . . . , n, be commuting operators on a Hilbert space H. Then (S1, . . . , Sn) is a pure Γn-isometry if and only if there exist a separable Hilbert space E and a unitary operator U : H → H 2(E) and functions Φ1, . . . , Φn−1 in H ∞L(E) and operators Ai ∈ L(E) i = 1, . . . , n − 1 such that (i) Si = U ∗MΦi U, i = 1, . . . , n − 1, Sn = U ∗MzU ; (ii) (γ1Φ1(z), . . . , γn−1Φn−1(z)) is a Γn−1-contraction for all z in T, where γi = n−i n 1, . . . , n − 1; for i = (iii) Φi(z) = Ai + A∗ (iv) [Ai, Aj] = 0 and [Ai, A∗ n−iz for 1 ≤ i ≤ n − 1; n−j] = [Aj, A∗ n−i] for 1 ≤ i, j ≤ n − 1, where [P, Q] = P Q − QP for two operators P, Q. Proof. Suppose (S1, . . . , Sn) is a pure Γn-isometry. By definition, there exist a Hilbert space K and a Γn-unitary ( S1, . . . , Sn) such that H ⊂ K is a common invariant subspace of Si's and Si = SiH, i = 1, . . . , n. From Theorem 3.2, it follows that S∗ n−i, 1 ≤ i, j ≤ n − 1. n By compression of Si to H, we have Sn = I and S∗ n Si = S∗ S∗ nSn = I and S∗ nSi = S∗ n−i, 1 ≤ i ≤ n − 1. Since Sn is a pure isometry and H is separable, there exist a unitary operator U : H → H 2(E), for some separable Hilbert space E, such that Sn = U ∗MzU , where Mz is the shift operator on H 2(E). Since Si's commute with Sn, there exist Φi ∈ H ∞L(E) such that Si = U ∗MΦi U for 1 ≤ i ≤ n − 1. As (S1, . . . , Sn) is a Γn-contraction, from Lemma 2.8 and Lemma 3.4 part (ii) follows. The relations S∗ nSi = S∗ n−i yield Let Φ(z) =Pn≥0 Cnzn, Ψ(z) =Pn≥0 Dnzn be in H ∞L(E). Then M¯zMΦ = M ∗ Cnzn−1 = D∗ 0 + D∗ D∗ n ¯zn Ψ implies M¯zMΦi = M ∗ Φn−i , 1 ≤ i ≤ n − 1. C0 ¯z + C1 +Xn≥2 1 ¯z +Xn≥2 for all z ∈ T and by comparing coefficients, we get C0 = D∗ 1 and C1 = D∗ 0. This gives that each Φi(z) is of the form Ai + Biz for some Ai, Bi ∈ L(E), where Ai = B∗ n−i for 1 ≤ i ≤ n − 1, which is part (iii). Now since Si's commutes, we have MΦiMΦj = MΦj MΦi and consequently (Ai + A∗ n−iz)(Aj + A∗ n−jz) = (Aj + A∗ n−jz)(Ai + A∗ n−iz) for 1 ≤ i, j ≤ n − 1 and for all z ∈ T. Comparing the constant term and the coefficients of z, we get part (iv). Conversely, suppose (S1, . . . , Sn) be the n-tuple satisfying conditions (i) to (iv). Consider the n-tuple (MΦ1 , . . . , MΦn−1 , Mz) of multiplication operators on L2(E) with symbols Φ1, . . . , Φn−1, z respectively. From condition (iv), it follows that MΦi's commutes with each other. Part of con- dition (iii) shows that M¯zMΦi = M ∗ by repeating calculations similar to above. Thus, it is easy to see from part (ii) of Theorem 3.2, that (MΦ1 , . . . , MΦn−1 , Mz) is a Γn-unitary and so is (U ∗MΦ1 U, . . . , U ∗MΦn−1 U, U ∗MzU ). Now, Si's, 1 ≤ i ≤ n, are the restrictions to the common invariant subspace H 2(E) of (MΦ1 , . . . , MΦn−1 , Mz) and hence (S1, . . . , Sn) is a Γn-isometry. Since Sn is a shift, (S1, . . . , Sn) is a pure Γn-isometry. (cid:3) Φn−i 10 BISWAS AND SHYAM ROY Next, we obtain characterization for Γn-isometry analogous to the Wold decomposition in terms of Γn-unitaries and pure Γn-isometries using Theorem 3.2 and Theorem 3.6. We prove this along the way of Agler and Young [1] and will use the following Lemma [1, Lemma 2.5]. Lemma 3.7. Let U and V be a unitary and a pure isometry on Hilbert space H1, H2 respectively, and let T : H1 → H2 be an operator such that T U = V T . Then T = 0. Theorem 3.8. Let Si, i = 1, . . . , n, be commuting operators on a Hilbert space H. The following are equivalent: (i) (S1, . . . , Sn) is a Γn-isometry; (ii) There exists a orthogonal decomposition H = H1 ⊕ H2 into common reducing subspaces of Si's, i = 1, . . . , n such that (S1H1 , . . . , SnH1) is a Γn-unitary and (S1H2 , . . . , SnH2) is a pure Γn-isometry; nSn = I, S∗ for i = 1, . . . , n − 1. n−i and (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction, where γi = n−i n nSi = S∗ (iii) S∗ Proof. Suppose (i) holds. By definition, there exists ( S1, . . . , Sn) a Γn-unitary on K containing H such that H is a invariant subspace of Si's and Si's are restrictions of Si's to H. From Theorem 3.2, it follows that Si's are satisfying the relations: S∗ n−i and (γ1 S1, . . . , γn−1 Sn−1) n is a Γn−1-contraction, where γi = n−i n for i = 1, . . . , n − 1. Compressing to the common invariant subspace H and by part (1) of remark 2.15, we obtain Sn = I, S∗ n Si = S∗ nSn = I, S∗ S∗ nSi = S∗ n−i and (γ1S1, . . . , γn−1Sn−1) is a Γn−1-contraction, where γi = n−i n implies (iii). for i = 1, . . . , n − 1. Thus (i) Suppose (iii) holds. By Wold decomposition, we may write Sn = U ⊕ V on H = H1 ⊕ H2 where H1, H2 are reducing subspaces for Sn, U is unitary and V is pure isometry. Let us write Si ="S(i) 11 S(i) S(i) 21 S(i) 22# 12 with respect to this decomposition, where S(i) SnSi = SiSn, we have jk is a bounded operator from Hk to Hj. Since "U S(i) V S(i) 11 U S(i) 21 V S(i) 22# ="S(i) 11 U S(i) 12 V S(i) 21 U S(i) 22 V# , i = 1, . . . , n − 1. 12 Thus, S(i) 21 U = V S(i) 21 and hence by Lemma 3.7, S(i) 21 = 0, i = 1, . . . , n − 1. Now S∗ nSi = S∗ n−i gives "U ∗S(i) 0 11 U ∗S(i) 12 V ∗S(i) 22# ="S(n−i)∗ 11 S(n−i)∗ 12 # , i = 1, . . . , n − 1. 0 S(n−i)∗ 22 It follows that S(i) From the matrix equation above, we have U ∗S(i) the remark 2.15 and part (ii) of Theorem 3.2, it follows that (S(1) 12 = 0, i = 1, . . . , n − 1. So H1, H2 are common reducing subspace for S1, . . . , Sn. , i = 1, . . . , n − 1. Thus by part (i) of , U ) is a Γn-unitary. , V ) is a pure Γn-isometry. Since V is a pure isometry H is separable, we can identify it with the shift operator Mz on the space of vector valued functions H 2(E), for some separable Hilbert space E. Since S(i) 22 's commute with V , there exists We now require to show that (S(1) 11 , . . . , S(n−1) 22 , . . . , S(n−1) 11 = S(n−i)∗ 11 11 22 Γn-ISOMETRIES 11 Φi ∈ H ∞L(E) such that S(i) a Γn-contraction, from Lemma 2.8 and Lemma 3.4, it follows that (γ1S(1) n for i = 1, . . . , n − 1. The relations V ∗S(i) Γn−1-contraction, where γi = n−i M¯zMΦi = M ∗ 22 can be identified with MΦi for 1 ≤ i ≤ n − 1. As (S1, . . . , Sn) is ) is a , 1 ≤ i ≤ n − 1. 22 22 , . . . , γn−1S(n−1) 22 = S(n−i)∗ yield 22 Φn−i Calculations similar to the first part of the proof of Theorem 3.6, we get each Φi(z) is of the form Ai + A∗ n−iz for some Ai's in L(E), i = 1, . . . , n − 1. Now since S(i) 22 's commutes, we have [Ai, Aj] = 0 and [Ai, A∗ n−j] = [Aj, A∗ n−i] for 1 ≤ i, j ≤ n − 1. Hence, by Theorem 3.6, (S1H2 , . . . , SnH2) is a pure Γn-isometry. Thus (iii) implies (ii). It is easy to see that (ii) implies (i). (cid:3) Corollary 3.9. Let Si, i = 1, . . . , n, be commuting operators on a Hilbert space H. (S1, . . . , Sn) is a Γn-co-isometry if and only if SnS∗ n = I, SnS∗ i = Sn−i and (γ1S∗ 1 , . . . , γn−1S∗ n−1) is a Γn−1-contraction, where γi = n−i n for i = 1, . . . , n − 1. 4. Characterization of invariants subspaces for Γn-isometries This section is devoted to characterize the joint invariant subspaces of pure Γn-isometries. Similar characterizations for pure Γ2-isometries appears in [9]. In the light of Theorem 3.6, we start with a characterization of pure Γn-isometries in terms of the parameters associated with them. For simplicity of notation, let (MΦ, Mz) denote the n-tuple of multiplication operators (MΦ1 , . . . , MΦn−1 , Mz) on H 2(E), where Φi ∈ H ∞L(E), i = 1, . . . , n − 1. Throughout this section, we will be using the canonical identification of H 2(E) with H 2 ⊗ E by the map znξ 7→ zn ⊗ ξ, where n ∈ N ∪ {0} and ξ ∈ E, whenever necessary. n−iz and Φi(z) = Ai + A∗ n−iz be in H ∞L(E) for some Ai ∈ L(E) Theorem 4.1. Let Φi(z) = Ai + A∗ and Ai ∈ L(F) respectively, i = 1, . . . , n − 1. Then the n-tuple (MΦ, Mz) on H 2(E) is unitarily equivalent to the n-tuple (M Φ, Mz) on H 2(F) if and only if the (n − 1)-tuples (A1, . . . , An−1) and ( A1, . . . , An−1) are unitarily equivalent. Proof. Suppose the n-tuple (MΦ, Mz) on H 2(E) is unitarily equivalent to the n-tuple (M Φ, Mz) on H 2(F). We can identify the map Φi (similarly Φi) by IH 2 ⊗ Ai + Mz ⊗ A∗ n−i. So, there exist a unitary U : H 2 ⊗ E → H 2 ⊗ F such that U (IH 2 ⊗ Ai + Mz ⊗ A∗ n−i)U ∗ = IH 2 ⊗ Ai + Mz ⊗ A∗ n−i, i = 1, . . . , n − 1, (4.1) and (4.2) From equation (4.2), it follows that there exists a unitary U : E → F such that U = IH 2 ⊗ U . Consequently, the equation (4.1) can be written as U (Mz ⊗ IE)U ∗ = Mz ⊗ IF. U Ai U ∗ + U A∗ n−iz U ∗z = Ai + A∗ n−iz, i = 1, . . . , n − 1, for all z ∈ T. Hence comparing the coefficients, we obtain U Ai U ∗ = Ai, i = 1, . . . , n − 1 which completes the proof in forward direction. Conversely, suppose there exist a unitary U : E → F that intertwines Ai and Ai, that is, U Ai U ∗ = Ai for each i = 1, . . . , n − 1. Let U : H 2 ⊗ E → H 2 ⊗ F be the map defined by 12 BISWAS AND SHYAM ROY U = IH 2 ⊗ U . Clearly, U is a unitary and from the computations similar to above, it is easy to see that U intertwines MΦi with M Φi for each i =, 1 . . . , n − 1, and Mz. This completes the proof. (cid:3) In the following corollary we express the above theorem in terms of pure Γn-isometries. Corollary 4.2. Let (S1, . . . , Sn) and ( S1, . . . , Sn) be a pair of pure Γn-isometries. Then (S1, . . . , Sn) and ( S1, . . . , Sn) are unitarily equivalent if and only if the (n − 1)-tuples (S∗ i=1 and ( S∗ n−i − SiS∗ i=1 are unitary equivalent. n−i − Si S∗ n)n−1 n)n−1 Proof. Let the pure Γn-isometry (S1, . . . , Sn) is equivalent to (MΦ, Mz) where Φi(z) = Ai + A∗ n−iz be in H ∞L(E) for some Ai ∈ L(E), i = 1, . . . , n − 1 satisfying conditions (ii) to (iv) in Theorem 3.2. Clearly the (n − 1)-tuple (S∗ i=1 . Note that with the canonical identification we have i=1 is equivalent to (M ∗ − MΦiM¯z)n−1 n−i − SiS∗ n)n−1 Φn−i M ∗ Φn−i − MΦiM¯z = (IH 2 ⊗ An−i + Mz ⊗ A∗ i )∗ − (IH 2 ⊗ Ai + Mz ⊗ A∗ n−i)(M¯z ⊗ IE) = (IH 2 − MzM¯z) ⊗ A∗ = PC ⊗ A∗ n−i n−i where PC is the orthogonal projection from H 2 to the scalars in H 2. Thus it follows for Theorem 4.1 that unitary equivalence of the n-tuple (MΦ, Mz) is determined by unitary equivalence of the (n − 1)-tuple (M ∗ i=1 , the unitary equivalence of (S1, . . . , Sn) is determined by (S∗ (cid:3) − MΦiM¯z)n−1 i=1 . This completes the proof. n−i − SiS∗ n)n−1 Φn−i Characterization of invariant subspaces of Γn-isometries essentially boils down to that of pure Γn-isometries due to Wold type decomposition in Theorem 3.8 which is again same as characterizing invariant subspace for the associated model space obtained in Theorem 3.6. The following theorem discusses this issue. A closed subspace M 6= {0} of H 2(E) is said to be (MΦ, Mz)-invariant if M is invariant under MΦi and Mz for all i = 1, . . . , n − 1. Let M 6= {0} be a closed subspace of H 2(E∗). It follows from Beurling-Lax-Halmos theorem that M is invariant under Mz if and only if there exists a Hilbert space E and an inner function Θ ∈ H ∞L(E, E∗) (Θ is an isometry almost everywhere on T) such that M = MΘH 2(E). Theorem 4.3. Let M 6= 0 be a closed subspace of H 2(E∗) and Φi, i = 1, . . . , n − 1 be in H ∞L(E∗) such that (MΦ, Mz) is a pure Γn-isometry on H 2(E∗). Then M is a (MΦ, Mz)-invariant subspace if and only if there exist Ψi, i = 1, . . . , n − 1 in H ∞L(E) such that (MΨ, Mz) is a pure Γn-isometry on H 2(E) and where Θ ∈ H ∞L(E, E∗) is the Beurling-Lax-Halmos representation of M. ΦiΘ = ΘΨi, i = 1, . . . , n − 1, Proof. We will prove only the forward direction as the other part is easy to see. Let M is invariant under (MΦ, Mz). Thus, in particular, M is invariant under Mz and hence the Beurling -Lax- Halmos represntation of M is ΘH 2(E) where ΘH ∞L(E, E∗) is an inner multiplier. We also have MΦiΘH 2(E) ⊆ ΘH 2(E) for each i = 1, . . . , n − 1. Thus, there exist Ψi's, i = 1, . . . , n − 1 in H ∞L(E) such that MΦiMΘ = MΘMΨi , that is, ΦiΘ = ΘΨi, i = 1, . . . , n − 1. Since Φi's commute, we Γn-ISOMETRIES 13 have MΘMΨiMΨj = MΦiMΦj MΘ = MΦj MΦiMΘ = MΘMΨj MΨi and hence ΨiΨj = ΨjΨi, i, j = 1, . . . , n − 1. Furthermore, we have p(MΦ1 , . . . , MΦn−1 )MΘ = MΘp(MΨ1 , . . . , MΨn−1 ) for all polynomials p ∈ C[z1, . . . , zn−1]. Therefore, kp(γ1MΨ1 , . . . , γn−1MΨn−1)k ≤ kM ∗ Θp(γ1MΦ1 , . . . , γn−1MΦn−1 )MΘk ≤ kpk∞,Γn−1, for all polynomials p ∈ C[z1, . . . , zn−1] and γi = n−i also note that n for i = 1, . . . , n − 1. Using Theorem 3.6, we M¯zMΨi = M¯zM ∗ ΘMΦiMΘ = M ∗ ΘM¯zMΦiMΘ = M ∗ ΘM ∗ Φn−i MΘ = M ∗ Ψn−i , i = 1, . . . , n − 1. From the observations made above and by Theorem 3.6, it follows that (MΨ, Mz) is a pure Γn- isometry on H 2(E) and this completes the proof. (cid:3) References [1] J. Agler and N. J. Young, A model theory for Γ-contractions, J. Operator Theory 49 (2003), 45-60. [2] T. Bhattacharyya, S. Pal and S. Shyam Roy, Dilations of Γ-contractions by solving operator equations, Adv. Math. 230 (2012), no. 2, 577-606. [3] A. Edigarian and W. Zwonek, Geometry of the symmetrized polydisc, Arch. Math., 84 (2005), 364-374. [4] J. P. Escofier, Galois theory, Translated from the 1997 French original by Leila Schneps. Graduate Texts in Mathematics, 204. Springer-Verlag, New York, 2001. xiv+280 pp. [5] A. Grinshpan, D. S. Kaliuzhnyi-Verbovetskyi, V. Vinnikov and H. Woerdeman, Classes of tuples of commuting contractions satisfying the multivariable von Neumann inequality, J. Funct. Anal. 256 (2009), no. 9, 3035-3054. [6] M. Marden, Geometry of polynomials, Second edition. Mathematical Surveys, No. 3 American Mathematical Society, Providence, R.I. 1966 xiii+243 pp. [7] V. Paulsen, Completely bounded maps and operator algebras, Cambridge Studies in Advanced Mathematics, 78. Cambridge University Press, Cambridge, 2002. xii+300 pp. [8] W. Rudin, Proper holomorphic mappings maps and finite reflection groups, Indiana Univ. Math. J., 31 (1982), 701-720 [9] J. Sarkar, Operator Theory on Symmetrized Bidisc, arXiv:1207.1862. [10] E. L. Stout, Polynomial convexity, Progress in Mathematics, 261. BirkhÃďuser Boston, Inc., Boston, MA, 2007. xii+439 pp. (Biswas) Theoretical Statistics and Mathematics Unit, Indian Statistical Institute, Kolkata 700108, India E-mail address: [email protected] (Shyam Roy) Indian Institute of Science Education and Research, Pincode 741252, Nadia, West Bengal, India E-mail address: [email protected]
1912.08105
1
1912
2019-12-17T15:58:47
The unbounded extension of Hille-Phillips functional calculus
[ "math.FA" ]
The extension of Hille-Phillips functional calculus of semigroup generators which leads to unbounded operators is given. Connections of this calculus to Bochner-Phillips functional calculus are indicated, and several examples are considered.
math.FA
math
The unbounded extension of Hille-Phillips functional calculus A. R. Mirotin [email protected] The extension of Hille-Phillips functional calculus of semigroup generators which leads to unbounded operators is given. Connections of this calculus to Bochner-Phillips functional calculus are indicated, and several examples are considered. Key wards: Hille-Phillips functional calculus, Bochner-Phillips functional calculus, fractional powers of operators, subordination. РАСШИРЕНИЕ ФУНКЦИОНАЛЬНОГО ИСЧИСЛЕНИЯ ХИЛЛЕ-ФИЛЛИПСА, ПРИВОДЯЩЕЕ К НЕОГРАНИЧЕННЫМ ОПЕРАТОРАМ А. Р. Миротин Дается расширение функционального исчисления Хилле-Филлипса генераторов C0-полугрупп, изложенного в их известной монографии, приводящее к неограничен- ным операторам. Указаны связи этого исчисления с исчислением Бохнера-Филлипса. Рассмотрены примеры. Ключевые слова: функциональное исчисления Хилле-Филлипса, функциональное исчисления Бохнера-Филлипса, дробные степени операторов, подчиненная полугруп- па. 1. Введение. В монографии [1], гл. XV -- XVI построено функциональное исчис- ление генераторов C0-полугрупп, использующее класс LM функций, представимых в виде преобразований Лапласа La(s) :=Z ∞ 0 estda(t) (s < 0) σ-конечных комплексных регулярных борелевских мер a на R+. Пространство таких мер будет обозначаться M(R+). При этом условия, налагаемые в [1] на меру и полу- группу, приводят к тому, что возникающие в результате операторы ограничены. Там же (с. 463) поставлена задача построения расширения этого исчисления, приводяще- го к неограниченным операторам. В работе развивается подход к такому расшире- нию, анонсированный ранее в [2]. Случай генераторов групп рассматривался в [3]. Подход, основанный на другой идее (причем для наборов нескольких генераторов), появился в [4]. Необходимость исчисления подобного типа вызвана также потреб- ностями функционального исчисления Бохнера-Филлипса, использующего класс T отрицательных функций Бернштейна [8] -- [15] (см. также [16]). Ниже будут установ- лены связи между этими исчислениями. Всюду ниже A есть генератор ограниченной 1 C0-полугруппы T в банаховом пространстве X с областью определения D(A) и обра- зом ImA. Через LB(Y, X) обозначается пространство линейных ограниченных опера- торов, действующих между банаховыми пространствами Y и X, LB(X) := LB(X, X). Если f есть функция на R+, то через bf будет обозначаться преобразование Лапласа меры f (t)dt. Конец доказательства или примера обозначается знаком (cid:3). 2. Основное определение. Следующее определение формально совпадает с определением, предложенным Хилле и Филлипсом в монографии [1], но мы отказы- ваемся от наложенных там ограничений, гарантирующих существование интеграла и ограниченность определяемого им оператора (см. также работу [3], посвященную генераторам групп). Определение 1. Для функции g из LM, g = La, a ∈ M(R+)) положим g(A)x = ∞Z 0 T (t)xda(t), где область определения D0(g(A)) этого оператора состоит из тех x ∈ X, для которых интеграл в правой части существует в смысле Бохнера. Следующий пример иллюстрирует определение 1. Пример 1. Пусть g(s) = s−1, s < 0. Тогда g = La, где a = −mes (mes (cid:22) мера Лебега на R+). Таким образом, в силу определения 1 при x ∈ D0(g(A)) g(A)x = − ∞Z 0 T (t)xdt. Предположим, что генератор A инъективен, а полугруппа T сильно устойчива (т. е. T (n)y → 0 при y ∈ X, n → ∞) и покажем, что из определения 1 следует равенство g(A) = A−1. Пусть x ∈ ImA и y = A−1x. Тогда g(A)x = − ∞Z 0 T (t)Aydt = − ∞Z 0 dT (t)y = y. Таким образом, D0(g(A)) ⊇ ImA и при x ∈ ImA имеем g(A)x = A−1x. Нам оста- лось доказать включение D0(g(A)) ⊆ ImA. С этой целью выберем произвольно x ∈ D0(g(A)) и рассмотрим последовательность yn := − nZ 0 T (t)xdt. Положим y := lim n→∞ yn. Как известно, yn ∈ D(A) и Ayn = −A nZ 0 T (t)xdt = x − T (n)x. 2 Поэтому Ayn → x (n → ∞), и в силу замкнутости оператора A имеем y ∈ D(A) и Ay = x, что и завершает доказательство.(cid:3) 3. Теоремы о замкнутости g(A). Прежде всего, нас интересуют условия, при которых оператор g(A) будет плотно определен и замкнут (замыкаем). Лемма 1. 1) Если ImA ⊂ D0(g(A)), то оператор g(A)A ограничен относительно A; 2) если дополнительно предположить, что оператор A инъективен, то g(A) замкнут на подпространстве D(A) ∩ D0(g(A)), наделенном нормой графика; 3) если R ∞ 0 kT (t)k d a (t) < ∞, то оператор g(A) ограничен на X, и рассматри- ваемое исчисление согласовано с классическим исчислением Хилле-Филлипса. Доказательство. 1) Для любого x ∈ D(A) определим операторы Bnx =Z n 0 T (t)Axda(t), (интеграл существует в смысле Бохнера, так как функция r 7→ kT (t)k ограничена на [0,n] по принципу равномерной ограниченности). Поскольку у нас Ax ∈ D0(g(A)), то Bnx → g(A)Ax (n → ∞) при всех x ∈ D(A). Далее, так как оператор A замкнут, то пространство Y = D(A), наделенное нормой графика kxkY = kxk + kAxk, банахово. Кроме того, Bn ∈ LB(Y, X), поскольку kBnxk ≤(cid:18)Z n 0 kT (t)k da(t)(cid:19) kxkY . В силу теоремы Банаха-Штейнгауза оператор g(A)A тоже принадлежит LB(Y, X), а потому ограничен относительно A. 2) Заметим сначала, что при x ∈ D(A) ∩ D0(g(A)) справедливо равенство Ag(A)x = g(A)Ax. (1) Действительно, с учетом замкнутости A и сходимости интегралов имеем g(A)Ax =Z ∞ 0 T (t)Axda(t) =Z ∞ 0 AT (t)xda(r) = Ag(A)x, поскольку ImA ⊂ D0(g(A)). Теперь, если A инъективен, то с помощью утверждения 1) получаем, что оператор g(A)x = A−1g(A)Ax замнут на подпространстве D(A) ∩ D0(g(A)) пространства Y как произведение замкнутого и ограниченного операторов. 3) Это следует из свойств интеграла Бохнера. (cid:3) Всюду далее для функции f на R+ через bf обозначается ее преобразование Ла- пласа, т. е. bf (s) :=Z ∞ 0 estf (t)dt (s < 0). 3 Следствие 1. Пусть g = bf есть преобразование Лапласа функции f, причем 0 T (t)xdf (t). Тогда ImA ⊂ D0(g(A)) при x из D(A) существуют lim n→∞ и справедливы все утверждения леммы 1. f (n)T (n)x и R ∞ Доказательство. Интегрируя по частям, получаем при всех x ∈ D(A) Z n 0 T (t)Axf (t)dt =Z n 0 f (t)dT (t)x = f (n)T (n)x − f (0)x −Z n 0 T (t)xdf (t), (2) причем правая часть имеет предел при n → ∞. (cid:3) Следствие 2. В условиях части 1 леммы 1 оператор g(A)A ограничен вместе с A. Следствие 3. Если в условиях части 1 леммы 1 оператор А инъективен, то g(A)ImA ограничен относительно A−1. Следствие 4. Если в условиях части 1 леммы 1 существует ограниченный обратный оператор A−1 на ImA, то g(A) ограничен на ImA. Теорема 1. Пусть g = bf , где функция f такова, что оператор Sx := T (t)xdf (t) ∞Z 0 ограничен, а последовательность f (n)T (n) сходится на D(A) сильно к оператору B ∈ LB(D(A)). Тогда 1) ImA ⊂ D0(g(A)) и оператор g(A)A ограничен на D(A); 2) если оператор A инъективен, то оператор g(A)D(A) ∩ D0(g(A)) замкнут, а если еще B = 0, то замкнут также и оператор g(A)Im(A). Доказательство. 1) Включение ImA ⊂ D0(g(A)) сразу вытекает из следствия 1. Переходя к пределу в формуле (2), получаем при x ∈ D(A) g(A)Ax =Z ∞ 0 T (t)Axf (t)dt = Bx − f (0)x −Z ∞ 0 T (t)xdf (t), (3) откуда и следует ограниченность g(A)A на D(A). 2) Здесь первое утверждение следует из 1) и равенства g(A)x = A−1g(A)Ax (x ∈ D(A) ∩ D0(g(A))) как в доказательстве леммы 1. Пусть теперь B = 0. Полагая в (3) y = Ax, имеем g(A)y = −f (0)A−1y −Z ∞ 0 T (t)A−1ydf (t). Но оператор y 7→ AZ ∞ 0 T (t)A−1ydf (t) =Z ∞ 0 AT (t)A−1ydf (t) =Z ∞ 0 T (t)ydf (t) = Sy 4 ограничен на ImA, а потому оператор g(A)y = −A−1 (f (0)y + Sy) замкнут на ImA как произведение замкнутого и ограниченного операторов. (cid:3) Отрицательные дробные степени генераторов полугрупп определяются, как при- вило, при условиях, гарантирующих ограниченность этих степеней (см., напр., [5, c. 32 -- 33]). В [6, глава 7] отрицательные дробные степени определялись при условиях позитивности и инъективности оператора. Наш подход позволяет снять или ослабить эти ограничения. Пример 2. Пусть g(s) = (−s)−α, s < 0, α > 0. Тогда g = bf , где f (r) = 1/Γ(α)rα−1. Следовательно, мы можем в соответствии с определением 1 положить (−A)−αx := 1 Γ(α) ∞Z 0 T (t)xtα−1dt, считая, что D0((−A)−α) состоит из тех x ∈ X, при которых интеграл существует в смысле Бохнера. Предположим, что α > 1, оператор A инъективен, а C0-полугруппа T удовлетворяет оценке kT (t)k ≤ C/tδ с константами δ > α − 1, C > 0. Тогда выпол- нены все условия теоремы 1, а потому ImA ⊂ D0((−A)−α) и оператор (−A)−αImA замкнут. Если же 0 < α < 1, то мы можем при δ > α положить (−A)−α := (−A)−(1+α)(−A). В этом случае D((−A)−α) = D(A), и при x ∈ D(A), интегрируя по частям, для (−A)−αx получим ту же формулу, что и выше. В самом деле, тогда tαT (t)x → 0 (t → ∞), а потому (−A)−αx := 1 Γ(α + 1) ∞Z 0 T (t)(−Ax)tαdt = − 1 αΓ(α)  tαT (t)x ∞ 0 − α ∞Z 0 T (t)xtα−1dt  = 1 Γ(α) ∞Z 0 T (t)xtα−1dt.(cid:3) 4. Связь с исчислением Бохнера-Филлипса. Следующие теоремы устанавли- вают связь между рассматриваемым исчислением и исчислением Бохнера-Филлипса (см., например, [9], [11]). Ниже через T будет обозначаться класс неположительных функций Бернштейна одного переменного. Функция ψ из T допускает интегральное представление ψ(s) = ψ(0) +Z ∞ где ρ ≥ 0 (cid:22) мера на R+, причем R r 0 0 dρ(u) < ∞, R ∞ (esu − 1) u−1dρ(u) (s < 0), (4) r u−1dρ(u) < ∞ при r > 0. Определение 2. Для неположительной функции Бернштейна ψ с интегральным представлением (4) и генератора A ограниченной C0-полугруппы T на банаховом 5 пространстве X ее значение на операторе A при x ∈ D(A)определяется интегралом Бохнера ψ(A)x = c0x +Z R+ (T (u) − I)xu−1dρ(u). Замыкание этого оператора, также обозначаемое ψ(A), тоже есть генератор ограни- ченной C0-полугруппы gt(A) на X (эта полугруппа называется подчиненной полу- группе T ). Теорема 2. Пусть ψ ∈ T , ψ(0) = 0. Тогда функция ψ(s) := ψ(s)/s принадлежит LM и при всех x ∈ D(A) ∩ D0( ψ(A)) справедливо равенство ψ(A)x = A ψ(A)x, (5) где ψ(A) понимается в смысле определения 1, а ψ(A) (cid:22) в смысле исчисления Бохнера- Филлипса. Доказательство. В силу формулы (4) и теоремы Фубини ψ(s) = ∞R0 (cid:0) esu−1 s (cid:1) u−1dρ(u) = esr(cid:18) ∞R0 ∞R0 ∞R0 (cid:18) ∞R0 1[0;u](r)esrdr(cid:19) u−1dρ(u) = 1[0;u](r)u−1dρ(u)(cid:19) dr = bf (s), где f (r) =R ∞ r u−1dρ(u), а 1A (cid:22) индикатор множества A. Следовательно, если x ∈ D(A) ∩ D0( ψ(A)), то ψ(A)x = ∞Z 0 T (t)xf (t)dt. (6) С другой стороны, по теореме Фубини для интеграла Бохнера ψ(A)x := ∞R0 ∞R0 (cid:18) ∞R0 (T (u) − I) xu−1dρ(u) = ∞R0 (cid:18) uR0 1[0;u](t)AT (t)xdt(cid:19) u−1dρ(u) = A AT (t)xdt(cid:19) u−1dρ(u) = ∞R0 T (t)x(cid:18) ∞R0 1[0;u](t)u−1dρ(u)(cid:19) dt = A ψ(A)x (теорема Фубини применима, поскольку интеграл Бохнера в (6) сходится). (cid:3) Следствие 5. Если оператор A имеет ограниченный обратный, то оператор ψ(A) замкнут на подпространстве D(A) ∩ D0( ψ(A)). Доказательство. В силу теоремы 4.1 из [10] и ее следствия формула (5) влечет равенство ψ(A)x = ψ(A)A−1x (x ∈ D(A) ∩ D0( ψ(A))), правая часть которого есть произведение замкнутого и ограниченного операторов. (cid:3) 6 Следствие 6. Оператор ψ(A) отображает D(A) ∩ D0( ψ(A)) в D(A). Следствие 7. Оператор A отображает D(A) ∩ D0( ψ(A)) в D0( ψ(A)). Положим LMT :=ng ∈ LM : g = La, ∀x ∈ X lim t→∞ a(t)kT (t)xk = 0o . Теорема 3 (ср. [11, теорема 1]). Пусть g ∈ LMT , ψ ∈ T . Тогда h := gψ ∈ LMT , D0(h(A)) ⊂ D(A) ∩ D0(g(A)) и при x ∈ D(A) ∩ D0(g(A)) справедливы равенства h(A)x = ψ(A)g(A)x = g(A)ψ(A)x. Доказательство. Пусть g = La, a(t) (cid:22) такая функция распределения меры a, что a(0) = 0. Тогда h = Lb, где b (cid:22) мера на R+ с функцией распределения b(t) = ψ(0)a(t) + ∞Z 0 (a(t − u) − a(t)) u−1dρ(u). (7) В самом деле, в силу [7, глава 1, §15, теорема 15 b] Lb(s) = ψ(0)La(s) + ∞Z 0 ψ(0)g(s) + u−1dρ(u) (a(t − u) − a(t))u−1dρ(u) estdt ∞Z   = ∞Z estdt (a(t − u) − a(t)) = 0 0 0 ∞Z  ∞Z  ∞Z 0 0 ψ(0)g(s) + ∞Z 0 estdta(t − u) − estda(t) ∞Z 0  u−1dρ(u) = ψ(0)g(s) + ∞Z 0 (cid:0)es(t+u) − est(cid:1) da(t)u−1dρ(u) = ψ(0)g(s) + estda(t) (eus − 1) u−1dρ(u) = g(s)ψ(s). ∞Z ∞Z 0 Далее, поскольку при x ∈ X, t > 0 0 b(t)T (t)x = ψ(0)a(t)T (t)x + ∞Z 0 (a(t − u) − a(t)) T (t)xu−1dρ(u) = 7 ψ(0)a(t)T (t)x + ∞Z 0 a(t)(T (t + u)x − T (t)x)u−1dρ(u) = a(t)T (t)ψ(A)x, то h ∈ LMT . Кроме того, полагая t → +0, получаем b(+0) = 0. С помощью интегрирования по частям легко проверить, что при x ∈ D(A) ∩ D0(g(A)) g(A)x =Z ∞ 0 T (t)(−Ax)a(t)dt. (8) Следовательно, при этих x справедливы равенства (T (u) − I)g(A)x = T (u + t)(−Ax)a(t)dt − T (t)(−Ax)a(t)dt = ∞R0 = T (t)(−Ax)(a(t − u) − a(t))dt. ∞R0 ∞R0 Поэтому и с учетом (7) при x ∈ D(A)∩D0(g(A)) имеем (не нарушая общности, можно считать ψ(0) = 0) ∞R0 ∞R0 (T (u) − I)) g(A)xu−1dρ(u) = T (t)(−Ax)(cid:18) ∞R0 (a(t − u) − a(t))u−1dρ(u)(cid:19) dt = ∞R0 ψ(A)g(A)x = T (t)(−Ax)(a(t − u) − a(t))dt(cid:19) u−1dρ(u) = ∞R0 (cid:18) ∞R0 T (t)(−Ax)b(t)dt = − b(t)dT (t)x = ∞R0 −b(t)T (t)x∞ t=0 + ∞Z 0 T (t)xdb(t) = h(A)x. Значит, D0(h(A)) ⊂ D(A) ∩ D0(g(A)) и первое равенство доказано. Наконец заметим, что операторы T (u) и g(A) коммутируют в том смысле, что при x ∈ D0(g(A)) T (u)g(A)x = ∞Z 0 T (u + t)xda(t) = ∞Z 0 T (t)T (u)xda(t) = g(A)T (u)x, а потому h(A)x = ψ(A)g(A)x = ∞R0 g(A) (T (u) − I) xu−1dρ(u) = = g(A) (T (u) − I) xu−1dρ(u) = g(A)ψ(A)x. (cid:3) ∞R0 Пример 3. Пусть g(s) = (−s)−α, s < 0, 0 < α < 1, а C0-полугруппа T удовле- творяет оценке kT (t)k ≤ C/tδ с константами δ > α, C > 0 как в примере 2. И пусть ψ(s) = −(−s)β, s < 0, 0 < β < α. Как известно, ψ ∈ T и −(−s)β = β Γ(1 − β) 8 ∞Z 0 (est − 1)t−β−1dt. Тогда выполнены все условия теоремы 3, причем в силу примера 2 D0(g(A)) = D0(h(A)) = D(A), а потому для генератора A полугруппы T при x ∈ D(A) спра- ведливо равенство (−A)β(−A)−αx = (−A)β−αx.(cid:3) В связи со идущим ниже следствием 8 отметим, что, если ψ ∈ T , то функция −1/ψ на (−∞, 0) абсолютно монотонна, а потому принадлежит LM. Следствие 8. Пусть ψ ∈ T , функция 1/ψ принадлежит LMT , и оператор (1/ψ)(A) (в смысле определения 1) определен и ограничен на X. Тогда оператор ψ(A) обратим, и ψ(A)−1 = (1/ψ)(A). Доказательство. Так как (1/ψ)(s)ψ(s) = 1, то силу теоремы 3 при x ∈ D(A) имеем (1/ψ)(A)ψ(A)x = ψ(A)(1/ψ)(A)x = x. Поскольку оператор ψ(A)(1/ψ)(A) замкнут как произведение замкнутого и ограни- ченного операторов, последнее равенство верно при всех x ∈ X. Следовательно, опе- ратор ψ(A) биективен и ψ(A)−1 = (1/ψ)(A). (cid:3) Пример 4. Пусть ψ(s) = − log(1 −s). Известно, что ψ ∈ T (см., напр., [9]). Кроме того, известно, что для функции ν(t, −1) = ∞Z 0 tξ−1 Γ(ξ) dξ. справедливо равенство 1/ log(−x) = \ν(t, −1)(x) при x < 0 [17, глава V, §5.7, (11)]. Следовательно, 1/ψ = [(−f ), где f (t) = e−tν(t, −1). Пусть полугруппа T равномерно устойчива, т. е. kT (t)k ≤ Meωt, где ω < 0. С помощью правила Лопиталя легко проверить, что 1/ψ ∈ LMT . Кроме того, оператор (1/ψ)(A) определен и ограничен на X, так как k(1/ψ)(A)xk ≤ kT (t)xkf (t)dt ≤ M log(1 − ω) kxk. ∞Z 0 Таким образом, по следствию 8 существует ограниченный обратный оператор (log(I − A))−1x = ∞Z 0 T (t)xf (t)dt.(cid:3) Список литературы [1] Хилле, Э. Функциональный анализ и полугруппы /Э. Хилле, Р. Филлипс. -- М. : ИЛ, 1962. -- 829 с. [2] А. Р. Миротин, Расширение функционального исчисления Хилле-Филлипса, Из- вестия ГГУ им. Ф. Скорины, 2009, № 5 (56), с. 66 -- 69. 9 [3] B. Baeumer, M. Haase and M. Kovacs, Unbounded functional calculus for bounded groups with applications, J. Evol. Equ., 2009, Vol. 9, no 1, p. 171 -- 195. [4] О. В. Лопушанский, С. В. Шарин, Обобщенное функциональное исчисление типа Хилле -- Филлипса для многопараметрических полугрупп, Сибирский ма- тем.журнал, 2014. Том 55, № 1, с. 131 -- 146. [5] Д.Хенри, Геометрическая теория полулинейных параболических уравнений, М.: Мир, 1985. [6] C. Martinez Carracedo and M. Sanz Alix, The Theory of Fractional Powers of Operators, North- Holland Mathematics Studies 187, Elsevier, Amsterdam - London - New York - Oxford - Paris - Shannon - Tokyo, 2001. [7] D. V. Widder, The Laplase Transform, Prinston, 1946. [8] Carasso A. S., On Subordinate Holomorphic Semigroups / A. S. Carasso, T. Kato // Trans. Amer. Math. Soc. -- 1991. -- Vol. 327, N 2. -- P. 867 -- 878. [9] Миротин, А. Р. О T -исчислении генераторов C0-полугрупп / А. Р. Миротин // Сибирский матем. журнал. -- 1998. -- Т. 39, N 3. -- С. 571 -- 583. A. R. Mirotin, "On the T -calculus of generators of C0-semigroups", Siberian Math. J., 39:3 (1998) [10] Миротин, А. Р. Многомерное T -исчисление генераторов C0-полугрупп/А. Р. Ми- ротин // Алгебра и анализ. -- 1999. -- Т. 11, № 2. -- С. 142 -- 170; A. R. Mirotin, "The multidimensional -calculus of generators of C0-semigroups", St. Petersburg Math. J., 11:2 (2000), 315 -- 335 [11] Mirotin, A. R. Criteria for Analyticity of Subordinate Semigroups / A. R. Mirotin // Semigroup Forum. -- 2009. -- Vol. 78, № 2. -- Р. 262 -- 275. [12] Миротин, А. Р. Функции класса Шенберга T действуют в конусе диссипативных элементов банаховой алгебры. II, Матем. заметки, 61:4 (1997), 630 -- 633 ; A. R. Mirotin, "Functions from the Schoenberg class T on the cone of dissipative elements of a Banach algebra", Math. Notes, 61:4 (1997), 524 -- 527. [13] Миротин, А. Р. Функции класса Шенберга T действуют в конусе диссипатив- ных элементов банаховой алгебры. II, Матем. заметки, 64:3 (1998), 423 -- 430; A. R. Mirotin, "Functions from the Schoenberg class T act in the cone of dissipative elements of a Banach algebra. II", Math. Notes, 64:3 (1998), 364 -- 370. [14] Mirotin, A. R. On multidimensional Bochner-Phillips functional calculus (2019) - arXiv preprint, arXiv:1902.08762. [15] Миротин, А. Р. О совместных спектрах наборов неограниченных операторов, Изв. РАН. Сер. матем., 79:6 (2015), 145 -- 170; A. R. Mirotin, "On joint spectra of families of unbounded operators", Izv. Math., 79:6 (2015), 1235 -- 1259. 10 [16] А. Р. Миротин, Об одном функциональном исчислении замкнутых операторов в банаховом пространстве. III. Некоторые вопросы теории возмущений, Изв. ву- зов. Матем., 2017, 12, 24 -- 34; A. R. Mirotin, On some functional calculus of closed operators on Banach space. III. Some topics of perturbation theory, Russian Math. (Iz. VUZ), 61:12 (2017), 19 -- 28 [17] Бейтмен Г., Эрдейи А. Таблицы интегральных преобразований. Т. 1. Преобра- зования Фурье, Лапласа, Меллина, М. : Наука, 1969. 11
1703.03724
1
1703
2017-03-10T15:45:51
Strong transitivity properties for operators
[ "math.FA" ]
Given a Furstenberg family $\mathscr{F}$ of subsets of $\mathbb{N}$, an operator $T$ on a topological vector space $X$ is called $\mathscr{F}$-transitive provided for each non-empty open subsets $U$, $V$ of $X$ the set $\{n\in \mathbb{Z}_+ : T^n(U)\cap V\neq\emptyset\}$ belongs to $\mathscr{F}$. We classify the topologically transitive operators with a hierarchy of $\mathscr{F}$-transitive subclasses by considering families $\mathscr{F}$ that are determined by various notions of largeness and density in $\mathbb{Z}_+$.
math.FA
math
Strong transitivity properties for operators J. Bès ∗, Q. Menet†, A. Peris‡and Y. Puig§ Abstract Given a Furstenberg family F of subsets of N, an operator T on a topological vector space X is called F -transitive provided for each non-empty open subsets U, V of X the set {n ∈ Z+ : T n(U )∩V (cid:54)= ∅} belongs to F . We classify the topologically transitive operators with a hierarchy of F -transitive subclasses by considering families F that are determined by various notions of largeness and density in Z+. 1 Introduction Throughout this paper X denotes a topological space and U(X) the set of non-empty open subsets of X. When X is a topological vector space, L(X) stands for the set of operators (i..e, linear and continuous self-maps) on X. An operator T ∈ L(X) is called hypercyclic if there exists a vector x ∈ X such that for each V in U(X) the time return set NT (x, V ) = N (x, V ) := {n ≥ 0 : T nx ∈ V } is non-empty, or equivalently (since X has no isolated points) an infinite set. When X is an F -space (that is, a complete and metrizable topological vector space), we know thanks to Birkhoff's transitivity theorem that T is hypercyclic if and only if it is topologically transitive, that is, provided NT (U, V ) = N (U, V ) := {n ≥ 0 : T n(U ) ∩ V (cid:54)= ∅} 7 1 0 2 r a M 0 1 ] . A F h t a m [ 1 v 4 2 7 3 0 . 3 0 7 1 : v i X r a is infinite for every U, V ∈ U(X). Since 2004, several refined notions of hypercyclicity based on the prop- erties of time return sets N (x, V ) have been investigated: frequent hyper- cyclicity [3, 2], U-frequent hypercyclicity [21, 9], reiterative hypercyclicity ∗Department of Mathematics and Statistics, Bowling Green State University, Bowling †Univ. Artois, EA 2462, Laboratoire de Mathématiques de Lens (LML), F-62300 Lens, ‡Departament de Matemàtica Aplicada, IUMPA, Universitat Politècnica de València, Green, OH 43403, USA. e-mail: [email protected] France. e-mail: [email protected] Edifici 7A, 46022 València, Spain. e-mail: [email protected] §Department of Mathematics, Ben-Gurion University of the Negev, P.O.B. 653, 84105 Beer Sheva, Israel. e-mail: [email protected] 1 [7]. More recently a general notion called A-hypercyclicity, which generalizes the abovementioned notions of hypercyclicity, has been used to investigate the different types of hypercyclic operators, see [7, 9]. Our aim here is to investigate refined notions of topological transitivity based on properties satisfied by the return sets N (U, V ). Some of these are already well-known, such as the topological notions of mixing, weak- mixing, and ergodicity, say. Recall that a continuous self-map T on X is called mixing provided N (U, V ) is cofinite for each U, V ∈ U(X). Also, T is called weakly mixing whenever T × T is topologically transitive on X × X, and this occurs precisely when the return set N (U, V ) is thick (i.e. contains arbitrarily long intervals) for each U, V ∈ U(X) [19]. Finally, T is topologically ergodic provided N (U, V ) is syndetic (i.e. has bounded gaps) for each U, V ∈ U(X). It is known that topologically ergodic operators are weakly mixing [14]. The above mentioned notions may be stated through the concept of a (Furstenberg) family. The symbols Z and Z+ denote the sets of integers and of positive integers, respectively. Definition 1.1. We say that a non-empty collection F of subsets of Z+ is a family provided that each set A ∈ F is infinite and that F is hereditarily upward (i.e. for any A ∈ F , if B ⊃ A then B ∈ F ). The dual family F ∗ of F is defined as the collection of subsets A of Z+ such that A∩ B (cid:54)= ∅ for every B ∈ F . Some standard families are the following: The family I of infinite sets, whose dual family I∗ coincides with the family of cofinite sets. The family T of thick sets, whose dual family is S = T∗, the family of syndetic sets. For a topologically transitive map T a distinguished family is NT := {A ⊂ Z+ : NT (U, V ) ⊆ A for some U, V ∈ U(X)}. From now on the symbol F will always denote a family. Definition 1.2. We say that a continuous map T on X is F -transitive (or an F -map, for short) provided NT ⊂ F , that is, provided N (U, V ) ∈ F for each U, V ∈ U(X). If in addition X is a topological vector space and T ∈ L(X) we call T an F -transitive operator (or F -operator for short). Hence the I-operators are precisely those operators which are topolog- ically transitive, and the I∗-operators and T-operators are precisely those which are mixing and weak mixing, respectively. The T∗ = S-operators, that is, the topologically ergodic operators. We present here some new classes of topologically transitive operators by considering families F defined in terms of various notions of density and largeness in Z+. A hierarchy of fourteen classes (which include the earlier mentioned classes defined by properties of return sets N (x, V )) appears in Figure 2 and summarizes our findings. We stress that while trivially any F1-map is an F2-map when F1 ⊂ F2, it is possible that the classes of 2 F1-operators and of F2-operators coincide even if F1 is strictly contained in F2 (see e.g., Proposition 5.1). The paper is organized as follows. In Section 2 we describe some gen- eral facts about families F and their corresponding F -transitive maps and operators. In Theorem 2.4 we provide an extension of the Hypercyclicity Cri- terion that ensures an operator to be F -transitive. We apply this criterion in Section 3 to characterize F -transitivity among unilateral and bilateral weighted backward shift operators on c0 and (cid:96)p (1 ≤ p < ∞) spaces. To il- lustrate, we establish in Corollary 3.4 that a unilateral backward shift Bw is topologically ergodic precisely when its weight sequence w = (wn)n satisfies that each set AM = {n : n(cid:89) wj > M} (M > 0) j=1 is syndetic. Section 4 is dedicated to F -operators induced by families F given by sets of positive or full (lower or upper) asymptotic density or Banach density. In Section 5, we look at F -operators induced by families F commonly used in Ramsey theory, and we compare the classes that we obtain with the class of reiteratively hypercyclic operators (Subsection 5.1). Some natural questions conclude the paper. 2 F -Transitivity In this section we introduce a sufficient condition for an operator to be an F -operator, which we call the F -Transitivity Criterion, and it is in the same vein of the Hypercyclicity Criterion. Moreover, we will study the notion of hereditarily F -operator. We will be interested in the following three special properties a family F can have: being a filter, being partition-regular, and being shift-invariant. We use the following notation: given two families F1 and F2 F1 · F2 := {A ∩ B : A ∈ F1, B ∈ F2}. Obviously, F1 ⊂ F1 · F2 and F2 ⊂ F1 · F2. A family F is a filter provided it is invariant under finite intersections (i.e., provided F · F ⊂ F ). Say, the family I∗ of cofinite sets is a filter while the families I and S of infinite sets and of syndetic sets are not. The second property, that of being partition regular, will be useful for us to identify filters. A family F on Z+ is said to be partition regular if for every A ∈ F and any finite partition {A1, . . . , An} of A, there exists some i = 1, . . . , n such that Ai ∈ F . The family I is an example of partition regular family, while the families I∗, T and S are not. Later we will see other examples of partition regular families: the family of piecewise syndetic sets (see Remark 2.5), the family of sets with positive upper (Banach) density (see Section 4), the families of ∆-sets and of IP-sets (see Section 5). 3 Lemma 2.1. Given a family F , the following are equivalent: (I) F is partition regular, (II) A ∩ A(cid:48) ∈ F for every A ∈ F and A(cid:48) ∈ F ∗ (i.e., F · F ∗ ⊂ F ), (III) F ∗ is a filter. Proof. (I) =⇒ (II): Given A ∈ F and A(cid:48) ∈ F ∗ it is clear that A ∩ A(cid:48) (cid:54)= ∅ by definition of dual family. Since (A∩ A(cid:48))∪ (A\ A(cid:48)) = A, either A∩ A(cid:48) ∈ F or A \ A(cid:48) ∈ F by (I). Since (A \ A(cid:48)) ∩ A(cid:48) = ∅, by definition of dual family we necessarily have A ∩ A(cid:48) ∈ F . (II) =⇒ (III): For arbitrary A(cid:48), B(cid:48) ∈ F ∗ and A ∈ F , by applying (II) and the definition of dual family we have A∩ (A(cid:48) ∩ B(cid:48)) = (A∩ A(cid:48))∩ B(cid:48) (cid:54)= ∅, which yields that F ∗ is a filter. (III) =⇒ (I): We will just show that, given A ∈ F and A1, A2 ⊂ Z+ such that A1 ∪ A2 = A and A1 ∩ A2 = ∅, then either A1 ∈ F or A2 ∈ F . The general case can be deduced by an inductive process. Since F = F ∗∗, we need to show that Ai ∩ A(cid:48) (cid:54)= ∅ for every A(cid:48) ∈ F ∗, for i = 1 or i = 2. Suppose that there exist A(cid:48), B(cid:48) ∈ F ∗ with A1 ∩ A(cid:48) = ∅ and A2 ∩ B(cid:48) = ∅. Since F ∗ is a filter, then C(cid:48) := A(cid:48) ∩ B(cid:48) ∈ F ∗. Thus, ∅ (cid:54)= A ∩ C(cid:48) ⊂ (A1 ∩ A(cid:48)) ∪ (A2 ∩ B(cid:48)) = ∅, which is a contradiction. Notice that (F ∗)∗ = F for any family F : the inclusion F ⊂ (F ∗)∗ is immediate. Conversely, if A ∈ (F ∗)∗, then Z+\A (cid:54)∈ F ∗ by the definition of a dual family. This means that there exists B ∈ F such that B∩(Z+\A) = ∅. That is, B ⊂ A, which gives A ∈ F . Thus any family is a dual family, and Lemma 2.1 also gives that a family F is a filter if and only if F ∗ · F ⊂ F ∗ and if and only if F ∗ is partition regular. Another consequence of Lemma 2.1 is that any family F that is both a filter and partition regular (called an ultrafilter) must satisfy F = F ∗. Finally, our third property: A family F on Z+ is said to be shift−- invariant provided for every i ∈ Z+ and each A ∈ F , we have (A−i)∩Z+ ∈ F . We say that F is called shift+-invariant if for every i ∈ Z+ and each A ∈ F , we have A + i ∈ F . When F is both, shift−-invariant and shift+- invariant, we simply call it shift invariant. For instance, the families of infinite sets, cofinite sets, thick sets and syndetic sets are shift invariant. We may gain shift invariance by reducing a family. Given a family F , we define(cid:102)F+ = {A ⊂ Z+ : ∀N ∈ Z+ ∃B ∈ F such that A ⊃ B + [0, N ]}, (cid:102)F− = {A ⊂ Z+ : ∀N ∈ Z+ ∃B ∈ F such that A ⊃ (B + [−N, 0]) ∩ Z+}, 4 invariant. (cid:102)F = {A ⊂ Z+ : ∀N ∈ Z+ ∃B ∈ F such that A ⊃ (B + [−N, N ]) ∩ Z+}. So for any family F we have the inclusions (cid:102)F ⊂ (cid:102)F+ ⊂ F and (cid:102)F ⊂ (cid:102)F− ⊂ F , and that (cid:102)F− is shift+-invariant, (cid:102)F+ is shift−-invariant, and (cid:102)F is shift Lemma 2.2. If F is a filter on Z+, so is (cid:102)F . Moreover, for any family F satisfying (cid:102)F · (cid:102)F ⊂ F the subfamily (cid:102)F is a filter. Proof. Let A1, A2 ∈ (cid:102)F . We have to show that A1 ∩ A2 ∈ (cid:102)F . Given N ∈ N, there are B1(N ), B2(N ) ∈ F such that (B1(N ) + [−2N, 2N ]) ∩ Z+ ⊂ A1 and (B2(N ) + [−2N, 2N ]) ∩ Z+ ⊂ A2. For i = 1, 2 we define ¯Ai(N ) := (Bi(J) + [−J, J]) ∩ Z+. (cid:91) J≥N Clearly ¯A1(N ), ¯A2(N ) ∈ (cid:102)F for each N ∈ N. By hypothesis, B(N ) := ¯A1(N ) ∩ ¯A2(N ) ∈ F , N ∈ N. To prove that A1 ∩ A2 ∈ (cid:102)F we just need to show that (B(N ) + [−N, N ]) ∩ Z+ ⊂ A1 ∩ A2 for every N ∈ N. Indeed, given N ∈ N and m ∈ (B(N ) + [−N, N ]) ∩ Z+, we write m = k(N ) + l(N ) with k(N ) ∈ B(N ) and l(N ) ∈ [−N, N ]. By definition of B(N ) we have k(N ) = k1(J1) + l1(J1) = k2(J2) + l2(J2) for some ki(Ji) ∈ Bi(Ji), li(Ji) ∈ [−Ji, Ji], Ji ≥ N, i = 1, 2. Thus m = k1(J1) + l1(J1) + l(N ) ∈ (B1(J1) + [−2J1, 2J1]) ∩ Z+ ⊂ A1, and, analogously, m ∈ A2, which yields the result. (cid:102)F -map is an F -map, since (cid:102)F ⊂ F . The next lemma gives conditions for The rest of the section is dedicated to F -maps and F -operators. Every the converse, and is used in Proposition 3.1. Lemma 2.3. Let F be a family on Z+ and let T be a F -map. The following are equivalent. (i) T is weakly mixing, (ii) T is an (cid:102)F -map. Proof. (i) implies (ii): Given N ∈ N and U, V ∈ U(X), since T is weakly mixing, by Furstenberg result we know that NT is a filter, so there are U(cid:48), V (cid:48) ∈ U(X) such that N (U(cid:48), V (cid:48)) ⊂ N (T −m(U ), V ) ∩ N (U, T −m(V )), for m = 0, . . . , N. By F -transitivity we have N (U(cid:48), V (cid:48)) ∈ F . We then conclude that (N (U(cid:48), V (cid:48))+[−N, N ])∩Z+ ⊂ N (U, V ), and T is (cid:102)F -transitive. (ii) implies (i): If T is an (cid:102)F -map, since every element of (cid:102)F is thick, we have that NT consists of thick sets and, as we already recalled in the introduction, this means that T is weakly mixing. 5 To state the F -Transitivity Criterion, we recall the notion of limit along a family F : Given a sequence {xn}n in X and x ∈ X, we say that F − limn xn = x, or that xn F→ x, provided {n ∈ Z+ : xn ∈ U} ∈ F for each neighbourhood U of x. Theorem 2.4. (F -Transitivity Criterion) Let T be an operator on a topo- logical vector space X and let F be a family on Z+ such that (cid:102)F is a filter. Suppose there exist D1, D2 dense sets in X, and (possibly discontinuous) mappings Sn : D2 → X, n ∈ N satisfying (a) F -limn T n(x) = 0 for every x ∈ D1 (b) F -limn(Sn(y), T nSn(y)) = (0, y) for every y ∈ D2. Then T is an (cid:102)F -operator. Proof. Let U, V ∈ U(X). We fix U(cid:48), V (cid:48) ∈ U(X) and a 0-neighbourhood W such that U(cid:48) +W ⊂ U and V (cid:48) +W ⊂ V . Given N ∈ N, pick x ∈ D1∩T −N U(cid:48) and y ∈ D2 ∩ T −N V (cid:48). By continuity of T we easily get (cid:102)F+ − lim which yields N (T −N U(cid:48), W ) ∈ (cid:102)F+. That is, there is A ∈ F such that n A + [0, 2N ] ⊂ N (T −N U(cid:48), W ). Therefore, T nx = 0, (A + [−N, N ]) ∩ Z+ ⊂ (N (T −N U(cid:48), W ) − N ) ∩ Z+ ⊂ N (U(cid:48), W ), for m = 0, . . . , 2N and for every n ∈ A. In particular, (A + [−N, N ])∩Z+ ⊂ Also, we find a 0-neighbourhood W (cid:48) ⊂ W with T m(W (cid:48)) ⊂ W and y + W (cid:48) ⊂ T −N V (cid:48), m = 0, . . . , 2N. There is A ∈ F such that Sny ∈ W (cid:48) and T nSn(y) ∈ y + W (cid:48) for all n ∈ A. Thus, and, since N was arbitrary, we have that N (U(cid:48), W ) ∈ (cid:102)F . (cid:0)T (n−m)(T mSn(y)), T mSn(y)(cid:1) ∈ (y + W (cid:48), T m(W (cid:48))) ⊂ (T −N V (cid:48), W ), N (W, V (cid:48)). Since N was arbitrary, we obtain that N (W, V (cid:48)) ∈ (cid:102)F . Therefore, N (U, V ) ⊃ N (U(cid:48) + W, V (cid:48) + W ) ⊃ N (U(cid:48), W ) ∩ N (W, V (cid:48)) ∈ (cid:102)F · (cid:102)F ⊂ (cid:102)F , that is, T is an (cid:102)F -operator. 1. By Lemma 2.2 the assumption that (cid:102)F be a filter is not a filter, and(cid:101)S = TS is the family of thickly syndetic sets, which is trivially satisfied in the case that F is a filter, but Theorem 2.4 applies beyond this case. For instance, the family F = S of syndetic sets is Remark 2.5. 6 intersection of a thick set with a syndetic set), then (cid:102)PS = T, and a filter. So every operator that satisfies the S-Transitivity Criterion is a TS-operator. In contrast, if we consider the family of piecewise syndetic sets PS = = T · S (i.e., A is piecewise syndetic if, and only if, it is the TS∗ ∅ ∈ T · T. Thus the hypotheses of Theorem 2.4 are not satisfied. Actually, it is not hard to construct an operator T such that conditions (a) and (b) in Theorem 2.4 are satisfied for F = PS, with T not even transitive. 2. Another remarkable case is provided by, given a strictly increasing sequence (nk)k in N, considering the filter F := {A ⊂ N : ∃j ∈ N with A ⊃ {nk : k ≥ j}}. In this case Theorem 2.4 turns out to coincide with the classical Hy- percyclicity Criterion. Moreover, since the Hypercyclicity Criterion characterizes the weakly mixing operators on separable F -spaces [8], we have that every weakly mixing operator T on a separable F -space X supports a strictly increasing sequence (nk)k in N such that T is an F -operator, where F := {A ⊂ N : ∀N ∈ N ∃j ∈ N with A ⊃ {nk : k ≥ j} + [−N, N ]}. 3. We note that for an (cid:102)F -operator T with (cid:102)F a filter it is not true in G ⊂ (cid:102)F : just consider the family F = I∗ of cofinite sets and the fact general that T must satisfy the G-Transitivity Criterion for some filter that there exist mixing operators not satisfying Kitai's Criterion [12, Theorem 2.5]. 4. Recall that for the case F = I, Furstenberg [10, Proposition II.3] showed that once T ⊕ T is an I-map on X 2, every direct sum ⊕r j=1T on X r is an I-map too (r ∈ N). The assumptions of the F -Transitivity Criterion on an operator T clearly ensure that (any direct sum ⊕r j=1T will satisfy the F -Transitivity Criterion on the space X r and thus that) ⊕r j=1T is an (cid:102)F -operator on X r, for every r ∈ N. We next introduce the concept of a hereditarily F -operator, and we establish links with that of an F -operator. Definition 2.6. We say that a continuous map T is a hereditarily F -map if N (U, V ) ∩ A ∈ F for every U, V ∈ U(X) and every A ∈ F (that is, NT · F ⊂ F ). In addition, if X is a topological vector space and T ∈ L(X), we say that T is a hereditarily F -operator. Clearly, hereditarily F -maps are F -maps. Moreover, they are automat- ically F ∗-maps since NT · F ⊂ F (cid:54)(cid:51) ∅. Also, for a filter F the concepts of F -map and hereditarily F -map are equivalent. More generally, we have: 7 Proposition 2.7. Let T be a continuous map on a complete separable met- ric space X without isolated points. (A) Let F be a partition regular family. Then the following are equivalent: (1) T is an F ∗-map; (2) T is a hereditarily F ∗-map; (3) T is a hereditarily F -map; (4) hcA := {x ∈ X : {T nx : n ∈ A} = X} is a dense (Gδ) set in X for any A ∈ F . (B) Let F be a filter. Then the following are equivalent: (i) T is an F -map; (ii) T is a hereditarily F -map; (iii) T is a hereditarily F ∗-map; (iv) hcA := {x ∈ X : {T nx : n ∈ A} = X} is a dense (Gδ) set in X for any A ∈ F ∗. Proof. We will just show (A) since (B) follows by taking duals and Lemma 2.1. Indeed, condition (1) is equivalent to (2) because F ∗ is a filter. The fact that (1) implies (3) is a consequence of Lemma 2.1 too, while the converse was already noticed before for general families. Finally the equivalence be- tween (1) and (4) can be shown in a similar way as Birkhoff's transitivity theorem [15]. Note that when considering the family F = I of infinite sets in Propo- sition 2.7 (A) we obtain the known equivalences for mixing maps. Remark 2.8. By the same argument for an operator T on a separable topological vector space X, the first three equivalences of statements (A) and (B) still hold. We also point out that as with the hypercyclic case we have the following comparison principle for F -maps and transference principle for F -operators, see [15, Chapter 12]. 1. 2. (F -Comparison Principle) Any quasifactor of an F -map is an F - map. Indeed, let T : X → X be an F -map and let S : Y → Y and φ : X → Y be maps so that φ ◦ T = S ◦ φ, where φ has dense range. Then for any non-empty open subsets U and V of Y we have NS(U, V ) = NT (φ−1(U ), φ−1(V )) ∈ F . (Transference Principle) Let F be a family and let T be an operator on a topological vector space X so that each operator S on an F -space that is quasi-conjugate to T via an operator (that is, it supports a dense range operator J : X → Y with JT = SJ) is an F -map. Then T is an F -map. 8 3 F -transitive weighted shift operators Each bounded bilateral weight sequence w = (wk)k∈Z, induces a bilateral weighted backward shift operator Bw on X = c0(Z) or (cid:96)p(Z) (1 ≤ p < ∞) given by Bwek := wkek−1, where (ek)k∈Z denotes the canonical basis of X. Similarly, each bounded sequence w = (wn)n∈N induces a unilateral weighted backward shift operator Bw on X = c0(Z+) or (cid:96)p(Z+) (1 ≤ p < ∞), given by Bwen := wnen−1, n ≥ 1 and Bwe0 := 0, where (en)n∈Z+ denotes the canonical basis of X. Our characterization of F -transitive weighted backward shifts will rely on the properties of the sets AM,j and ¯AM,j defined as (cid:110) (cid:110) AM,j := ¯AM,j := n ∈ N : n ∈ N : (cid:111) wi > M j+n(cid:89) 1(cid:81)j i=j−n+1 wi > M i=j+1 (cid:111) , where M > 0 and j ∈ Z. In the case j = 0, we just write AM , ¯AM instead of AM,0, ¯AM,0 respectively. We note that Salas' [20] characterization of hy- percyclic (i.e., transitive) bilateral weighted shifts on the above sequence spaces may be formulated as Bw is hypercyclic ⇔ ∀M > 0 ∀N ∈ N (AM,j ∩ ¯AM,j) (cid:54)= ∅. N(cid:92) j=−N In other words, since AM(cid:48),j ⊂ AM,j and ¯AM(cid:48),j ⊂ ¯AM,j whenever M(cid:48) > M > 0, the collection of subsets {AM,j, ¯AM,j}M >0,j∈Z should form a filter subbase for the hypercyclicity of Bw. In that case, we denote by Aw the generated filter. Therefore, for the characterization of weighted shifts Bw that are F -operators for a certain family F we need to assume that Aw is a filter. When Bw is hypercyclic (i.e., when Aw is a filter), we can describe a filter base of Aw, which will be very useful in the characterization of weighted shifts that are F -operators, and it is given by the collection of sets {AM,j ∩ ¯AM,j : M > 0 and j ∈ N}. Actually, this is a consequence of the observation that, if M1, M2 > 0 and j1, j2, j3 ∈ Z with j3 > max{j1 ,j2}, then there is M3 > 0 such that AM3,j3 ⊂ AM1,j1 ∩ AM2,j2 and Indeed, let M := supi∈Z wi. We fix M3 > K(M1 + M2)(1 + M )2j3, where ¯AM3,j3 ⊂ ¯AM1,j1 ∩ ¯AM2,j2. m2(cid:89) wi−1 . K := 1 + max −j3≤m1≤m2≤j3 i=m1 9 If n ∈ AM3,j3 then j1+n(cid:89) wi = (cid:32) j3+n(cid:89) (cid:33) (cid:81)j3 (cid:81)j3+n i=j1+1 wi i=j1+n+1 wi > M3 wi (cid:81)j3 i=j1+1 wi M j3−j1−1 > M1. i=j1+1 i=j3+1 That is, n ∈ AM1,j1. The same argument shows n ∈ AM2,j2. Analogously, we also have ¯AM3,j3 ⊂ ¯AM1,j1 ∩ ¯AM2,j2. Proposition 3.1. Let Bw be a bilateral weighted backward shift on X = c0(Z) or (cid:96)p(Z), 1 ≤ p < ∞. Then the following are equivalent: (1) Bw is an (cid:102)F -operator; (2) Bw is an F -operator; (3) for every j ∈ N and M > 0, AM,j ∩ ¯AM,j ∈ F ; (4) Bw is hypercyclic, Aw ⊂ F , and Bw satisfies the Aw-Criterion. In addition, if (cid:102)F is a filter, then the above conditions are equivalent to (5) for every j ∈ N and M > 0 we have AM,j ∈ F and ¯AM,j ∈ F . Proof. Obviously, (1) implies (2). The reverse implication is a consequence of Lemma 2.3 since transitive weighted shifts are weakly mixing. Also, (4) implies (2). To show that (2) implies (3), given N, j ∈ N arbitrary, we must find nonempty open sets U, V ⊂ X such that N (U, V ) ⊂ AN,j ∩ ¯AN,j. (3.1) Indeed, we fix R > N, U := {x ∈ X : xj > } ∩ {x ∈ X : (cid:107)x(cid:107) < 1}, 1 R and we set If m ∈ N (U, V ) and x ∈ U is such that Bm (cid:13)(cid:13) < (cid:9). 1 R2 w x ∈ V , then 1 R2 < 1 and if l (cid:54)= j. (3.2) (3.3) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) wi xj+m − (N + 1) V =(cid:8)x ∈ X :(cid:13)(cid:13)x − (N + 1)ej (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < (cid:32) j+m(cid:89) (cid:33) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:32) l+m(cid:89) j+m(cid:89) (cid:33) (cid:32) j+m(cid:89) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < 1 R2 (cid:33) wi xl+m i=j+1 i=l+1 i=j+1 i=j+1 10 wi > wi xj+m > N, Since x ∈ U, we deduce from (3.2) that which implies that m ∈ AN,j. On the other hand, Bm intersect. Thus, l := j − m (cid:54)= j, and (3.3) implies w x ∈ V forces m > 0 since U and V do not (cid:33) (cid:33) (cid:32) j(cid:89) (cid:32) j(cid:89) wi < i=j−m+1 i=j−m+1 wi R xj < 1 R < 1 N , that yields m ∈ ¯AN,j. Thus the inclusion (3.1) is satisfied, and property (3) holds. To prove that (3) implies (4), since Bw is hypercyclic (i.e., Aw is a filter) and Aw ⊂ F because F contains a basis of Aw, we just need to show that Bw satisfies the Aw-criterion. Let D be the set of all finitely supported vectors in X and let Sw be the weighted forward shift defined on D by 1 Swei := ei+1. wi+1 wSnx = x for every x ∈ D. It suffices w then we have Bn If we consider Sn := Sn to show that wx = 0 for every x ∈ D; • Aw-limn Bn • Aw-limn Snx = 0 for every x ∈ D. For the rest of the proof we assume that X = (cid:96)p(Z) with 1 ≤ p < ∞. The proof is similar if X = c0(Z). Let x ∈ D, ε > 0 and Vε := {x ∈ (cid:96)p(Z) : wx ∈ Vε} ∈ Aw. Since x ∈ D, we (cid:107)x(cid:107) < ε}. First, we show that {n ∈ N : Bn (cid:33) j=−m xjej for some m ∈ N and we then have can write x =(cid:80)m (cid:32) j+n(cid:89) Bn wx = m−n(cid:88) Let M = (cid:107)x(cid:107)∞2m/ε and n ∈(cid:84)m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p wx(cid:107)p = m(cid:88) (cid:107)Bn i=j−n+1 j=−m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) wi j=−m−n i=j+1 wi xj+nej. ¯AM,j ∈ Aw. We have j=−m (cid:18) m(cid:88) j=−m xjp < ε (cid:107)x(cid:107)∞2m (cid:19)p xjp < εp, which implies thus {n ∈ N : Bn Vε} ∈ Aw. Indeed, we have wy ∈ Vε} ∈ Aw. It remains to show that {n ∈ N : Snx ∈ j(cid:89) m(cid:92) j=−m ¯AM,j ⊆ {n ∈ N : Bn wy ∈ Vε}, Snx = Sn wx = xj(cid:81)j+n i=j+1 wi ej+n. m(cid:88) j=−m 11 Let M = (cid:107)x(cid:107)∞2m/ε and n ∈(cid:84)m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) m(cid:88) xj(cid:81)j+n (cid:107)Snx(cid:107)p = j=−m i=j+1 wi (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p which implies j=−m AM,j. We then have < 2mεp (2m)p ≤ εp, m(cid:92) j=−m AM,j ⊆ {n ∈ N : Sny ∈ Vε}. Consequently, {n ∈ N : Sny ∈ Vε} ∈ Aw, and Bw is an F -operator. Certainly, condition (3) implies (5). If (5) holds, the argument preceding this Proposition yields that, for each j ∈ N and for every M > 0, the sets AM,j and ¯AM,j belong to (cid:102)F , which gives (3) since (cid:102)F is a filter. When F = I∗ is the filter of cofinite sets, we obtain as a consequence the well known characterization of mixing bilateral weighted shifts. On the other hand, the case F = S offers again an interesting result. Corollary 3.2. Let Bw be a bilateral weighted backward shift on X = c0(Z) or (cid:96)p(Z), 1 ≤ p < ∞. Then the following are equivalent: (1) Bw is a topologically ergodic operator; (2) for every j ∈ N and M > 0, AM,j and ¯AM,j are syndetic sets. The unilateral version of Proposition 3.1 we provide next relies only on the sets AM,j. Notice that for a hypercyclic unilateral weighted shift Bw the collection of sets {AM,j : M > 0 and j ∈ N} forms a base of a filter (which we call again Aw) since, as before, if M1, M2 > 0 and j1, j2, j3 ∈ N with j3 > max{j1, j2}, then there is M3 > 0 such that AM3,j3 ⊂ AM1,j1 ∩ AM2,j2. This fact yields a simplification of the corresponding characterization of uni- lateral weighted shifts that are F -operators, which can be further simplified if F is a shift−-invariant family. The unilateral version of Proposition 3.1 can be stated as follows. Proposition 3.3. Let Bw be an unilateral weighted backward shift on c0(Z+) or (cid:96)p(Z+) (1 ≤ p < ∞). The following are equivalent: (1) Bw is an (cid:102)F -operator; (2) Bw is an F -operator; (3) for every j ∈ N and M > 0, the set AM,j ∈ F ; (4) Bw is hypercyclic, Aw ⊂ F , and Bw satisfies the Aw-Criterion. 12 If in addition F is shift−-invariant, the above conditions are equivalent to (5) for every M > 0 the set AM ∈ F . Proof. We only prove that if F is shift−-invariant then condition (5) implies (3). Let M > 0 and j ∈ N. We fix M(cid:48) > M (supi∈N wi)j such that AM(cid:48) ⊂ (cid:81)n [j + 1, +∞[. Given n ∈ AM(cid:48), we have (cid:81)j s=1 ws s=1 ws > (supi∈N wi)j > M. ws = n(cid:89) M(cid:48) s=j+1 This implies that AM(cid:48) − j ⊂ AM,j. Since F is a shift−-invariant family, we conclude that AM,j ∈ F . In consequence we have the following characterization of topologically ergodic unilateral backward weighted shifts. Corollary 3.4. Let Bw be an unilateral weighted backward shift on X = c0(Z+) or (cid:96)p(Z+), 1 ≤ p < ∞, then the following are equivalent: (1) Bw is topologically ergodic; (2) for every M > 0 the set AM is syndetic. We conclude this section by considering finite products of F -maps. Proposition 3.5. Let T1, . . . , Tm be continuous maps on X, then (1) for n ≥ 1, T n 1 is an F -map on X if and only if T1 is an Fn-map where Fn := {A ⊂ Z+ : 1 1 is an F -map on X if and only if for every U, V ∈ U(X), NT1(U, V ) ∩ nZ+ ∈ nF . (2) If F is a filter then T1 × T2 × ··· × Tm is an F -map on X m if and n(A∩ nZ+) ∈ F}. In other words, T n only if Tl is an F -map on X for every 1 ≤ l ≤ m. 1 is an F -map on X if and only if NT n (U, V ) ∈ F Proof. (1) If n ≥ 1, then T n n(NT1(U, V ) ∩ nZ+). for every U, V ∈ U(X). We remark that NT n (cid:84)m Therefore, NT n (2) Note that T1 × T2 × ··· × Tm is an F -map on X m if and only if l=1 ∈ (U(X)× U(X))m. The conclusion l=1 NTl(Ul, Vl) ∈ F , for any (Ul, Vl)m follows since F is a filter. (U, V ) ∈ F if and only if NT1(U, V ) ∈ Fn. (U, V ) = 1 1 1 1 Hence by Proposition 3.1 and Proposition 3.5 we have the following corollary. Corollary 3.6. Let F be a filter and Bw be a bilateral weighted backward shift on X = (cid:96)p(Z) or c0(Z). Then, for every m ∈ N, the following are equivalent: (1) Bw ⊕ B2 (2) For every 1 ≤ l ≤ m, M > 0 and j ∈ Z, AM,j ∩ lZ+ ∈ lF and ¯AM,j ∩ lZ+ ∈ lF . w is an F -operator on X m; w ⊕ ... ⊕ Bm 13 4 Return sets and densities The purpose of this section is to analyze which kind of density properties the sets N (U, V ) can have for a given hypercyclic operator, and classify the hypercyclic operators accordingly. We first recall the definitions of the asymptotic densities and the Banach densities in Z+. Definition 4.1. Let A ⊆ Z+ be given. The upper and lower asymptotic density of A are defined respectively by d(A) = lim sup n→∞ A ∩ {1, 2, ..., n} n and d(A) = lim inf n→∞ A ∩ {1, 2, ..., n} n . The upper and lower Banach density of A are defined by Bd(A) = lim s→∞ αs/s and Bd(A) = lim s→∞ αs/s, where for each s ∈ Z+ αs = lim sup k→∞ A ∩ [k + 1, k + s] and αs = lim inf k→∞ A ∩ [k + 1, k + s]. In general we have Bd(A) ≤ d(A) ≤ d(A) ≤ Bd(A) and d(A) + d(Z+ \ A) = 1 and Bd(A) + Bd(Z+ \ A) = 1. (4.1) We will consider the following families. D = {A ⊆ Z+ : d(A) > 0}, D = {A ⊆ Z+ : d(A) > 0}, BD = {A ⊆ Z+ : Bd(A) > 0}, BD = {A ⊆ Z+ : Bd(A) > 0}, D1 = {A ⊆ Z+ : d(A) = 1}, D 1 = {A ⊆ Z+ : Bd(A) = 1}, BD1 = {A ⊆ Z+ : Bd(A) = 1}. 1 = {A ⊆ Z+ : d(A) = 1}, BD Notice that each of these families is shift invariant, and that D are filters. Moreover, 1 and BD 1 1. BD1 = T, the family of thick sets, 2. BD = S, the family of syndetic sets, 3. BD ⊃ PS, the family of piecewise syndetic sets, 1 ⊂ TS, the family of thickly syndetic sets, 4. BD 1 = BD 5. BD 1 = D , D ∗ ∗ , D1 = D∗, and BD1 = BD∗ by (4.1). In consequence, T is weakly mixing if and only if T is a BD1-map. Weighted shift operators and Proposition 3.3 help us to provide some counterexamples which allow us to distinguish the different notions of F - operators. 14 Proposition 4.2. Let X = c0(Z+), then (1) there exists a BD1-operator which is not D-operator. (2) there exists a D1-operator which is not D-operator. (3) there exists a D 1-operator which is not BD-operator. Proof. (1) Consider the weight sequence w = (1, . . . , 1 (cid:124) (cid:123)(cid:122) (cid:125) , 2, 2−1, 1, . . . , 1 , 2, 2, 2−2, 1, . . . , 1 , 2, 2, 2, 2−3, 1, . . . , 1 , . . . ) (cid:124) (cid:123)(cid:122) (cid:125) (cid:124) (cid:123)(cid:122) (cid:125) (cid:124) (cid:123)(cid:122) (cid:125) (cid:81)n n2 (cid:124) m0 n0 m1 n1 (cid:123)(cid:122) m2 ,···(cid:1). (cid:125) , 2,··· , 2 (cid:124) (cid:123)(cid:122) (cid:125) (cid:16)(cid:110) n ∈ N : (cid:81)n (cid:124) (cid:123)(cid:122) (cid:125) (cid:125) (cid:124) , 2,··· , 2 m2 m0 m1 We first observe that supn see Chapter 4 in [15]. In other words Bw is BD1-operator. i=1 wi is infinite, hence Bw is weakly mixing, On the other hand, by Proposition 3.3, we know that it suffices to show that d(A1) = 0 in order to deduce that Bw is not a D-operator. In other words, it suffices to show that d = 0 and this holds if (mk) grows sufficiently rapidly. i=1 wi > 1 (cid:111)(cid:17) m3 Thanks to Proposition 3.3, it suffices to find sequences (mk)k, (nk)k such that , 2,··· , 2 , 2−n1, 1,··· , 1 , 2−n0, 1,··· , 1 (2) Consider the weight w =(cid:0) 1,··· , 1 (cid:124) (cid:123)(cid:122) (cid:125) (cid:123)(cid:122) (cid:124) (cid:123)(cid:122) (cid:125) i=1 wi = 1}(cid:1) = 1 • d(cid:0){n :(cid:81)n i=1 wi > M}(cid:1) = 1, for every M > 0. • d(AM ) = d(cid:0){n :(cid:81)n i=1 wi = 1}(cid:1) = 1 then Indeed, if d(cid:0){n :(cid:81)n n(cid:89) wi ≤ 1}(cid:1) = 0. wi > 1}(cid:1) = 1 − d(cid:0){n : d(cid:0){n : k∈N Ak and B =(cid:83) So A =(cid:83) w =(cid:0)1, 2, 2−1, 1, 1, 2,··· , 2 (cid:124) (cid:123)(cid:122) (cid:125) (cid:124) (cid:123)(cid:122) (cid:125) d(AM ) = d(cid:0){n : The set A1 = {n :(cid:81)n wi > M}(cid:1) = 1. , 2−m2,···(cid:1). , 2−m0, 1, 1, 1, 2,··· , 2 (cid:124) (cid:123)(cid:122) (cid:125) 2,··· , 2 Define sequences of intervals in the following way: Ak = [1022k+1, 1022k+2[ and Bk = [1022k+2, 1022k+3[ for every k ∈ Z+. Hence, setting mk = Ak, nk = Bk for every k, we are done. (3) Let mk = 102k for every k ∈ Z+. We consider the weight k∈N Bk are disjoint with d(A) = d(B) = 1. i=1 wi > 1} has arbitrarily large gaps, hence Bw is not an BD-operator by Proposition 3.3. On the other hand, we have for every M > 1 , 2−m1, 1, 1, 1, 1, n(cid:89) i=1 n(cid:89) m0 m1 i=1 m2 Hence, Bw is D 1-operator by Proposition 3.3. i=1 15 Mixing operators obviously are BD this is the argument of the next result. 1-operators, but the converse is false, 1-operator on c0(Z+) which is not mix- Proposition 4.3. There exists a BD ing. Proof. Consider the weight w = (wn)∞ n=1 defined by w = (1, 2, 2−1, 2, 2, 2−2, . . . , 2, . . . , 2 (cid:124) (cid:123)(cid:122) (cid:125) The weighted shift Bw is not mixing since (cid:81)n n , 2−n, . . . ). i=1 wi does not tend to infinity as n tends to infinity (see, e.g., Chapter 4 in [15]). It remains to show that Bd(AM ) = 1 for every M ≥ 1. Let M > 1 and n ∈ N such that 2n−1 < M ≤ 2n. If k > n(n + 1)/2 and s ≥ (n + 1) + (n + 2) + ··· + 2n = n(3n + 1)/2, then there is ls > 1 such that (ls − 1)n((ls + 1)n + 1)/2 ≤ s < (ls)n((ls + 2)n + 1)/2. An easy computation shows that we have AM ∩ [k, k + s] ≥ s − ls(n2 + n) > (l2 s/2 − ls − 1)n2 − lsn. Therefore, αs := lim inf k→∞ AM ∩ [k, k + s] ≥ (l2 s/2 − ls − 1)n2 − lsn, and thus Bd(AM ) = lim s→∞ αs s ≥ lim s→∞ s/2 − ls − 1)n2 − lsn (l2 s/2 + ls)n2 + lsn (l2 = 1. We conclude by Proposition 3.3. Proposition 4.4. There exists a BD-operator on (cid:96)1(Z+) which is not a D1-operator. Proof. Let An = [2, . . . , 2 n−times sider the weight sequence , 2−n], B1 = A1, Bn = [Bn−1, An, Bn−1], and con- (cid:124) (cid:123)(cid:122) (cid:125) (cid:125) (cid:125), A3, A1, A2, A1 (cid:123)(cid:122) (cid:123)(cid:122) (cid:124) (cid:125), A4, A1, A2, A1 (cid:123)(cid:122) (cid:124) (cid:124) (cid:125) (cid:125), A3, A1, A2, A1 (cid:123)(cid:122) (cid:123)(cid:122) (cid:125), . . . ). (cid:124) w = (A1, A2, A1 (cid:123)(cid:122) (cid:124) (cid:124) Since AM has bounded gaps for every M > 0, we have from Corollary 3.4 that Bw is topologically ergodic, i.e., it is a BD-operator. In view of Proposition 3.3, it now suffices to show that (cid:16){k ∈ N : d We first notice that Bn = 3 · 2n − n − 3 and βn := i=1 wi > 1}(cid:17) k(cid:89) (cid:12)(cid:12)(cid:12)(cid:8)k ≤ Bn : 16 < 1. wi = 1(cid:9)(cid:12)(cid:12)(cid:12) = 2n − 1. k(cid:89) i=1 Now we observe that(cid:81)k i=1 wi ≥ 1 for all k ∈ N. Therefore, we have (cid:111)(cid:17) (cid:16)(cid:110) k(cid:89) (cid:12)(cid:12)(cid:12)(cid:8)k ∈ [1, n] :(cid:81)k i=1 wi > 1(cid:9)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:8)k ≤ Bn + n + 1 :(cid:81)k n i=1 wi > 1(cid:9)(cid:12)(cid:12)(cid:12) wi > 1 k ∈ N : d = lim sup n = lim sup i=1 Bn + n + 1 n Bn − βn + n + 2 Bn + n + 1 = lim n 2 · 2n 3 · 2n − 2 = 2 3 < 1. = lim n Figure 1 below summarizes the results of this section. We remark that: • by Proposition 4.2 (1), there exists a BD1-operator which is not a D1-operator and a BD-operator which is not a D-operator; • by Proposition 4.2 (2), there exists a D1-operator which is not a D operator and a D-operator which is not a D-operator; 1- • by Proposition 4.2 (3), there exists a D 1-operator which is not a BD 1 and a D-operator which is not a BD-operator. On the other hand, by Proposition 4.4, there exists a • BD-operator which is not a BD 1-operator; • BD-operator which is not a D1-operator; • D-operator which is not a D 1-operator; • D-operator which is not a D1-operator; • D-operator which is not a D1-operator. 17 Figure 1: Densities and transitivity properties 5 Some special families In this section we study new classes of F -transitive operators given by families commonly used in Ramsey Theory. For a rich source on these families see [16]. For instance, we will consider the families of ∆-sets and of IP-sets, as well as their dual families. ∆ = {A ⊆ Z+ : (B − B) ∩ Z+ ⊆ A, for some infinite subset B of Z+} IP = {A ⊆ Z+ : ∃(xn)n ⊆ N with (cid:88) xn ∈ A, ∀F ⊂ Z+ finite}. n∈F The families ∆∗ and IP∗ are filters since ∆ and IP are partition regular. In addition, we have I∗ (cid:36) ∆∗ (cid:36) IP∗ (cid:36) S I∗ (cid:36) PS∗ (cid:36) S, (5.1) see [6] for details. In linear dynamics, some of the widely studied classes are the mixing and weakly mixing operators. As we already mentioned, an operator T is mixing if and only if it is an I∗-operator and T is weakly mixing 18 if and only if T is a T-operator. We recall that the class of TS-operators coincides with the class of topologically ergodic operators by Lemma 2.3 (see also the exercises in [15, Chapter 2]). Moreover, since TS = PS∗ and TS is a filter, we know that PS∗ is partition regular (Lemma 2.1). With the help of Proposition 2.7 applied to F = PS we can therefore complete the picture. Proposition 5.1. Let T ∈ L(X), where X is a separable F -space. The following are equivalent: (1) T is a topologically ergodic operator; (2) T is a hereditarily TS-operator; (3) T is a TS-operator; (4) T is a hereditarily PS-operator; (5) hcA := {x ∈ X : {T nx : n ∈ A} = X} is a dense (Gδ) set in X for any A ∈ PS. We will distinguish different classes of F -operators by means of Propo- sition 3.3. Given a family F , the following are two standard ways to induce shift-invariant families (cid:91) (cid:92) k∈Z k∈Z F+ := (F + k) F• := (F + k), where F + k := {A ⊂ Z+ : ∃B ∈ F with (B + k) ∩ Z+ ⊂ A}, k ∈ Z. We have (cid:102)F ⊂ F• ⊂ F ⊂ F+. Moreover, for any A ⊆ Z+ we have A ∈ (F ∗)• if and only if A ∈ (F+) ∗ . (5.2) It is well-known that ∆ and IP are not shift invariant, while PS is. Also, if F = ∆, IP or PS and G = F or F+ then A ∈ G ∗ if and only if Z+ \ A /∈ G , (5.3) since G is partition regular. Proposition 5.2. Every F -operator is an F•-operator. Proof. Let U, V ∈ U(X) and k ≥ 0. We have N (U, T −kV ) + k ⊂ N (U, V ). Moreover, since X has no isolated points, by transitivity we can find non- empty open sets U(cid:48) ⊂ U and V (cid:48) ⊂ V such that N (T −kU(cid:48), V (cid:48)) ⊂ [k, +∞[. Thus we have N (T −kU(cid:48), V (cid:48))(cid:1) − k ⊂ N (U, V ). We can conclude that every F -operator is an F•-operator. 19 We next compare the notions of mixing operator, ∆∗-operator, IP∗- operator and topologically ergodic operator. Proposition 5.3. There exists a topologically ergodic weighted backward shift on X = c0(Z+) or (cid:96)p(Z+), 1 ≤ p < ∞, which is not an IP∗-operator. Proof. Consider the set (cid:110)(cid:88) 22n : F finite set of N(cid:111) . B = n∈F Clearly B ∈ IP and thus Z+ \ B /∈ IP∗ by (5.3). Let (bn) be the increasing enumeration of B. We define the weight w = (wm)∞ m=1 as follows 1 1 wb2 , 2, . . . , 2, , 2, . . . , 2, 1 (cid:124) (cid:123)(cid:122) (cid:125) 2b3−b2−1 , 2, . . . ). (5.4) (cid:124) (cid:123)(cid:122) (cid:125) 2b2−b1−1 w = (2, . . . , 2, 2b1−1 wb1 (cid:124)(cid:123)(cid:122)(cid:125) Now, A1 := {n ≥ 1 : (cid:81)n i=1 wi > 2j} = Z+ \(cid:16)(cid:83)j 1 :(cid:81)n i=1 wi > 1} = Z+ \ B, hence Bw is not an IP∗- operator by Propositon 3.3. On the other hand, it is easy to see that B /∈ PS. Then (B +i) /∈ PS for every i ≥ 0, since PS is shift invariant. Hence, by (5.3) the set Z+ \ (B + i) ∈ PS∗ for every i ≥ 0. Now observe that A2j := {n ≥ i=0(Z+ \ (B + i)) ∈ PS∗, since PS∗ is a filter. Hence Bw is a PS∗-operator, or equivalently a topologically ergodic operator, by Proposition 3.3. =(cid:84)j i=0 B + i (cid:17) wb3 Proposition 5.4. There exists a weighted backward shift operator on X = c0(Z+) or (cid:96)p(Z+), 1 ≤ p < ∞, which is an IP∗-operator but not a ∆∗- operator. Proof. Let B be an infinite subset of N with unbounded gaps and let (bn)n be an increasing enumeration of B. So there exists an increasing sequence (nk) such that 1 :(cid:81)n (5.5) m=1 given by (5.4). As before {n ≥ Consider the weight sequence w = (wm)∞ i=1 wi > 1} = Z+ \ B, so it would be desirable that B ∈ ∆ and thus that Z+ \ B /∈ ∆∗ since this would imply that Bw is not a ∆∗-operator. On the other hand, it can be verified that for every M > 0 and j ∈ N there exists a finite subset F of Z such that AM,j = Z+ \ (∪i∈F B + i). Hence, in order to conclude that Bw is an IP∗-operator, by Proposition 3.3 bnk+1 − bnk → ∞. and condition (5.3) we need to verify(cid:91) (B + i) /∈ IP i∈F (5.6) for any finite subset F of Z. Now, since IP is partition regular, condition (5.6) is obtained if B /∈ IP+ and this in turn is equivalent to Z+ \ B ∈ 20 (cid:0)IP∗(cid:1) [6, Theorem 2.11 (1)] ensures the existence of a set E ∈ (cid:0)IP∗(cid:1) not (cid:83) (cid:0)∆∗ + n(cid:1)-set in N, hence not ∆∗-set. In addition, Z+ \ E has • by (5.3) and (5.2). Now, an obvious modification in the proof of • which is unbounded gaps. Setting B = Z+ \ E we are done. n∈Z+ Evidently, every mixing operator is a ∆∗-operator but the converse is not true. Proposition 5.5. There exists a ∆∗-weighted backward shift on c0(Z+) or (cid:96)p(Z+), 1 ≤ p < ∞, which is not mixing. Proof. Let B = {bi : b1 = 2, bi+1 = bi + i + 2, i ∈ N}. Consider the weight sequence w = (wm)∞ m=1 given by (5.4), so we have w =(cid:0)2, 2−1, 2, 2, 2−2, 2, 2, 2, 2−3, . . .(cid:1). i=1 wi does not tend to infinity as n tends to infinity. On the other hand, it can be verified that for every M > 0 and j ∈ N there exists a finite subset F of Z such that AM,j = Z+ \ (∪i∈F B + i). Hence, in order to conclude that Bw is a ∆∗-operator, by i∈F B + i /∈ ∆, for every finite subset F of Z. So, let F be a finite subset of Z with N = maxa,b∈F a − b. Suppose that i∈F B + i is a ∆-set. Then, there exists an increasing sequence (dm)m such We know that Bw is not mixing since(cid:81)n Proposition 3.3 and condition (5.3) we need to verify(cid:83) (cid:83) i∈F B +i = ∆(cid:0)(dm)m that(cid:83) of (dm)m defined by ∆(cid:0)(dm)m which means that the distance dj2 − dj1 between elements of (cid:83) the way in which B was defined. We thus conclude that(cid:83) (cid:1), where ∆(cid:0)(dm)m (cid:1) denotes the set of differences (cid:1) = {dj − di : 1 ≤ i < j}. Fix dj1, dj2(j1 < j2) such that dj2 − dj1 > N. Then for each m ∈ N we have dj2 − dj1 = (djm − dj1) − (djm − dj2) , i∈F B + i is attained infinitely many times, which is not the case taking into account i∈F B +i /∈ ∆. 5.1 Connection with A-hypercyclicity In this subsection we investigate the connection between the classes of hypercyclic operators considered throughout this work and the notion of A-hypercyclicity studied in [7]. Given a family A on Z+, an operator T ∈ L(X) is called A-hypercyclic if there exists x ∈ X such that N (x, V ) ∈ A for each V in U(X). Such a vector x is called an A-hypercyclic vector for T . When A = D, the operator T is called frequently hypercyclic. This class was introduced by Bayart and Grivaux in [3], [2]. When A = D, the operator T is called U-frequently hypercyclic; this class was introduced by Shkarin [21]. When A = BD, the operator T is called reiteratively hypercyclic [18] (see a detailed study in [7]). 21 The frequently hypercyclic operators constitute by far the most exten- sively studied class of operators amongst the three classes mentioned above. Clearly any frequently hypercyclic operator is an U-frequently hypercyclic operator, which in turn is reiteratively hypercyclic. The hierarchy between frequently hypercyclic and U-frequently hypercyclic operators as well as a full characterization for weighted shift operators have been established by Bayart and Ruzsa [5]. A complementary study of this kind, taking into account reiterative hypercyclicity can be found in [7]. In particular, we already know that there exists a mixing weighted shift which is not reiteratively hypercyclic as shown in [7]. On the other hand, there exists a frequently hypercyclic (hence reiteratively hypercyclic) op- erator which is not mixing, see [1]. Reiteratively hypercyclic operators are topologically ergodic [7, 13]. One can therefore wonder whether any reiter- atively hypercyclic operator is a ∆∗-operator or an IP∗-operator. Proposition 5.6. Let T ∈ L(X) be a reiteratively hypercyclic operator. Then N (U, V ) ∈ (cid:92) (cid:16) (cid:17) ∆∗ + t , t∈N (U,V ) for every U, V non-empty open sets in X. Proof. Let U, V ∈ U(X) and n ∈ N (U, V ). The set Un = U ∩ T −nV is a non-empty open set. Since T is reiteratively hypercyclic, there exists x ∈ X such that Bd (N (x, Un)) > 0. Let s1, s2 ∈ N (x, Un). We have T s1−s2+n(T s2x) = T n(T s1x) ∈ V. In other words, N (x, Un) − N (x, Un) + n ⊆ N (U, V ). (5.7) The desired result then follows from Theorem 3.18 in [11], which implies that A − A ∈ ∆∗ whenever A ∈ BD. The family ∆∗ is not shift invariant (2N := {2n : n ∈ N} ∈ ∆∗ while 2N + 1 /∈ ∆∗). Hence, we cannot deduce from Proposition 5.6 that every reiteratively hypercyclic operator is a ∆∗-operator. In fact, we are not able to answer in general the following question: is any reiteratively hypercyclic operator either a ∆∗-operator or an IP∗-operator? However we can show that the answer is yes if we consider bilateral or unilateral weighted shifts. Proposition 5.7. If Bw is reiteratively hypercyclic on X = (cid:96)p(Z), 1 ≤ p < ∞, or X = c0(Z), then Bw is an ∆∗-operator. In order to prove Proposition 5.7, we first state two lemmas. The first one directly follows from Proposition 5.6. 22 Lemma 5.8. Let U, V non-empty open sets in X such that U ∩ V (cid:54)= ∅, if T is reiteratively hypercyclic on X then N (U, V ) ∈ ∆∗. Let X = (cid:96)p(Z), 1 ≤ p < ∞, or c0(Z). The second lemma will rely on the non-empty open sets UR,j defined for every R > 1 and every j ∈ Z by UR,j = {U ∈ U(X) : xj > ,∀x ∈ U}. 1 R In particular, we remark that if M R > 1 then B((M + 1)ej; where B(y; ) stands for the open ball centered at y with radius . Lemma 5.9. Let M > 0, j ∈ Z and R > 1 such that M R > 1. Suppose M R ) ∈ UR,j, there exists U ∈ UR,j such that for any non-empty open subset (cid:101)U of U it holds N ((cid:101)U , B((M + 1)ej; M R )) ∈ ∆∗. Then AM,j ∈ ∆∗ and ¯AM,j ∈ ∆∗. 1 1 Proof. Let (z(m))m be a dense set in X such that z(m) = (z(m)1, . . . , z(m)m, 0 . . . ) 1 1 1 1 M R ). wi i=j+1 1 M R (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < xj+r − (M + 1) and Um = B(z(m); 1/m). Let U ∈ UR,j such that for any non-empty open M R )) ∈ ∆∗. Then there exists M R )) ∈ ∆∗. Pick m such that Um ⊂ U and hence N (Um, B((M + 1)ej; wx ∈ r ∈ N (Um, B((M + 1)ej; B((M + 1)ej; subset (cid:101)U of U we have N ((cid:101)U , B((M + 1)ej; Then, we have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) M R )) with r > m and x ∈ Um such that Br (cid:33) (cid:32) t+r(cid:89) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) > (cid:32) j+r(cid:89) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r(cid:89) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) > M, and for every t (cid:54)= j xt+r (cid:33) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r(cid:89) On the other hand, by (5.9), we get(cid:81)j j(cid:89) j(cid:89) wi+j where the first inequality follows since r > m. We conclude that N (Um, B((M + 1)ej; M R )) \ {1 . . . m} ⊆ AM,j and thus AM,j ∈ ∆∗. i=j−r+1 wixj < 1 By (5.8) we get, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < M R, hence 1 . M R i=1 i=1 wi+jxr+j wi i=t+1 (5.8) (5.9) 1 We deduce that(cid:81)j wi 1 R < wixj < 1 . M R i=j−r+1 i=j−r+1 wi < 1 i=j−r+1 M and thus ¯AM,j ∈ ∆∗. 23 Proof of Proposition 5.7 Suppose Bw is not a ∆∗-operator on X, then by Proposition 3.1, there exists M > 0 and j ∈ Z such that AM,j /∈ ∆∗ or ¯AM,j /∈ ∆∗. Let R > 1 such that M R > 1. By Lemma 5.9, it follows that ∀U ∈ UR,j ∃(cid:101)U ⊆ U : N ((cid:101)U , B((M + 1)ej; )) /∈ ∆∗. M R ) ∈ UR,j, we can consider U = B((M + 1)ej; there thus exists (cid:101)U ⊆ U such that N ((cid:101)U , U ) /∈ ∆∗, which by Lemma 5.8, is Since B((M + 1)ej; 1 M R ) and 1 M R 1 not possible if Bw is reiteratively hypercyclic. This concludes the proof of Proposition 5.7. Analogously, we have the following result for unilateral weighted shifts. Proposition 5.10. If Bw is reiteratively hypercyclic on X = (cid:96)p(Z+), 1 ≤ p < ∞, or on X = c0(Z+), then Bw is a ∆∗-operator. Proposition 5.11. There exists a reiteratively hypercyclic operator on c0(Z+) which is not a D1-operator. Proof. Let Bw be the weighted shift on c0(Z+) given by  k−1(cid:89) 2 w−1 ν ν=1 1 wk = if k ∈ S if k ∈ (S + 1)\S otherwise. where S :=(cid:83) j,l≥1]l10j − j, l10j + j[. It is shown in [7, Theorem 17] that Bw is reiteratively hypercyclic and that k(cid:89) d({k ∈ N : wi ≥ 2j}) → 0. (cid:81)k In particular, we deduce that there exists j ≥ 1 such that d({k ∈ N : i=1 wi ≥ 2j}) < 1 and in view of Proposition 3.3, we can conclude that Bw is not a D1-operator. i=1 Figure 2 summarizes what we know after this work. 24 Figure 2: Known relations We recall the following questions that remain open. Question 5.12. Does there exist a D-operator which is not a BD1-operator? In other words, does there exist T ∈ L(X) being a D-operator but not weakly mixing? Note that if it were the case, then such operator T must not be weighted shift. Question 5.13. Is any reiteratively hypercyclic operator an ∆∗-operator or an IP∗-operator? References [1] C. Badea and S. Grivaux. Unimodular eigenvalues, uniformly dis- tributed sequences and linear dynamics. Adv. Math. 211 (2007), no. 2, 766-793. [2] F. Bayart and S. Grivaux. Invariant Gaussian measures for operators on Banach spaces and linear dynamics, Proc. Lond. Math. Soc. (3) 94 (2007), no.1, 181-210. 25 [3] F. Bayart and S. Grivaux. Frequently hypercyclic operators. Trans. Amer. Math. Soc. 358 (2006), no.11, 5083-5117. [4] F. Bayart and E. Matheron. Dynamics of linear operators. Cambridge Tracts in Mathematics, 179. Cambridge University Press, Cambridge (2009). [5] F. Bayart and I. Ruzsa. Difference sets and frequently hypercyclic weighted shifts. Ergodic Theory Dynam. Systems 35 (2015), no.3, 691- 709. [6] V. Bergelson and T. Downarowicz. Large sets of integers and hierarchy of mixing properties of measure preserving systems. Colloq. Math. 110 (2008), no. 1, 117-150. [7] J. Bès, Q. Menet, A. Peris and Y. Puig. Recurrence properties of hy- percyclic operators. Math. Ann. 366 (2016), no. 1-2, 545–572. [8] J. Bès and A. Peris, Hereditarily hypercyclic operators, J. Funct. Anal. 167 (1999), 94–112. [9] A. Bonilla and K.-G. Grosse-Erdmann, Upper frequent hypercyclicity and related notions, Preprint 2016. ArXiv:1601.07276v1. [10] H. Furstenberg, Disjointness in ergodic theory, minimal sets, and a problem in Diophantine approximation, Math. Systems Theory 1(1967), 1–49. [11] H. Furstenberg, Recurrence in Ergodic Theory and Combinatorial Number Theory, Princeton University Press, Princeton, NJ, 1981. [12] S. Grivaux, Hypercyclic operators, mixing operators and the bounded steps problem, J. Op. Theory 54:1 (2005), 147–168. [13] K.-G. Grosse-Erdmann and A. Peris, Frequently dense orbits, C. R. Math. Acad. Sci. Paris 341 (2005), 123–128. [14] K.-G. Grosse-Erdmann and A. Peris, Weakly mixing operators on topo- logical vector spaces, Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Math. RACSAM 104 (2010), no. 2, 413–426. [15] K.-G. Grosse-Erdmann and A. Peris. Linear chaos. Universitext. Springer, London, 2011. [16] N. Hindman and D. Strauss. Algebra in the Stone-Čech compactifi- cation. Theory and applications. Expositions in Mathematics, 27. De Gruyter, Berlin (1998). [17] Q. Menet. Linear chaos and frequent hypercyclicity. Trans. Amer. Math. Soc., to appear. doi:10.1090/tran/6808. 26 [18] A. Peris. Topologically ergodic operators. Function Theory on Infi- nite Dimensional Spaces IX, Madrid (2005). http://www.mat.ucm.es/ ~confexx/web_confe_05/index.htm [19] A. Peris and L. Saldivia. Syndetically hypercyclic operators. Integr. Equ. Oper. Theory 51 (2005), 275-281. [20] H. N. Salas, Hypercyclic weighted shifts, Trans. Amer. Math. Soc. 347 (1995), 993–1004. [21] S. Shkarin. On the spectrum of frequently hypercyclic operators. Proc. Amer. Math. Soc. 137 (2009), no. 1, 123-134. 27
1601.02035
1
1601
2016-01-08T21:40:49
On isomorphically polyhedral $\mathcal L_\infty$-spaces
[ "math.FA" ]
We show that there exist $\mathcal L_\infty$-subspaces of separable isomorphically polyhedral Lindenstrauss spaces that cannot be renormed to be a Lindenstrauss space.
math.FA
math
ON ISOMORPHICALLY POLYHEDRAL L∞-SPACES JES ´US M. F. CASTILLO AND PIER LUIGI PAPINI Abstract. We show that there exist L∞-subspaces of separable isomorphically poly- hedral Lindenstrauss spaces that cannot be renormed to be a Lindenstrauss space. 1. Isomorphically polyhedral spaces A Banach space is said to be polyhedral if the closed unit ball of every finite dimen- sional subspace is the closed convex hull of a finite number of points. Polyhedrality is a geometrical notion: c0 is polyhedral while c is not. It is also an hereditary notion: every subspace of a polyhedral space is polyhedral. The isomorphic notion associated with polyhedrality is: A Banach space is said to be isomorphically polyhedral if it admits a polyhedral renorming. The simplest examples of isomorphically polyhedral spaces are the C(α) spaces for α an ordinal, and their subspaces. In [5] we surveyed what is known about polyhedral L∞-spaces, which can be summarized as follows: (1) There are polyhedral spaces which are not L∞: indeed, any non L∞ subspace of c0(Γ) -- recall from [11] that subspaces of c0(Γ) are L∞-spaces if and only if they are isomorphic to c0(I). (2) There are Lindenstrauss spaces not polyhedral: C[0, 1]. (3) A result of Fonf [8] asserts that preduals of ℓ1 are isomorphically polyhedral. (4) Fonf informed us [9] that the result fails for ℓ1(Γ): Kunen's compact K provides, under CH, a scattered, non metrizable, compact so that C(K) space has the rare property that every uncountable set of elements contains one that belongs to the closure of the convex hull of the others. And this property was used by Jim´enez and Moreno [13] to show that every equivalent renorming of C(K) has only a countable number of weak*-strongly exposed points. Thus, no equivalent renorming can be polyhedral (see [10]). At the same time C(K)∗ = ℓ1(Γ) since K is scattered. (5) The trees T for which C(T ) is isomorphically polyhedral are characterized in [10]. Thus, there are scattered compact K (not depending on CH as it occurs with Kunen's compact) such that C(K) is not isomorphically polyhedral. This research was partially supported by project MTM2013-45643-C2-1-P. The research of the sec- ond author has been partially supported by GNAMPA of the Instituto Nazionale di Alta Matematica. 1 2 JES ´US M. F. CASTILLO AND PIER LUIGI PAPINI Fonf [9] asked [5, Section 4, problem 5] whether isomorphically polyhedral L∞-spaces are isomorphically Lindenstrauss. The purpose of this note is to show that the answer is no. 2. Preliminaries A Banach space X is said to be an L∞,λ-space if every finite dimensional subspace F of X is contained in another finite dimensional subspace of X whose Banach-Mazur distance to the corresponding space ℓn ∞ is at most λ. The space X is said to be an L∞- space if it is an L∞,λ-space for some λ. The basic theory and examples of L∞-spaces can be found in [18, Chapter 5]. A Banach space X is said to be a Lindenstrauss space if it is an isometric predual of some space L1(µ). Lindenstrauss spaces correspond to L∞,1+-spaces. A Lindenstrauss space is an L∞,1-space if and only if it is polyhedral (i.e., the unit ball of every finite dimensional subspace is a polytope) [18, p.199]. A Banach space X is said to have Pe lczy´nski's property (V ) if each operator de- fined on X is either weakly compact or an isomorphism on a subspace isomorphic to c0. Pe lczy´nski shows in [19] that C(K)-spaces enjoy property (V ), and Johnson and Zippin [14] that Lindenstrauss spaces also have (V ). Let α : A → Z and β : B → Z be operators acting between Banach spaces. the pull-back space PB is defined as PB = PB(α, β) = {(a, b) ∈ A ⊕∞ B : α(a) = β(b)}. It has the property of yielding a commutative diagram (1) PB ′α y B ′β −−−→ A  α y −−−→ Z β in which the arrows after primes are the restriction of the projections onto the corre- sponding factor. Needless to say (1) is minimally commutative in the sense that if the operators ′′β : C → A and ′′α : C → B satisfy α ◦ ′′β = β ◦ ′′α, then there is a unique operator γ : C → PB such that ′′β = ′βγ and ′′β = ′βγ. Clearly, γ(c) = (′′β(c), ′′α(c)) and kγk ≤ max{k′′αk, k′′βk}. Quite clearly ′α is onto if α is. As a consequence of this, if one has an exact sequence (2) 0 −−−→ Y ı−−−→ X π−−−→ Z −−−→ 0 and an operator u : A → Z then one can form the pull-back diagram of the couple (π, u): 0 −−−→ Y ı−−−→ X ′ux   PB π−−−→ Z −−−→ 0 u x   ′π−−−→ A ON ISOMORPHICALLY POLYHEDRAL L∞-SPACES 3 Recalling that ′π is onto and taking j(y) = (0, ı(y)), it is easily seen that the following diagram is commutative: (3) 0 −−−→ Y (cid:13) (cid:13) (cid:13) 0 −−−→ Y ı−−−→ X ′ux   −−−→ PB j π−−−→ Z −−−→ 0 u x   ′π−−−→ A −−−→ 0 Thus, the lower sequence is exact, and we shall refer to it as the pull-back sequence. The well-known (see e.g., [2]) splitting criterion is: the pull-back sequence splits if and only if u lifts to X; i.e., there is an operator U : A → X such that πU = u. 3. An isomorphically polyhedral L∞-space that is not Lindenstrauss Theorem 1. There is a separable isomorphically polyhedral L∞ space that is not iso- morphically Lindenstrauss. Moreover, it is a subspace of an isomorphically polyhedral Lindenstrauss space. Proof. We need to recall from [1] the existence of nontrivial exact sequences 0 −−−→ C(ωω) −−−→ Ω q −−−→ c0 −−−→ 0 in which the quotient map q is strictly singular. This fact makes Ω fail Pe lczy´nski's property (V ). Since Lindenstrauss spaces share with C(K)-spaces Pe lczy´nski's prop- erty (V), the space Ω is not isomorphic to a Lindenstrauss space. Of course it is an L∞-space since this is a 3-space property. Thus, our purpose is to show that there is an Ω as above that is isomorphically polyhedral. We recall from [1, Section3] the parameter ρN (c0), defined as the the least constant such that if T : c0 → ℓ∞(ωN ) is a bounded linear operator such that dist(T x, C(ωN )) ≤ kxk for all x ∈ c0 then there is a linear map L : c0 → C(ωN ) with kT − Lk ≤ ρN (c0). Theorems 3.1 and Lemma 3.2 in [1] show that lim ρN (c0) = +∞. Now we need a spe- cific choice for each N: this is provided by [1, Prop. 4.6]: there is a bounded operator TN : c0 → ℓ∞(ωN ) so that dist(TN x, C(ωN )) ≤ kxk for all x ∈ c0 but such that if E ⊂ c0 is a subspace of c0 almost isometric to c0 then ρN (c0) ≤ 2kTN − Lk for any linear map L : c0 → C(ωN ). Let, for each N, a linear continuous operator TN : c0 → ℓ∞(ωN ) as above. We form the twisted sum space C(ωN ) ⊕TN c0 = (cid:0)C(ωN ) × c0, k · kTN (cid:1) endowed with the norm k(h, x)kTN = max{kh − TN xk, kxk}. This yields an exact sequence 0 −−−→ C(ωN ) iN−−−→ C(ωN ) ⊕TN c0 qN−−−→ c0 −−−→ 0 4 JES ´US M. F. CASTILLO AND PIER LUIGI PAPINI with embedding iN (f ) = (f, 0) and quotient map qN (f, x) = x. The identity map id : C(ωN ) ⊕TN c0 −→ C(ωN ) ⊕∞ c0 is an isomorphism since kTN k−1k(f, x)kTN ≤ k(f, x)k∞ ≤ kTN kk(f, x)kTN and therefore the space C(ωN ) ⊕TN c0 is isomorphically polyhedral. We need now to use the main result in [7] asserting that in a separable isomorphically polyhedral space every norm can be approximated by a polyhedral norm. Let k · kPN be a polyhedral norm in C(ωN ) ⊕TN c0 that is 2-equivalent to k · kTN . The sequence (3) splits, but the norm of the projection goes to infinity with N: Indeed, if P : C(ωN ) ⊕TN c0 → C(ωN ) is a linear continuous projection then P has to have the form P (f, x) = (f − Lx, 0), where L : c0 → C(ωN ) is a certain linear map. Thus, if x ∈ c0 is a norm one element, one gets P (TN x, x) = (TN x − Lx, 0) and thus kTN x − Lxk ≤ kP kkxk, hence kTN − Lk ≤ kP k. The choice of TN forces limN→∞ inf kP k = +∞. Therefore, the c0-sum 0 −−−→ c0(C(ωN )) −−−→ c0(C(ωN ) ⊕PN c0) (qN ) −−−→ c0(c0) −−−→ 0 cannot split. The space c0(C(ωN ) ⊕PN c0) is isomorphically polyhedral as any c0-sum of polyhedral spaces [12]. We now define a suitable operator ∆ so that when making the pull-back diagram 0 −−−→ c0(C(ωN )) −−−→ c0(C(ωN ) ⊕PN c0) (qN ) −−−→ c0(c0) −−−→ 0 (cid:13) (cid:13) (cid:13) 0 −−−→ c0(C(ωN )) −−−→ δ x   Ω x  ∆  c0 −−−→ 0 −−−→ q the map q is strictly singular. That prevents Ω from being Lindenstrauss under any equivalent renorming. Pick as ∆ the diagonal operator c0 → c0(c0) induced by the scalar sequence (ρN (c0)−1/2) ∈ c0; i.e., ∆(x) = (ρN (c0)−1/2x)N . Assume that q is not strictly singular. Then, there is a subspace E of c0 and a linear bounded map V : E → Ω so that qV = ∆E. By the c0 saturation and the distortion properties of c0, there is no loss of generality assuming that E is an almost isometric copy of c0. By the commutativity of the diagram (qN )δV = ∆E, which in particular means that qN δV (e) = ρN (c0)−1/2e for all e ∈ E. This means that the map δV has on E the form (LN e, ρN (c0)−1/2e)N where LN : E → C(ωN ) is a linear map; by continuity, there is a constant M so that k(LN e, ρN (c0)−1/2e)k ≤ Mkek, which means kLN e − TN ρN (c0)−1/2ek ≤ Mkek ON ISOMORPHICALLY POLYHEDRAL L∞-SPACES 5 and thus kρN (c0)1/2LN − TN k ≤ MρN (c0)1/2. This contradicts the fact that E = c0, the definition of ρN (c0) and the choice of TN . To conclude the proof, the definition of pull-back space implies that Ω is actually a (cid:3) subspace of c0(C(ωN ) ⊕PN c0) ⊕∞ c0, hence isomorphically polyhedral. Since c0(C(ωN )) ≃ C(ωN), the space Ω above yields a twisted sum 0 −−−→ C(ωω) −−−→ Ω q −−−→ c0 −−−→ 0 in which q is strictly singular. The dual sequence 0 −−−→ ℓ1 −−−→ Ω∗ −−−→ ℓ1 −−−→ 0 necessarily splits and thus Ω∗ can be renormed to be ℓ1, although Ω cannot be endowed with an equivalent norm · so that (Ω, · )∗ = ℓ1. Moreover, Ω is actually a subspace of the isomorphically polyhedral Lindenstrauss space c0(C(ωN ) ⊕PN c0) ⊕ c0. 4. An isomorphically polyhedral L∞ space that is not a Lindenstrauss-Pe lczy´nski space We show now that one can produce an L∞-variation of Ω still farther from Linden- strauss spaces. Lazar [15] and Lindenstrauss [16] showed that Lindenstrauss polyhedral spaces X enjoy the property that compact X-valued operator admit equal norm ex- tensions. In [3], the authors introduce the Lindenstrauss-Pe lczy´nski spaces (in short LP-spaces) as those Banach spaces E such that all operators from subspaces of c0 into E can be extended to c0. The spaces are so named because Lindenstrauss and Pe lczy´nski first proved in [17] that C(K)-spaces have this property. Lindenstrauss spaces have also the property (see [17, 6]) as well as L∞-spaces not containing c0 [3] and, of course, all their complemented subspaces. The construction of the space Ω above has been modified in [4] to show that for every subspace H ⊂ c0 there is an exact sequence 0 −−−→ C(ωω) −−−→ ΩH −−−→ c0 −−−→ 0 in which the space ΩH is not a Lindenstrauss-Pe lczy´nski space [17]; more precisely, there is an operator H → ΩH that cannot be extended to the whole c0. Proposition 1. There is an isomorphically polyhedral L∞-space that is not an LP- space. Proof. Consider the exact sequence 0 → C(ωω) → Ω → c0 → 0 with strictly singular quotient constructed above. Since every quotient of c0 is isomorphic to a subspace of c0, we can consider that there is an embedding uH : c0/H → c0. The pull-back p sequence 0 → C(ωω) → PH → c0/H → 0 also has strictly singular quotient map. We form the commutative diagram 6 JES ´US M. F. CASTILLO AND PIER LUIGI PAPINI 0 x   0 −−−→ C(ωω) −−−→ PH x   k 0 −−−→ C(ωω) −−−→ ΩH −−−→ −−−→ c0/H −−−→ 0 x  t  c0 −−−→ 0 0 x   = p Q = jx   H x  0 x  i  H x  0 to show, exactly as in [17] that ΩH is not an LP-space since j cannot be extended to c0 through i. The space ΩH has been obtained from a pull-back diagram 0 −−−→ c0(C(ωN )) −−−→ Ω −−−→ c0 −−−→ 0 (cid:13) (cid:13)(cid:13) x  0 −−−→ C(ωω) −−−→ PH x   k 0 −−−→ C(ωω) −−−→ ΩH −−−→ p Q uH x  −−−→ c0/H −−−→ 0 t x   c0 −−−→ 0, and thus it is a subspace of Ω ⊕ c0, hence isomorphically polyhedral. (cid:3) References [1] F. Cabello S´anchez, J.M.F. Castillo, N.J. Kalton and D.T. Yost, Twisted sums with C(K) spaces, Trans. Amer. Math. Soc. 355 (2003) 4523-4541. [2] J.M.F. Castillo and M. Gonz´alez. Three-space problems in Banach space theory, Lecture Notes in Math. 1667. Springer-Verlag, 1997. [3] J.M.F. Castillo, Y. Moreno and J. Su´arez. On Lindenstrauss-Pe lczy´nski spaces, Studia Math. 174 (2006) 213 -- 231. [4] J.M.F. Castillo, Y. Moreno and J. Su´arez, On the structure of Lindenstrauss-Pe lczy´nski spaces, Studia Math. 194 (2009) 105 -- 115. [5] J.M.F. Castillo and P.L. Papini, Hepheastus account on Trojanski's polyhedral war, Extracta Math. 29 (2014) 35 - 51. [6] J.M.F. Castillo and J. Su´arez, Extension of operators into Lindenstrauss spaces, Israel J. Math. 169 (2009) 1-27. [7] R. Deville, V. Fonf and P. Hajek, Analytic and polyhedral approximation of convex bodies in separable polyhedral Banach spaces, Israel J. Math. 105 (1998) 139 -- 154. ON ISOMORPHICALLY POLYHEDRAL L∞-SPACES 7 [8] V.P. Fonf, Massiveness of the set of extreme points of the dual ball of a Banach spaces and polyhedral spaces, Fuct. Anal. Appl. 12 (1978) 237-239. [9] V.P. Fonf, personal communications. [10] V.P. Fonf, A.J. Pallares, R.J. Smith and S. Troyanski, Polyhedral norms on non-separable Banach spaces, J. Funct. Anal. 255 (2008) 449 - 470. [11] G. Godefroy, N.J. Kalton and G. Lancien. Subspaces of c0(N) and Lipschitz isomorphisms Geom. Funct. Anal. 10 (2000) 798 - 820. [12] A.B. Hansen and N.J. Nielsen, On isomorphic classification of polyhedral preduals of L1, Preprint Series Aarhus University, 1973/74 No. 24. [13] M. Jim´enez Sevilla and Jos´e P. Moreno, Renorming Banach spaces with the Mazur intersection property J. Funct. Anal. 144 (1997) 486-504. [14] W.B. Johnson and M. Zippin, Separable L1 preduals are quotients of C(∆). Israel J. Math. 16 (1973) 198 -- 202. [15] A. J. Lazar, Polyhedral Banach spaces and the extensions of compact operators, Israel J. Math. 7 (1970) 357 - 364. [16] J. Lindenstrauss, On the extension of compact operators, Mem. Amer. Math. Soc. 48 (1964). [17] J. Lindenstrauss and A. Pe lczy´nski, Contributions to the theory of the classical Banach spaces, J. Funct. Anal. 8 (1971) 225 -- 249. [18] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces. Lecture Notes in Math. 338. Springer- Verlag, 1973. [19] A. Pe lczy´nski, On C(S)-subspaces of separable Banach spaces, Studia Math. 31 (1968) 513 -- 522. Departamento de Matem´aticas, Universidad de Extremadura, Avenida de Elvas, 06071-Badajoz, Spain E-mail address: [email protected] via Martucci 19, 40136 Bologna, Italia E-mail address: [email protected]
1508.05794
1
1508
2015-08-24T13:15:29
Graph Laplacians do not generate strongly continuous semigroups
[ "math.FA" ]
We show that for graph Laplacians $\Delta_G$ on a connected locally finite simplicial undirected graph $G$ with countable infinite vertex set $V$ none of the operators $\alpha\,\mathrm{Id}+\beta\Delta_G, \alpha,\beta\in\mathbb{K},\beta \ne 0$, generate a strongly continuous semigroup on $\mathbb{K}^V$ when the latter is equipped with the product topology.
math.FA
math
GRAPH LAPLACIANS DO NOT GENERATE STRONGLY CONTINUOUS SEMIGROUPS T. KALMES, C. SCHUMACHER Abstract. We show that for graph Laplacians ∆G on a connected locally finite simplicial undirected graph G with countable infinite vertex set V none of the operators α Id +β∆G, α, β ∈ K, β 6= 0, generate a strongly continuous semigroup on KV when the latter is equipped with the product topology. 1. Introduction Let G = (V, E) be a connected locally finite simplicial undirected graph with a countably infinite vertex set V and the set of edges E ⊆ V × V . Two vertices v, w ∈ V of G are adjacent, v ∼ w, if (v, w) ∈ E, i. e. (v, w) is an edge of G. Recall that a graph is locally finite if for every vertex v of G the number deg(v) of adjacent vertices is finite and that G is connected if for any pair of different vertices v, w ∈ V there is a path connecting v and w, i.e. a finite number of edges (v0, v1), . . . , (vn−1, vn) of G with v ∈ {v0, vn} and w ∈ {v0, vn}. The length of a path is the number of its edges. The graph G is simplicial if it does have no multiple edges, which we excluded implicitly by E ⊆ V × V , and no loops loops, i. e. (v, v) is not an edge of G for any vertex v. Finally, G is undirected if (w, v) is an edge of G whenever (v, w) is an edge of G. A Laplacian on G is a linear operator acting on the real valued functions on V , ∆G : KV → KV , defined by ∀ v ∈ V : ∆G f (v) = Xv∼w γv,w(f (v) − f (w)) with positive weights γv,w > 0. Common choices are γv,w = 1 or γv,w = 1/ deg(v). If we equip KV with the product topology it is clear that KV is a locally convex vector space on which ∆G is a continuous linear mapping. ∆G has been investigated e.g. in [1], [2], and [4]. The definitions and most basic results for semigroups on locally convex spaces X are the same as for Banach spaces, we refer to [5], [6], [7], [8]. A strongly continuous semigroup T on X is thus a morphism from the semigroup ([0, ∞), +) to that of continuous linear operators (L(X), ◦) such that all orbits t 7→ Tt(x), x ∈ X, are continuous. Although under rather weak assumptions on X many results from the Banach space setting carry over to strongly continuous semigroups of locally convex spaces there are some crucial differences, e. g. in general not every continuous linear oper- ator on X is the generator of a strongly continuous semigroup on X. The purpose of this short note is to proof the following theorem. For the defini- tion of a quasiadjacency matrix with respect to G see section 2. 2010 Mathematics Subject Classification. 47D06, 05C63, 46A04 Key words and phrases: graph Laplacians, strongly continuous semigroup on locally convex spaces. 1 2 T. KALMES, C. SCHUMACHER Theorem 1. For every λ ∈ K and each quasiadjacency matrix B with respect to G the continuous linear operator λ Id −B does not generate a strongly contin- uous semigroup on ω. In particular α Id +β∆G is not the generator of a strongly continuous semigroup on KV for any α, β ∈ K, β 6= 0. 2. Proof of Theorem 1 Enumeration of the vertices V = {vk; k ∈ N} of G with vk 6= vj for j 6= k ∈ N gives an isomorphism of KV onto KN. In order to keep notation simple, for a linear mapping A : KV → KV we denote by A also the linear operator on KN induced by the isomporphism between KV and KN. We equip KN with its usual Fr´echet space topology, i. e. the locally convex topol- ogy defined by the increasing fundamental system of seminorms (pk)k∈N given by ∀ f = (fj)j∈N ∈ KN : pk(f ) = k Xj=1 fj. As usual, we denote this Fr´echet space by ω. ω is the projective limit of the projec- tive spectrum of Banach spaces (Kn)n∈N with surjective linking maps πn m : Km → Kn, πn m(x1, . . . , xm) = (x1, . . . , xn), m ≥ n, thus ω is a quojection. Moreover, let πm : KN → Km, πm((xn)n∈N) = (x1, . . . , xm). Using that G is connected and locally finite, a straightforward calculation shows that A : ω → ω is continuous if and only if the inducing A : KV → KV has finite hopping range, i. e. if for every v ∈ V there is n ∈ N such that ∀ f, g ∈ KV : fUn(v) = gUn(v) =⇒ A(f )(v) = A(g)(v), where Un(v) is the n-neighbourhood of v, i. e. Un(v) is the union of {v} with the set of all endpoints of paths in G starting in v and of length not exceeding n. Obviously, ∆G has finite hopping range and is thus is a continuous linear mapping on ω. Finally, we call a matrix B = (bk,l) ∈ KN×N quasiadjacency matrix with respect to G if there is α ∈ K such that α bk,l ≥ 0 for all k, l and bk,l 6= 0 precisely when vk and vl are adjacent. Obviously, ∆G = Id −B for some quasiadjacency matrix B with respect to G. Moreover, for each quasiadjacency matrix B with respect to G and for every λ ∈ K the linear operator λ Id −B has finite hopping range and is thus continuous on ω. Since ω is a quojection, according to [3, Theorem 2] λ Id −B generates a strongly continuous semigroup if and only if (1) ∀ n ∈ N ∃ m ∈ N ∀ k ∈ N0, x ∈ ω : (cid:0)πm(x) = 0 ⇒ πn((λ Id −B)kx) = 0(cid:1). Denoting the unit vectors by el = (δk,l)k∈N we have for l ≥ 2 and k ∈ N π1(cid:0)(λ Id −B)kel(cid:1) = k Xj=1 (cid:18)k j(cid:19)(−1)jλk−j π1(Bjel). Since B is a quasiadjacency matrix with respect to G it follows that π1(Bj el) does not vanish precisely when there is a path of length j in G connecting v1 and vl. Now fix m and l > m so that πm(el) = 0. Moreover, set k := min{length j of a path connecting e1 and el}. Since G is connected, k ∈ N and π1(cid:0)(λ Id −B)kel(cid:1) = k Xj=1 (cid:18)k j(cid:19)(−1)jλk−j π1(Bjel) = (−1)kπ1(Bkel) 6= 0. GRAPH LAPLACIANS DO NOT GENERATE STRONGLY CONTINUOUS SEMIGROUPS 3 Since m is arbitrary, (1) is not fulfilled for n = 1 proving the theorem. (cid:3) References [1] T. Ceccherini-Silberstein, M. Coornaert, A note on Laplace operators on groups, in: Limits of graphs in groups theory and computer science, 37 -- 40, EPFL Press, Lausanne, 2009. [2] T. Ceccherini-Silberstein, M. Coornaert, J. Dodziuk, The surjectivity of the combinatorial Laplacian on infinite graphs, Enseign. Math. (2), vol. 58(1-2):125 -- 130, 2012. [3] L. Frerick, E. Jord´a, Enrique, T. Kalmes, J. Wengenroth, Strongly continuous semigroups on some Fr´echet spaces, J. Math. Anal. Appl. 412:121 -- 124, 2014. [4] T. Kalmes, A short remark on the surjectivity of the combinatorial Laplacian on infinite graphs, arXiv-preprint 1412.5854, 2014. [5] H. Komatsu, Semi-groups of operators in locally convex spaces, J. Math. Soc. Japan 16:230 -- 262,1964. [6] T. K¯omura, Semigroups of operators in locally convex spaces, J. Functional Analysis 2:258 -- 296, 1968. [7] S. ¯Ouchi, Semi-groups of operators in locally convex spaces, J. Math. Soc. Japan 25:265 -- 276, 1973. [8] K. Yosida, Functional analysis, Die Grundlehren der Mathematischen Wissenschaften, Band 123, Academic Press Inc., New York 1965. Technische Universitat Chemnitz, Fakultat fur Mathematik, 09107 Chemnitz, Germany E-mail addresses: [email protected], [email protected]
1307.4553
2
1307
2016-08-17T20:51:25
Tail Behaviour of Mexican Needlets
[ "math.FA" ]
In this paper we study the tail behaviour of Mexican needlets, a class of spherical wavelets introduced by Geller and Mayeli (2009). More specifically, we provide an explicit upper bound depending on the resolution level $j$ and a parameter $s$ governing the shape of the Mexican needlets
math.FA
math
Tail Behaviour of Mexican Needlets Claudio Durastantia,b,1 aDipartimento di Matematica, Università di Roma "Tor Vergata" bFakultät für Matematik, Ruhr Universität, Bochum Abstract In this paper we study the tail behaviour of Mexican needlets, a class of spherical wavelets introduced by Geller and Mayeli in [11]. More specifically, we provide an explicit upper bound depending on the resolution level j and a parameter s governing the shape of the Mexican needlets. Keywords: Wavelets, Mexican needlets, sphere, concentration properties, spatial localization. 2010 MSC: 42C40, 42C10, 42C15. 1. Introduction A lot of interest has recently been focussed on various forms of spherical wavelets (see, for example, [2, 4, 7, 14, 24, 31] and the references therein). This attention has also been fuelled by strong applied motivations, for instance in Astrophysics and Cosmology. More specifically, we refer to spherical Mexican hat wavelets (see, for instance, [20]), axisymmetric, directional, and steerable wavelets (cf., for example, [19, 21, 22, 31]), ridgelets and curvelets (see, for instance, [23, 28]). Many theoretical and applied papers have been concerned, in particular, with the so-called spherical needlets, which were introduced into the Functional Analysis literature by [24, 25]. Loosely speaking, the latter can be envisaged as a convolution of the spherical harmonics with a weight function which is smooth and compactly supported in the harmonic domain (more details will be given below). Localization properties in this framework were fully investigated by [24, 25]. Needlets have been recently generalized to various directions. Spin and mixed needlets were constructed over spin fiber bundles in [8, 9], respectively. Needlet-like wavelets were also developed on the unitary ball in [6, 26] and over compact manifolds in [15]. This framework has been also extended to allow for an unbounded support in the frequency domain by [11], see also [10, 12]; the latter construction is usually labelled as Mexican needlets. Examples of Email address: [email protected] (Claudio Durastanti) 1This research is supported by the ERC Grant n. 277742 Pascal and by the German DFG grant GRK 2131. Preprint submitted to Journal of Mathematical Analysis and ApplicationsSeptember 30, 2018 applications, mainly related to the study of Cosmic Microwave Background (CMB) radiation, can be found in [3, 5, 16, 18, 27]. As described in details below, Mexican needlets enjoy excellent localization properties in the real domain; in this paper, we investigate the relationship between the tail decay and the exact shape of the weight function. Indeed, the aim of this work is to provide analytic expressions to bound the tail behaviour in the real domain. We prove here that the tails are Gaussian up to a polynomial term, whose dependence on the choice of the kernel can be identified explicitly. More specifically, for any (multi)resolution level j, consider a partition of the sphere into a set (of cardinality Nj) of spherical subregions of area λjk and midpoint ξjk, k = 1, . . . , Nj. For any k, let ϑ := ϑjk (x) denote the geodesic distance between a generic coordinate x ∈ S2 and ξjk. For the scale parameter B > 1 and the shape parameter s ∈ N, we shall consider wavelet filters of the form Ψjk;s (ϑ) := pλjk 2π ∞Xℓ=0(cid:18)(cid:18) ℓ + 1 Bj (cid:19)(cid:19)2s 2 exp −(cid:18) ℓ + 1 Bj (cid:19)2! (2ℓ + 1) Pℓ (cos ϑ) , 2 where Pℓ (·) denotes the standard Legendre polynomial of degree ℓ. In Theorem 3.1, we shall be able to show that Ψjk;s (ϑ) ≤ CsBje−(2Bj ε)2(cid:0)1 +(cid:12)(cid:12)H2s(cid:0)Bjϑ(cid:1)(cid:12)(cid:12)(cid:1) , where Cs is a positive constant and H2s (·) identifies the Hermite polynomial of degree 2s. It is important to remark that in [11] the authors obtained an analogous expression for the n-dimensional sphere, limiting their investigation to the case of the shape parameter s = 1, which can be linked to the spherical Mexican hat wavelets (see Remark 3.3 and [27]). In this paper, we will extend this bound for any choice of s ∈ N. Our argument exploits a technique similar to the one used by Narcowich, Petrushev and Ward in [24] (see also [25] and the text- book [17, Section 13.3]). Furthermore, an analogous method was developed in [22] to establish concentration properties of spin directional wavelets. In our proof, we will also exploit the analytic form of the weight function to compute exactly its Fourier transform in terms of Hermite polynomials; this will also allow us to investigate explicitly the roles of the resolution level j and of the shape parameter s. We also establish bounds on the Lp-norms of the Mexican needlets, depending on the resolution level j and on the scale parameter B (see Corollary 3.2). Furthermore, in Proposition 2.1 we provide an explicit connec- tion between Mexican needlets with different shape by means of the spherical Laplacian operator. The plan of this paper is as follows. In Section 2 we recall the definition and some pivotal properties of Mexican needlets; in Section 3 we exploit our main theorem while Section 4 collects some auxiliary results. 2 2. The construction of Mexican needlets In this Section we shall review Mexican needlets, as developed by Geller and Mayeli, see [10, 11, 12]. As already mentioned and similarly to standard needlets (cf. [24, 25]), Mexican needlets can be viewed as a combination of Legendre polynomials weighted by a smooth window function. On one hand, recall the well-known decomposition of L2(cid:0)S2(cid:1), the space of the square-integrable functions over the sphere, given by L2(cid:0)S2(cid:1) =Mℓ≥0 Hℓ, where Hℓ is the space of the homogeneous polynomials of degree ℓ, spanned by the spherical harmonics {Yℓm, m = −ℓ, . . . , ℓ}. Therefore, spherical harmonics referred, for example, to the textbook [29]); here we just recall the so-called summation formula, i.e., for any ℓ ≥ 0 and for any x, y ∈ S2, provide an orthonormal basis for L2(cid:0)S2(cid:1). For further details, the reader is ℓXm=−ℓ Y ℓm (x) Yℓm (y) = 2ℓ + 1 4π Pℓ (hx, yi) , (1) where h·,·i denotes the geodesic distance over the sphere and Pℓ (·) is the Leg- endre polynomial of order ℓ, given by Pℓ (u) := 1 2ℓℓ! dℓ duℓ(cid:0)u2 − 1(cid:1)ℓ , u ∈ [−1, 1] , see, for example, [1, Chapter 22, Eqq. (22.1.6) and (22.2.10)]. On the other hand, consider the window (or weight) function fs : R → R+ fs (t) := t2se−t2 , t ∈ R, (2) for s ∈ N, so that, for any t ∈ R, we have that where as B → 1, B > 1, and f 2 s(cid:18) t Bj(cid:19) ≤ MB;s log B < ∞, 1 2 − 1 B 1 2 !2 2 !2 1 2 − 1 B 1 log B log B 1 2 − 1 B 1 1 2 − 1 B 1 2 !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  , 2 !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  , 0 < mB;s log B ≤ ∞Xj=−∞ mB;s := ηs1 − O(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) B MB;s := ηs1 + O(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) B ηs :=Z ∞ 0 f 2 s (tℓ) dt t = Γ (2s) 22s+1 , 3 see also see also [11]. In [11] (see also [10, 12]), it was proven that, for any given resolution level j ∈ (−∞,∞), there exists a finite set of measurable subregions of the sphere {Ejk}Nj k=1, of diameter ρjk and area λjk, such that ∪Nj k=1 Ejk = S2, Ejk1 ∩ Ejk2 = ∅, for any k1 6= k2, ρjk ≤ cBB−j, where cB > 0, B > 1. Each of these regions can be indexed by a point ξjk ∈ Ejk, typically chosen as its midpoint. For x, y ∈ S2, let Wj;s : S2× S2 → R be defined as follows Pℓ (hx, yi) . 2 fs(cid:18) ℓ + 1 Bj (cid:19)(cid:0)ℓ + 1 2(cid:1) Consider now the kernel operator Kj;s on L2(cid:0)S2(cid:1): Wj;s (x, y) :=Xℓ≥0 Kj;sF (x) =ZS2 2π For δ0 > 0 sufficiently small, it is shown in [11] that if, for any k and for any j , then, for ε > 0, Wj;s (x, y) F (y) dy, F ∈ L2(cid:0)S2(cid:1) . so that cBB−j < δ0, the area λjk is comparable with(cid:0)cBB−j(cid:1)2 (mB;s − ε)kFk2 λjk Kj;sF (ξjk)2 ≤ (MB;s + ε)kFk2 L2(S2) ≤ L2(S2) . ∞Xj=−∞ NjXk=1 Remark 2.1. Let us introduce preliminarily the notation a ≈ b if there exist c′, c′′ > 0 so that c′b ≤ a ≤ c′′b. In many practical applications, the set of Ejk, labelled by the pair (ξjk, λjk), can be identified with those evaluated by common packages such as HealPix (see for instance [13]), where, for any j, λjk ≈ 4πρ2 jk. Partitioning the sphere into Nj ≈ B2j regions Ejk, we have that µ(cid:16)∪Nj k=1Ejk(cid:17) = 4π = µ(cid:0)S2(cid:1) , where µ denotes the area. Hence, for the sake of simplicity, we consider, for any j, λjk ≈B−2j, k = 1, ..., Nj, Nj ≈B2j. Let us now define Ψjk;s : S2 → R as Ψjk;s (x) :=pλjkKj;s (x, ξjk) . 4 (3) An alternative form of Mexican needlets can be given by ∞Xℓ=0 ψjk;s (x) :=pλjk pλjk fs(cid:18)√−eℓ Bj (cid:19) ℓXm=−ℓ fs(cid:18)√−eℓ Bj (cid:19) 2ℓ + 1 ∞Xℓ=0 4π Y ℓm (ξjk) Yℓm (x) Pℓ (hx, ξjki) , (4) {eℓ} denoting the spectrum of the spherical Laplacian ∆S2 associated to the eigenfunctions {Yℓm}, i.e., eℓ = −ℓ(ℓ + 1), whence (∆S2 − eℓ) Yℓm (x) = 0. Note that the last equality in (4) is due to (1). Remark 2.2. (4) is the standard definition of Mexican needlets in the litera- ture, while Theorem 3.1 is focussed on (3). Consider, anyway, that the difference between (3) and (4) concerns only the argument of fs (·). In the former, the argument is given by the square root of the eigenvalue of the Laplacian operator −eℓ, while in the latter it is replaced by ℓ + 1 2 . This formulation is instrumental for the derivation of the localization property in Section 3 (cf., more specifically, (7)), as in [24] for the standard needlet case (see also Section 13.3 in the Ap- pendix of [17]). This is a minor difference, asymptotically negligible considering that, trivially, lim j→∞ lim ℓ→∞ fs(cid:16)√−eℓ Bj (cid:17) Bj (cid:17) = lim fs(cid:16) ℓ+ 1 j→∞ 2 lim ℓ→∞ (cid:16) ℓ(ℓ+1) B2j (cid:17)s B2j (cid:19)s (cid:18) (ℓ+ 1 2 )2 B2j (cid:17) exp(cid:16)− ℓ(ℓ+1) B2j (cid:19) = 1 . exp(cid:18)− (ℓ+ 1 2 )2 As a consequence, for any x ∈ S2, we obtain an asymptotic equivalence between Ψjk;s (x) and ψjk;s (x). Hence, the localization property, proved in Theorem 3.1 for (3), holds also for the Mexican needlets given by (4). For F ∈ L2(cid:0)S2(cid:1) and for any j, k, let the Mexican needlet coefficients be given by βjk;s := hF, ψjk;siL2(S2) . It is proven in [11] that there exists a constant C0 = C0 (B, cB, fs) such that (mB;s − C0)kFk2 L2(S2) ≤ βjk2 ≤ (MB;s + C0)kFk2 L2(S2) . Hence, if mB;s − C0 > 0, {ψjk;s} is a frame for L2(cid:0)S2(cid:1) bounded by (mB;s − C0) and (MB;s + C0). It holds that NjXk=1 ∞Xj=−∞ = 1 + O(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) B 5 MB;s + C0 mB;s − C0 ∼ MB;s mB;s 1 2 !2 2 − 1 B 1 log B 1 2 − 1 B 1 2 !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  , as B → 1, B > 1, and that NjXk=1 ∞Xj=−∞ βjk;s2 = where δ := δ (B) = O (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18) B 2 (cid:19)2 2 −1 B that 1 1 log(cid:18) B 1 2 −1 B 1 2 (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) δ (B) = 0, lim B→1 cf. [11]. ηs (1 + δ) log B kFk2 L2(S2) , ! as B → 1, B > 1, is such Remark 2.3. As well-known in the literature, standard needlets describe a tight frame with tightness constant equal to 1, allowing for an exact reconstruction formula (cf. [24, 25] and the textbook [17, Section 10.3]). On the other hand, Mexican needlets are characterized by a non-compact support in the harmonic domain, and this makes perfect reconstruction unfeasible for the lack of an exact cubature formula. Despite these features, Mexican needlets enjoy some remark- able advantages with respect to the standard ones. More specifically, they are characterized by an extremely good concentration properties in the real domain. In addition, it is possible to choose the measurable disjoint sets Ejk with mini- mal conditions, and still ensure frame constants arbitrarily close to unity (and hence almost exact reconstruction). Note that, in this paper, we investigate the exact dependence of localization properties upon s, an issue which is extremely relevant for applications (see, for example, [27]). It may be noted that the choice of s represents a trade- off between localization in real and harmonic domain; the latter improves as s increases, while the reverse holds for the former. We add here the following result, which establishes a link between Mexican needlets with different shape parameter s ∈ N. Proposition 2.1. For any s ∈ N, where s > 1, and x ∈ S2, ψjk;s (x) = (−1)s B−2js (∆S2 )s ψjk;1 (x) . Proof. Easy calculations lead to − B−2j∆S2 ψjk;s (x) = −∆S2pλjk =pλjkXℓ≥0(cid:18)−eℓ B2j(cid:19)s B2j Xℓ≥0(cid:18)−eℓ B2j(cid:19)s+1 exp(cid:18)−eℓ exp(cid:16) eℓ B2j(cid:19) ℓXm=−ℓ B2j(cid:17) ℓXm=−ℓ = ψjk;s+1 (x) . Y ℓm (ξjk) Yℓm (x) Y ℓm (ξjk) Yℓm (x) Iterating the procedure, we obtain the statement. 6 Before concluding this Section, for ξjk, x ∈ S2, let us label the geodesic distance by so that we can express the Mexican needlets given by (3) in terms of ϑ ϑ := ϑjk (x) = hx, ξjki, Ψjk;s (ϑ) :=pλjk 1 2π ∞Xℓ=0 fs (cid:0)ℓ + 1 2(cid:1)Bj !(cid:18)ℓ + 1 2(cid:19) Pℓ (cos ϑ) . (5) 3. The localization property The aim of this Section is to achieve an exhaustive proof of the so-called localization property, i.e., to establish an upper bound for the supremum of the modulus of the Mexican needlet defined by (5), remarking its dependence on the resolution level j and on the shape parameter s, up to a multiplicative constant. This result is given in the Theorem 3.1. We stress again that this achievement was pursued implicitely by Geller and Mayeli in [11], where the authors anyway found a similar result studying (5) for small and large angles, even if they limited their investigations to the case s = 1. Here, instead, we extend this result to any value of the shape parameter s in (2), holding for any value of ϑ by means of a unique procedure, which resembles the one employed by Narcowich, Petrushev and Ward in [24] to exploit the localization property for standard needlets on the n-dimensional sphere Sn (see also [25] and the textbook [17, Section 13.3]). In this case, howsoever, we will take advantage of the explicit formulation of the weight function (2), which allows us to compute exactly its Fourier transform in terms of Hermite polynomials and, through that, to exploit precisely the dependence on the resolution level j of the supΨjk;s (ϑ). For the sake of simplicity, let us introduce the following notation ε = ε (B, j) := B−j , so that, we can define Ψε;s (ϑ) := 1 2π Remark 3.1. Observe that fs(cid:18)ε(cid:18)ℓ + 1 2(cid:19)(cid:19)(cid:18)ℓ + 1 2(cid:19) Pℓ (cos ϑ) . (6) ∞Xℓ=0 Ψjk;s (x) = Ψjk;s (ϑ) =pλjkΨε;s (ϑ) , Furthermore, in (6), while the index ε substitutes j, the index k is no more necessary. As stressed above, λjk does not appear in (6), while Theorem 3.1 holds for any ϑ ∈ [0, π]. The dependence on k arises when we choose ϑ = ϑjk (x). Theorem 3.1. Let Ψjk;s (ϑ) be given by (3). Then, for any s ∈ N and k = 1, . . . , Nj, there exists Cs > 0 such that Ψjk;s (x) ≤ CsBje− B2j 4 ϑ2(x)(cid:16)1 +(cid:12)(cid:12)Bj ϑ (x)(cid:12)(cid:12)2s(cid:17) , uniformly over j. 7 Proof. By using the Mehler-Dirichlet representation formula (see, for instance, [30, Formula 4.8.7, pag. 86]), the Legendre polynomial of degree ℓ can be written as ϑ dφ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dφ Hence, using (6), we have that ϑ 1 1 1 = dφ. √2 1 2π Pℓ (cos ϑ) = 2(cid:1) φ(cid:1) sin(cid:0)(cid:0)ℓ + 1 π Z π √cos ϑ − cos φ Ψε;s (ϑ) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2(cid:1) φ(cid:1) sin(cid:0)(cid:0)ℓ + 1 2(cid:19)Z π 2(cid:19)(cid:19)(cid:18)ℓ + fs(cid:18)ε(cid:18)ℓ + ∞Xℓ=0 √cos ϑ − cos φ 2(cid:1) φ(cid:1)(cid:12)(cid:12) ϑ (cid:12)(cid:12)P∞ℓ=0 fs(cid:0)ε(cid:0)ℓ + 1 2(cid:1)(cid:1)(cid:0)ℓ + 1 2(cid:1) sin(cid:0)(cid:0)ℓ + 1 2πZ π √cos ϑ − cos φ 2πZ π √cos ϑ − cos φ 2(cid:19)(cid:19)(cid:18)ℓ + fs(cid:18)ε(cid:18)ℓ + ∞Xℓ=0 2(cid:19) gε,φ;s(cid:18)ℓ + ∞Xℓ=0 gε,φ;s(cid:18)ℓ + ∞Xℓ=−∞ 2(cid:19) sin(cid:18)(cid:18)ℓ + 2(cid:19) φ(cid:19) 2(cid:19) . Λε;s (φ) Λε,s (φ) dφ , ≤ 1 2 := = 1 1 = 1 1 1 where ϑ 1 (7) In the last equality, we use the fact that gε,φ;s (u) := fs (εu) u sin (uφ) , u ∈ R is an even function. Using Lemma 4.1, we obtain . ϑ ε2 (8) 2ε )2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Note that Z π H2s+1(cid:16) φ 2ε(cid:17) dφ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) e−( φ Ψε;s (ϑ) ≤ eC2s+1 √cos ϑ − cos φ 2 (cid:17) sin(cid:16) φ−ϑ 2 (cid:17) cos ϑ − cos φ = 2(cid:0)φ2 − ϑ2(cid:1) sin(cid:16) ϑ+φ (cid:16) ϑ+φ 2 (cid:17)(cid:16) φ−ϑ 2 (cid:17) 2 , π(cid:3). Finally, in case III, we 2(cid:1), where 0 < δ < ε. In case II, ϑ ∈(cid:2) π 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) . 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s(cid:18) ϑ that ϑ ∈(cid:0)δ, π there exists a constant eCs so that Ψε;s (ϑ) ≤ eCs have that ϑ ∈ [0, δ], 0 < δ < ε, as in [24]. In cases I and II, we will prove that In order to estimate (8), we consider three different cases. In case I, we have ε2 e−( ϑ (9) . 8 We distinguish between the two cases for technical reasons. Indeed, in case II, the integral in (8) will be estimated by using supplementary angles. As far as case III is concerned, we will prove that there exists C′′′s such that Ψε;s (ϑ) ≤ C′′′s ε2 . Case I. We have that 0 < 0 ≤ ϑ + φ 2 ≤ 2 ≤ φ − ϑ 3 4 π 2 π , . On one hand, note that (9) can be bounded as: where CI > 0. On the other hand, the integral (8) can be rewritten as cos ϑ − cos φ ≥ Ψε;s (ϑ) ≤ eC′2s+1 ε2 √2 2 4 π 4 3π 2 1 2(cid:0)φ2 − ϑ2(cid:1) √2 = CI(cid:0)φ2 − ϑ2(cid:1) , (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z π e−( φ 2ε )2 H2s+1(cid:16) φ 2ε(cid:17) p(φ2 − ϑ2) ϑ , dφ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) where eC′2s+1 > 0. Recall that, for n odd and u ∈ R, the Hermite polynomials can be rewritten as Hn (u) = n! (2u)2r+1 , (10) (see, for instance, [1, Chapter 22]). Therefore, we have that Ψε;s (ϑ) ≤ eC′2s+1 ε2 where Using Lemma 4.2, which establishes that n−1 n−1 2 −r (−1) 2 − r(cid:1)! (−1)s−r 22r+1 (2r + 1)! (s − r)! 2Xr=0 (2r + 1)!(cid:0) n−1 (2s + 1)!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) sXr=0 2ε )2(cid:16) φ 2ε(cid:17)2r+1 Qε,r (ϑ) :=Z π p(φ2 − ϑ2) 2ε(cid:19)2r 2ε )2(cid:18) ϑ Qε,r (ϑ) ≤ Cre−( ϑ e−( φ dφ. ϑ , , Qε,r (ϑ) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 9 where Cr > 0, we get ε2 Ψε;s (ϑ) ≤ eC′2s+1 ≤ eC′2s+1 C′s ε2 e−( ϑ ε2 = e−( ϑ (2s + 1)! (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) sXr=0 (−1)s−r 22r+1 (2r + 1)! (s − r)! 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) sXr=0 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) , 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s(cid:18) ϑ Qε,r (ϑ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2ε(cid:19)2r(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (2r)! (s − r)! (cid:18) ϑ (−1)s−r 22r+1 (2s)! Cr where C′s > 0. Case II. As in [24] (see also [17, Section 13.3], we use the supplementary angles to φ and ϑ, denoted by eφ = π − φ and eϑ = π − ϑ, respectively. We can easily observe that 2 ≤ 0 ≤ eϑ +eφ 0 ≤ eϑ −eφ 2 ≤ π 2 π 2 ; , so that we get where CII > 0. By substitution in (8), following the same procedure as above yields 2ε (cid:17)2 e−(cid:16) π−eφ 2 (cid:17) sin(cid:16) eϑ+eφ 2 (cid:17) coseφ − coseϑ ≥ 2(cid:16)eϑ2 −eφ2(cid:17) sin(cid:16) eϑ−eφ 2 (cid:17)(cid:16) eϑ+eφ 2 (cid:17) (cid:16) eϑ−eφ ≥ CII(cid:16)eϑ2 −eφ2(cid:17) , (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) eQε,r (ϑ) :=Z eϑ deφ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) H2s+1(cid:16) π−eφ 2ε (cid:17) r(cid:16)eϑ2 −eφ2(cid:17) (2r + 1)! (s − r)!eQε,r (ϑ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) sXr=0 2ε(cid:17)2(cid:16) π−eφ 2ε (cid:17)2r+1 e−(cid:16) eφ r(cid:16)eϑ2 −eφ2(cid:17) Ψε;s (ϑ) ≤ eC′′2s+1 ≤ eC′′2s+1 (−1)s−r Z eϑ deφ. (2s + 1)! ε2 ε2 0 0 , 10 where eC′′2s+1 > 0 and Finally, using Lemma 4.2 leads to and, as a straightforward consequence, we get Case III. Following again [24], we have that Ψε;s (ϑ) ≤ Λε;s (φ) ≤ 1 , 1 1 CIII CIII e−(cid:16) eϑ 2ε(cid:17)2 C′′s ε2 e−( ϑ eQε,r (ϑ) ≤ eCr eϑ 2ε!2r 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s(cid:18) ϑ 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) . 2(cid:19)(cid:19) ε2(cid:18)ℓ + 2(cid:19) fs(cid:18)ε(cid:18)ℓ + (εℓ)2s+1 e−(εℓ)2Z e(ℓ+1) ε2Xℓ≥0 ε2 Xℓ≥0 ε2 Z ∞ Γ(cid:0)s + 3 2(cid:1) (cid:18)1 +(cid:12)(cid:12)(cid:12)(cid:12)H2s(cid:18) ϑ ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) , 4 ϑ2(x)(cid:16)1 +(cid:12)(cid:12)Bj ϑ (x)(cid:12)(cid:12)2s(cid:17) , e−( ϑ ε2 u2s+1e−u2 du ≤ ≤ = CIII ε2 = C′′′s ε2 . 0 2 2ε )2 du εℓ Combining these results yields Ψε;s (ϑ) < Cs for ϑ = [0, π]. From Remark 3.1 and since λjk ≤ cB−2j , we have that Ψjk;s (x) ≤ CsBje− B2j as claimed. Remark 3.2. In view of Remark 2.2, it holds ψjk;s (x) ≤ CsBje− B2j 4 ϑ2(x)(cid:16)1 +(cid:12)(cid:12)Bjϑ (x)(cid:12)(cid:12)2s(cid:17) . Remark 3.3. As suggested in [27], Mexican needlets in the case s = 1 pro- vide a valid asymptotic approximation to the spherical Mexican hat wavelets . Recall that the discretized version of spherical Mexican hat wavelets use a stere- ographic projection on the sphere (see [2]). These wavelets conserve the most crucial properties of the flat Mexican hat wavelets and, for this reason, they are widely used in Astrophysics and Cosmology, even if they lack of a reconstruction formula (see, for example, [20], for a quick review). In [27], it is proved that the bound between the absolute value of the difference between spherical Mexican hat wavelets and Mexican needlets is of order B−j min(cid:0)ϑ4B4j, 1(cid:1). This bound 11 matches exactly with the results proved in Theorem 3.1. Indeed, fixed s = 1, the bound for small angles is controlled by B−j, which depends, on one hand, on the normalization factor of the spherical Mexican hat wavelet and, on the other, onpλjk. Up to a proper normalization, for larger angles, this factor has to be multiplied by a series expansion of even powers of ϑ, controlled by lead- ing term of order 4 (cf. [10]). Heuristically, it implies that spherical Mexican hat wavelets can be approximated to Mexican needlets in the corresponding Ejk and that this approximation is better for large j. For this reason, the spatial concentration properties here discussed and the correlation properties studied in [16, 18] can be helpful for the study of the asymptotic behaviour of the random spherical Mexican wavelet coefficients hat in the high-frequency limit. Before concluding this Section, as for the standard needlets (see [24, 25]), we can also establish the order of the Lp-norms of Mexican needlets as follows. Corollary 3.2 (Bounds on Lp(cid:0)S2(cid:1)-norms). For any p ∈ [1,∞) , there exist cp, Cp ∈ R such that cpB2j( 1 2 − 1 p ) ≤ kΨjk;skLp(S2) ≤ CpB2j( 1 2 − 1 p ). Furthermore, there exist c∞, C∞ ∈ R such that c∞Bj ≤ kΨjk;skL∞(S2) ≤ C∞Bj. the estimate of the bounds for L2(cid:0)S2(cid:1) norms. Proof. The proof of this Corollary is very close to the one developed in the standard needlet framework in [25]. The only remarkable difference concerns In [25], this bound is proven as corollary of the tight-frame property. We establish a similar result for the Mexican needlet framework as follows. Let dx denote the uniform spherical measure. Hence, we have that Bj(cid:19) fs(cid:18) ℓ′ Bj(cid:19) Yℓm (x) Y ℓ′m′ (x) Y ℓm (ξjk) Yℓ′m′ (ξjk) dx Y ℓm (ξjk) Yℓm (ξjk) δℓ′ ℓ δm′ m kΨjk;sk2 L2(S2) =ZS2 Ψjk;s (x)2 dx fs(cid:18) ℓ =λjkZS2 ∞Xℓ=0 ∞Xℓ′=0 ℓ′Xm′=−ℓ′ ℓXm=−ℓ Bj(cid:19) ℓXm=−ℓ s(cid:18) ℓ ∞Xℓ=0 s(cid:18) ℓ Bj(cid:19) 2ℓ + 1 ∞Xℓ=0 =λjk =λjk × f 2 f 2 4π . 12 On one hand, we get λjk ∞Xℓ=0 f 2 s(cid:18) ℓ Bj(cid:19) 2ℓ + 1 4π ≤ c 2π = ≤ ≤ Bj f 2 c 2π 1 Bj 1 Bj Bj(cid:19) ℓ + 1/2 s(cid:18) ℓ ∞Xℓ=0 Bj(cid:19)4s ∞Xℓ=0(cid:18) ℓ Bj(cid:19)4s ∞Xℓ=0(cid:18) ℓ 2πZ ∞ u4s+1e−2u2 e−2( ℓ c 2π c 0 du ≤ C2. e−2( ℓ Bj )2 ℓ + 1/2 Bj Bj )2 ℓ Bj Bj Z ℓ+1 ℓ Bj du On the other hand, we obtain λjk ∞Xℓ=0 Following [25], we get f 2 Bj(cid:19) 2ℓ + 1 s(cid:18) ℓ 4π ≥ c2. cpB2j( 1 2 − 1 p ) ≤ kΨjk;skLp(S2) ≤ CpB2j( 1 2 − 1 p ), as claimed. 4. Auxiliary results In this Section we collect some auxiliary results, concerning the upper bounds of Λε;s, Qε,r and eQε,r. We introduce preliminarily the following notation for the Fourier transform of a function f ∈ L1 (R): F [f ] (ω) :=ZR f (u) e−iωudu =: bf (ω) . Let us also recall two standard properties for the Fourier transforms. Under standard conditions, we have that Finally, the Poisson Summation Formula can be defined as follows. If, for ω ∈ [0, 2π] and α > 0, dα dωαbf (ω) = (−i)α F [uαf (u)] (ω) ; F(cid:20) dα duα f (u)(cid:21) (ω) = (−i)α ωαF [f (u)] (ω) . f (u) +(cid:12)(cid:12)(cid:12)bf (ω)(cid:12)(cid:12)(cid:12) ≤ 1 + uα+1 , Ca 13 then: f (τ ) e−iωτ = ∞Xτ =−∞ ∞Xν=−∞bf (ω + 2πν) . (11) For more details and discussions about Fourier transforms, the reader is referred, for instance, to the textbook [29]. Lemma 4.1. Let Λε;s (φ) be given by (7). Then there exists eC2s+1 > 0 such that Proof. First, note that straightforward calculations lead to On one hand, we have that ε2 e−( φ 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) . 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ 2(cid:19)(cid:21) (ω) = ei ω 2bgε,φ;s (ω) . Λε;s (φ) ≤ eC2s+1 F(cid:20)gε,φ;s(cid:18)u + F [fs (εu) u] (ω) = ZR fs (εu) ue−iωu 1 1 ε(cid:17) . ε2 F [fs (u) u](cid:16) ω On the other hand, note that = Fhe−u2i (ω) = √πe− ω2 4 . Combining (2) and (12) yields (12) , F [fs (u) u] (ω) = i2s+1 d2s+1 dω2s+1 Fhe−u2i (ω) d2s+1 dω2s+1 e− ω2 2 ) √πH2s+1(cid:16) ω 4 2(cid:17) e− ω2 4 = i2s+1√π = (−1)(s+ 1 where H2s+1 (·) is the Hermite polynomial of order 2s + 1. Recall that the polynomials composing Hn (·) are all even (odd) if n is even (odd) - for more details, see, for instance, [1, Chapter 22]. Collecting all these results, we get F [fs (εu) u] (ω) = Hence, we obtain: 2 ) √π (−1)(s+ 1 ε2 H2s+1(cid:16) ω 2ε(cid:17) e−( ω 2ε )2 . bgε,φ;s (ω) =F [sin (φu)] (ω) ∗ F [fs (εu) u] (ω) 2ε (cid:19) e−( ω−φ (cid:18)H2s+1(cid:18) ω − φ (−1)s π ε2 = 3 2 2ε )2 −H2s+1(cid:18) ω + φ 2ε (cid:19) e−( ω+φ 2ε )2(cid:19) . 14 Using (11), we have that gε,φ;s(cid:18)ℓ + 1 2(cid:19) = ∞Xℓ=−∞ = 2 3 2 ei 2πν ei 2πν ∞Xν=−∞ 2 bgε,φ;s (2πν) ∞Xν=−∞ (−1)s π ε2 2ε (cid:19) e−( 2πν+φ − H2s+1(cid:18) 2πν + φ ∞Xν=−∞ (−1)s+1 π ei 2πν ε2 3 2 2ε 2 =2 )2(cid:19) (cid:18)H2s+1(cid:18) 2πν − φ 2ε (cid:19) e−( 2πν−φ 2ε )2 H2s+1(cid:18) 2πν + φ 2ε (cid:19) e−( 2πν+φ 2ε )2 , where the last equality takes into account that H2s+1 (·) is odd. Therefore, we have that Then, we obtain Λε;s (φ) = where Note that where Λε;s (φ) = 3 2 (−1)s+1 π ε2 ∞Xν=−∞ ei 2πν )2 . 2ε 2 H2s+1(cid:18) 2πν + φ 2ε (cid:19) e−( 2πν+φ )2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ )2(cid:12)(cid:12)(cid:12)(cid:12) . 2ε (cid:19) e−( 2πν+φ 3 2 2ε + V+ + V− , π ε2 Vε,s (φ) , Vε,s (φ) = π 3 2 ε2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Vε,s (φ) =(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ 2ε 2ε )2 2ε (cid:19) e−( 2πν+φ H2s+1(cid:18) 2πν + φ ∞Xν=−∞ ∞Xν=−∞(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) 2πν + φ 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−( φ 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−( 2πν+φ ∞Xν=1(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) 2πν + φ −∞Xν=−1(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) 2πν + φ 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−( 2πν+φ 2Xk=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 − k(cid:1)! (2k + 1)!(cid:0) n−1 ≤ C′n un . (−1) 2 −k n−1 n−1 2ε 2ε V+ = V− = Hn (u) ≤ n! )2 , )2 . (2u)2k+1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Using (10), for u > 1, we have that 15 (13) (14) Hence, we get (cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18)2πν + φ 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)e−( 2πν+φ 2ε )2 Note that 2ε 2s+1 2πν + φ (cid:12)(cid:12)(cid:12)(cid:12) ≤C2s+1(cid:12)(cid:12)(cid:12)(cid:12) 2ε (cid:19)2s+1 =C2s+1(cid:18) 2πν + φ C2s+1"(cid:18) πν C2s+1(cid:20)(cid:16) πν 2ε )2 ≤e−( φ =e−( φ 2ε )2 + + ε ε πu ε + π 2ε < 2πu ε ; e−( 2πν+φ 2ε )2 e−( φ 2ε )2 e−( πν ε )2 e− 2πνφ 4ε2 φ 2ε(cid:19)2s+1 2ε(cid:17)2(s+1) π e−( πν ε )2 e− 2πνφ 4ε2 # e−( πν ε )2(cid:21) . )2(cid:12)(cid:12)(cid:12)(cid:12) ε )2 ε )2 furthermore, observing that e−νφ < 1, we obtain 2ε ε π + C2s+1 e−( πν V+ ≤C2s+1 2ε )2 ≤e−( φ ∞Xν=1(cid:16) πν ∞Xν=1(cid:12)(cid:12)(cid:12)(cid:12)H2(s+1)(cid:18) 2πν + φ 2ε (cid:19) e−( 2πν+φ 2ε(cid:17)2s+1 ε (cid:17)2s+1 C2s+122s+1 ∞Xν=1(cid:16) πν 2ε )2 ≤e−( φ 2ε )2 ≤C′2s+1e−( φ Indeed, the series P∞ν=1(cid:0) πν ε (cid:1)2s+1 (cid:17)2s+1 ν→∞(cid:16) π(ν+1) (cid:0) πν ε (cid:1)2s+1 e−( πν means of the D'Alembert's criterion, i.e., for all ν > 1. On the other hand, if u ≤ 1, we obtain ν→∞(cid:18)1 + v(cid:19)2s+1 e−( π(ν+1) e−( πν e−( πν = lim ε )2 ε )2 lim )2 1 . ε ε Hence, we have that Hn (u) ≤ n! 1 (2k + 1)!(cid:0) n−1 n−1 2Xk=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−( 2πν+φ 2ε )2 (cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) 2πν + φ is convergent, as easily proved by (15) ε2 (2ν + 1)(cid:19) = 0, π2 exp(cid:18)− 22k+1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C′n. )2 2ε 2 − k(cid:1)! ≤C′2s+1e−( 2πν+φ 2ε )2 =C′2s+1e−( φ =e−( φ 2ε )2 ≤e−( φ 2ε )2 16 e−( πν ε )2 ε )2 C′2s+1he−( πν C′2s+1e−( πν ε )2 . 4ε2 e− 2πνφ e− 2πνφ 4ε2 i V− = Using (14) yields (cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18)φ − 2πν′ 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−(cid:16) φ−2πν ′ Since φ < π, straightforward calculations lead to Therefore, we obtain V+ ≤C2s+1 2ε (cid:19) e−( 2πν+φ 2ε )2(cid:12)(cid:12)(cid:12)(cid:12) C2s+1 ∞Xν=1(cid:12)(cid:12)(cid:12)(cid:12)H2(s+1)(cid:18) 2πν + φ C2s+122s+1 ∞Xν=1 ∞Xν=1 e−( πν e−( πν ε )2 ε )2 2ε )2 ≤e−( φ 2ε )2 ≤e−( φ 2ε )2 ≤C′2s+1e−( φ ε )2 , since the seriesP∞ν=1 e−( πν Consider now the sum V−. Let us define ν′ = −ν, so that is convergent. (cid:17)2 . 2ε ∞Xν ′=1(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ − 2πν′ ≤C2s+1(cid:12)(cid:12)(cid:12)(cid:12) 2ε )2 (cid:17)2 2ε 2ε 2ε (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−(cid:16) φ−2πν ′ (cid:12)(cid:12)(cid:12)(cid:12) φ − 2πν′ 2s+1 φ − 2πν′ 2ε 2ε =e−( φ C2s+1"(cid:12)(cid:12)(cid:12)(cid:12) ∞Xν ′=1(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ − 2πν′ (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) e−(cid:16) φ−2πν ′ (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ε (cid:19)2s+1 exp(cid:18)− φ − 2πν′ π − 2πν′ C′′2s+1 C2s+1 C2s+1 2s+1 2s+1 2πν ′φ 2ε 2ε π2ν ′ e e 2ε (cid:17)2 (cid:12)(cid:12)(cid:12)(cid:12) ∞Xν ′=1"(cid:12)(cid:12)(cid:12)(cid:12) ∞Xν ′=1"(cid:12)(cid:12)(cid:12)(cid:12) ∞Xν ′=1"(cid:18) πν′ ε (cid:17)2s+1 V− ≤C2s+1 2ε )2 ≤e−( φ 2ε )2 ≤e−( φ 2ε )2 ≤e−( φ 2ε )2 ≤C′′2s+1e−( φ . e−(cid:16) φ−2πν ′ 2ε (cid:17)2 2s+1 e 2πν ′φ 4ε2 e−(cid:16) πν ′ ε (cid:17)2# . π2 2ε2 e− π2 ε (cid:17)2# 4ε2 e−(cid:16) πν ′ ε2 (ν ′)2# ε2 ν′(cid:20)ν′ − 2(cid:3)(cid:17)(cid:21) can be proved to be 2(cid:21)(cid:19)# (16) 1 Indeed, the seriesP∞ν ′=1(cid:20)(cid:16) πν ′ exp(cid:16)− π2 ε2 ν′(cid:2)ν′ − 1 convergent by means of the D'Alembert's criterion, as above. Combining (15) and (16) in (13), the term corresponding to ν = 0 is dominant. 17 Hence, we have that Vε,s (φ) ≤e−( φ Thus as claimed. ϑ Qε,r (ϑ) :=Z π eQε,r (ϑ) :=Z eϑ Then, there exist Cr,eCr > 0 so that 0 ≤C2s+1e−( φ Λε;s (φ) ≤ eC2s+1 ε2 dφ, e−( φ deφ. 2ε )2(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) + C′2s+1 + C′′2s+1(cid:19) 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) . 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ 2ε(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) , 2ε )2(cid:12)(cid:12)(cid:12)(cid:12)H2s+1(cid:18) φ 2ε )2(cid:16) φ 2ε(cid:17)2r+1 p(φ2 − ϑ2) 2ε (cid:17)2r+1 2ε(cid:17)2(cid:16) π−eφ e−(cid:16) eφ r(cid:16)eϑ2 −eφ2(cid:17) 2ε )2(cid:18) ϑ 2ε(cid:19)2r Qε,r (ϑ) ≤ Cre−( ϑ 2ε!2r eQε,r (ϑ) ≤ eCr eϑ e−(cid:16) eϑ ϑ )2(cid:16) φ 2ε(cid:19)2r+1Z π r(cid:16) φ ϑ(cid:17)2 t =(cid:16)(φ/ϑ)2 − 1(cid:17) 1 2ε(cid:19)2r+1Z (cid:16)( π −1(cid:17) 1 2ε(cid:19)2r+1Z ∞ 2ε(cid:19)2r+1 2ε )2(cid:18) ϑ 2ε )2(cid:18) ϑ 2ε )2(cid:18) ϑ t2(cid:0)t2 + 1(cid:1)r ϑ(cid:17)2r+1 (I1 + I2) , e−( ϑ 2ε · φ e−( ϑ e−( ϑ 2ε )2 − 1 2ε(cid:17)2 . ϑ )2 2 1 ϑ dφ . 0 0 2 , , 18 2ε )2 t2(cid:0)t2 + 1(cid:1)r dt dt Lemma 4.2. For any ϑ ∈ [0, π], let Qε,r (ϑ) and eQε,r (ϑ) be given by e−( φ Proof. First, note that Qε,r (ϑ) =(cid:18) ϑ ϑ Then, we use the substitution in order to obtain Qε,r (ϑ) =e−( ϑ ≤e−( ϑ =e−( ϑ where On one hand, for t ∈ [0, 1], we have that On the other hand, for t ∈ (1,∞), we obtain Hence, we get I1 ≤2rZ 1 I2 ≤Z ∞ 0 1 e−( ϑ 2ε )2 e−( ϑ 2ε )2 Straightforward calculations lead to I1 :=Z 1 I2 :=Z ∞ 1 0 e−( ϑ 2ε )2 e−( ϑ 2ε )2 dt, dt. t2(cid:0)t2 + 1(cid:1)r t2(cid:0)t2 + 1(cid:1)r ≤ 2r. t2 (cid:0)t2 + 1(cid:1)r (cid:0)t2 + 1(cid:1)r ≤ (2t)2r . dt ≤ 2rZ ∞ (2t)2r dt ≤ 4rZ ∞ 2ε(cid:19)−1 e−( ϑ t2 0 0 , I1 ≤2r−1√π(cid:18) ϑ I2 ≤(cid:18) ϑ 2ε )2 t2 dt, e−( ϑ 2ε )2 t2 t2rdt. ′′ . , 2 r e−( ϑ as claimed. ≤Cre−( ϑ Qε,r (ϑ) ≤C Therefore, we obtain 2ε(cid:19)−2r! 2ε(cid:19)−(2r+1) Γ(cid:0)r − 1 2(cid:1) 2ε(cid:19)2r 1 +(cid:18) ϑ 2ε )2(cid:18) ϑ 2ε(cid:19)2r 2ε )2(cid:18) ϑ As far as eQε,r (ϑ) is concerned, note that the following inequality holds 2ε!2 = exp"−(cid:16) π 2ε# + eφ 2ε(cid:17)2 πeφ ≤ exp(cid:20)− 2ε(cid:18) 1 2ε − 1(cid:19)(cid:21) . 2u(cid:18) 1 2u − 1(cid:19)(cid:21) . exp− π −eφ 2ε !2 Then, let the function γ (·) on R be given by γ (u) := exp(cid:20)− + 2 π2 π2 19 π2 2 so that so that lim u→±∞ 1 π2 γ′′ (u) = 3π2 γ′ (u) = π2 Finally, note that Furthermore, since Hence, we have that 1 u2 + 1 u − On the other hand, we have that 2u − 1(cid:19)(cid:21) , 2u − 1(cid:19)(cid:21) , 2u(cid:18) 1 2u2(cid:18) 1 u − 1(cid:19) exp(cid:20)− γ′ (u) = 0 ⇐⇒ u = 1. 2u(cid:18) 1 3(cid:21) exp(cid:20)− 2u2(cid:20)− 4 (cid:19) < 0. exp(cid:18) π2 γ′′ (1) = − γ (u) = 0 < exp(cid:18) π2 4 (cid:19) = γ (1) . 2ε!2 ≤ exp(cid:20) π2 exp− π −eφ 2ε !2 + eφ 4 (cid:21) . 2ε!2r+1 2ε !2r+1 2ε!2r+1(cid:18) π = eφ π −eφ ≤ eφ eφ − 1(cid:19)2r+1 eϑ(cid:17)2(cid:16) eφ 2ε(cid:17)2r+1 eϑ(cid:17)2r+1(cid:16) eϑ e−(cid:16) eϑ 2ε eφ eQε,r (ϑ) ≤ eC′rZ eϑ s(cid:18)1 −(cid:16) eφ eϑ(cid:17)2(cid:19) Using the substitution t =r1 −(cid:16)eφ/eϑ(cid:17)2 2ε!2r+1 eQε,r (ϑ) ≤eC′′r eϑ 2ε!2r+1 ≤eC′′r eϑ ≤eCr eϑ 2ε!2r e−(cid:16) eϑ 2ε(cid:17)2Z 1 e−(cid:16) eϑ 2ε(cid:17)2Z 1 e−(cid:16) eϑ t2(cid:0)1 − t2(cid:1)r 2ε(cid:17)2 e(cid:16) eϑ 2ε(cid:17)2 e(cid:16) eϑ t2 dt deφ. 1 eϑ we obtain as claimed. , dt Note that γ (·) achieves its absolute maximum for u = 1. Indeed, on one hand, we have that π2 0 , we get 0 0 2ε(cid:17)2 , 20 Acknowledgement. The author wishes to thank Federico Cacciafesta for the helpful conversations and his enlightening suggestions and Domenico Marinucci for his precious hints and his fundamental corrections. This paper is dedicated to Amalia Olivieri. References [1] Abramowitz, M. and Stegun, I. (1946). Handbook of Mathematical Functions. Dover, New York. [2] Antoine, J.-P. and Vandergheynst, P. (2007). Wavelets on the Sphere and Other Conic Sections. J. Fourier Anal. Appl., 13, 4, 369 -- 386. [3] Cammarota, V. and Marinucci, D. (2015). On the Limiting Behaviour of Needlets Polyspectra. Ann. Henri Poincaré Probab. Stat., 51, 3, 1159 -- 118. [4] Dahlke, S., Steidtl, G. and Teschke, G. (2007). Frames and Coorbit Theory on Homogeneous Spaces with a Special Guidance on the Sphere. J. Fourier Anal. Appl., 13, 4, 387 -- 404. [5] Durastanti, C. and Lan, X., (2013). High-Frequency Tail Index Esti- mation by Nearly Tight Frames, Amer. Math. Soc. Contemp. Math., 603. [6] Durastanti, C., Fantaye, Y.T., Hansen, F.K., Marinucci, D. and Pesenson, I.Z. (2014). Simple proposal for radial 3D needlets. Phys. Rev. D, 90, 103532. [7] Freeden, W. and Schreiner, M. (1998). Orthogonal and nonorthogo- nal multiresolution analysis, scale discrete and exact fully discrete wavelet transform on the sphere. Constr. Approx., 14, 4, 493 -- 515. [8] Geller, D. and Marinucci, D. (2010). Spin Wavelets on the Sphere. J. Fourier Anal. Appl., 16, 6, 840 -- 884. [9] Geller, D. and Marinucci, D. (2011). Mixed Needlets. J. Math. Anal. Appl, 375, 2, 610 -- 630. [10] Geller, D. and Mayeli, A. (2009). Continuous Wavelets on Manifolds. Math. Z., 262, 895 -- 927. [11] Geller, D. and Mayeli, A. (2009). Nearly Tight Frames and Space- Frequency Analysis on Compact Manifolds. Math. Z., 263, 235 -- 264. [12] Geller, D. and Mayeli, A. (2009). Besov Spaces and Frames on Com- pact Manifolds. Indiana Univ. Math. J., 58, 2003 -- 2042. [13] Gorski, K.M., Hivon, E., Banday, A.J., Wandelt, B.D., Hansen, F.K., Reinecke, M. and Bartelman, M. (2005). HEALPix, a Frame- work for High Resolution Discretization, and Fast Analysis of Data Dis- tributed on the Sphere. Astrophys. J., 622, 759 -- 771. 21 [14] Holschneider, M. and Iglewska-Nowak, I. (2007). Poisson Wavelets on the Sphere. J. Fourier Anal. Appl., 13, 4, 405 -- 420. [15] Kerkyacharian, G., Nickl, R. and Picard, D. (2012). Concentration inequalities and confidence bands for needlet density estimators on compact homogeneous manifolds. Probab. Theory Related Fields, 153, 1, 363 -- 404. [16] Lan, X. and Marinucci, D. (2009). On the Dependence Structure of Wavelet Coefficients for Spherical Random Fields. Stochastic Process. Appl., 119, 110, 3749 -- 3766. [17] Marinucci, D. and Peccati, G. (2011). Random Fields on the Sphere. Representation, Limit Theorem and Cosmological Applications. Cambridge University Press. [18] Mayeli, A. (2010). Asymptotic Uncorrelation for Mexican Needlets. J. Math. Anal. Appl., 363, 1, 336 -- 344. [19] McEwen, J.D., Hobson, M.P. and Lasenby A.N. (2008). Optimal filters on the sphere. IEEE Trans. Sig. Proc., 56, 8, 3813 -- 3823. [20] McEwen, J.D., Vielva, P., Wiaux, Y., Barreiro, R.B., Cayon, L., Hobson, M.P., Lasenby, A.N., Martinez-Gonzalez, E. and Sanz, J. (2007). Cosmological Applications of a Wavelet Analysis on the Sphere. J. Fourier Anal. Appl., 13, 495 -- 510. [21] McEwen, J.D., Leistedt, B., Büttner, M., Peiris, H. V. and Wiaux, Y.(2015). Directional spin wavelets on the sphere. IEEE Trans. Sig. Proc., submitted. [22] McEwen, J.D., Durastanti C., and Wiaux Y. (2016). Localisation of directional scale-discretised wavelets on the sphere. Applied Comput. Harm. Anal., in press. [23] Chan, J. Y. H., Leistedt, B., Kitching, T. D. and McEwen J. D. (2015). Second-generation curvelets on the sphere. IEEE Trans. Sig. Proc., in press, 2015. [24] Narcowich, F.J., Petrushev, P. and Ward, J.D. (2006). Localized Tight Frames on Spheres. SIAM J. Math. Anal. 38, 574 -- 594. [25] Narcowich, F.J., Petrushev, P. and Ward, J.D. (2006). Decompo- sition of Besov and Triebel-Lizorkin Spaces on the Sphere. J. Funct. Anal., 238, 2, 530 -- 564. [26] Petrushev, P. and Xu,Y.(2008). Localized Polynomial Frames on the Ball. Constr. Approx., 27, 121 -- 148. 22 [27] Scodeller, S., Rudjord, O. Hansen, F.K., Marinucci, D., Geller, D. and Mayeli, A. (2011). Introducing Mexican needlets for CMB analysis: Issues for practical applications and comparison with standard needlets. Astrophys. J., 733, 2, 121. [28] Starck, J.-L , Moudden, Y. , Abrial P. and Nguyen M. (2006). Wavelets, Ridgelets and Curvelets on the Sphere. A&A, 446, 1191 -- 1204. [29] Stein, E., Weiss, G. (1971). Introduction to Fourier Analysis on Eu- clidean Spaces. Princeton University Press. [30] Szego, G. (1975). Orthogonal Polynomials. American Mathematical So- ciety, 23. [31] Wiaux, Y., McEwen, J.D. and Vielva, P. (2007). Complex Data Processing: Fast Wavelet Analysis on the Sphere. J. Fourier Anal. Appl., 13, 4 477 -- 494. 23
1606.04874
1
1606
2016-06-15T17:32:45
On certain product of Banach modules
[ "math.FA" ]
Let $A$ and $B$ be Banach algebras and let $B$ be an algebraic Banach $A-$bimodule. Then the $\ell^1-$direct sum $A\times B$ equipped with the multiplication $$(a_1,b_1)(a_2,b_2)=(a_1a_2,a_1\cdot b_2+b_1\cdot a_2+b_1b_2),~~ (a_1, a_2\in A, b_1, b_2\in B)$$ is a Banach algebra denoted by $A\bowtie B$. Module extension algebras, Lau product and also the direct sum of Banach algebras are the main examples satisfying this framework. We obtain characterizations of bounded approximate identities, spectrum, and topological center of this product. This provides a unified approach for obtaining some known results of both module extensions and Lau product of Banach algebras.
math.FA
math
The Extended Abstracts of The 4th Seminar on Functional Analysis and its Applications 2-3rd March 2016, Ferdowsi University of Mashhad, Iran ON CERTAIN PRODUCT OF BANACH MODULES MOHAMMAD RAMEZANPOUR1 ∗ 1 School of Mathematics and Computer Science, Damghan University, P. O. Box 36716, Damghan 41167, Iran. [email protected] Abstract. Let A and B be Banach algebras and let B be an algebraic Banach A−bimodule. Then the ℓ1−direct sum A × B equipped with the multiplication (a1, b1)(a2, b2) = (a1a2, a1·b2+b1·a2+b1b2), (a1, a2 ∈ A, b1, b2 ∈ B) is a Banach algebra denoted by A ⊲⊳ B. Module extension algebras, Lau product and also the direct sum of Banach algebras are the main examples satisfying this framework. We obtain characteriza- tions of bounded approximate identities, spectrum, and topological center of this product. This provides a unified approach for obtain- ing some known results of both module extensions and Lau product of Banach algebras. 1. Introduction Let A and B be Banach algebra and B is a Banach A−bimodule, we say that B is an algebraic Banach A−bimodule if a · (b1b2) = (a · b1)b2, (b1b2) · a = b1(b2 · a), (b1 · a)b2 = b1(a · b2), for each b1, b2 ∈ B and a ∈ A. Then the Cartesian product A × B with the algebra multiplication (a1, b1)(a2, b2) = (a1a2, a1 · b2 + b1 · a2 + b1b2), 2010 Mathematics Subject Classification. Primary 46H05; Secondary 46H20, 46H25, 46H05. Key words and phrases. Banach algebras, extensions, topological center. ∗ Speaker. 1 2 M. RAMEZANPOUR and with the norm k(a, b)k = kak + kbk becomes a Banach algebra provided ka · bk ≤ kakkbk, which we denote it by A ⊲⊳ B. We note that if we identify A × {0} with A, and {0} × B with B, in A ⊲⊳ B, then B is a closed ideal while A is a closed subalgebra of A ⊲⊳ B, and (A ⊲⊳ B)/B is isometrically isomorphic to A. In other words, A ⊲⊳ B is a strongly splitting Banach algebra extension of A by B. Besides giving a new method of constructing Banach algebras, the product ⊲⊳ has relevance with the following known products. (a) (Direct product of two Banach algebras) Let A and B be Banach algebra. If we define a · b = b · a = 0, then B is an algebraic Banach A-bimodule and (a1, b1)(a2, b2) = (a1a2, b1b2). Therefore, A ⊲⊳ B is the direct product A ×1 B. (b) (The module extensions) Let X be a Banach A-bimodule. Define x1x2 = 0, then X is an algebraic Banach A-bimodule, and (a1, x1)(a2, x2) = (a1a2, a1 · x2 + x1 · a2). Therefore, A ⊲⊳ X is the module extension A ⊕1 X. Module exten- sions are known as a rich source of (counter-)examples in various situations in abstract harmonic analysis and functional analysis, [5]. (c) (θ−Lau product of Banach algebras) Let A and B be Banach al- gebra and θ ∈ ∆(A), the set of all non zero multiplicative lin- ear functional on A. Then B with a module actions given by a · b = b · a = θ(a)b is an algebraic Banach A-bimodule and (a1, b1)(a2, b2) = (a1a2, θ(a1)b2 + θ(a2)b1 + b1b2). Thus A ⊲⊳ B is the θ-Lau product A θ× B. This product was intro- duced by Lau [2] for certain class of Banach algebras and followed by Sangani Monfared [4] for the general case. An elementary very familiar example is the case that A = C with θ as the identity character i that we get the unitization B♯ = C θ× B of B. (d) (T −Lau product of Banach algebras) Let A and B be Banach al- gebra and T : A → B be an algebra homomorphism with kT k ≤ 1. Define a · b = T (a)b = b · a. Then B is an algebraic Banach A- bimodule and (a1, b1)(a2, b2) = (a1a2, T (a1)b2 + b1T (a2) + b1b2). Thus A ⊲⊳ B is the T −Lau product A T× B. This type of product was first introduced by Bhatt and Dabhi for the case where B is commutative and was extended by Javanshiri and Nemati for the general case [1]. ON CERTAIN PRODUCT OF BANACH MODULES 3 The purpose of the present note is to determine the Gelfand space of A ⊲⊳ B which turns out to be non trivial even though A ⊲⊳ B need not be commutative and to discuss the topological center of A ⊲⊳ B. These topics are central to the general theory of Banach algebras. 2. Main results Let B be a Banach A−bimodule, we recall that B is called symmetric if a · b = b · a for all a ∈ A and b ∈ B. We start with the following propositions which characterize the basic properties of A ⊲⊳ B in terms of A and B. These results extend related results in [1, 4]. Proposition 2.1. [3] Let B be an algebraic Banach A-bimodule. Then A ⊲⊳ B is commutative if and only if B is a symmetric Banach A- bimodule and both A and B are commutative. Proposition 2.2. [3] Let B be an algebraic Banach A-bimodule. Then (a0, b0) is an identity for A ⊲⊳ B if and only if a0 is an identity for A, b0 · a = a · b0 = 0 for all a ∈ A and a0 · b + b0b = b · a0 + bb0 = b for all b ∈ B. Proposition 2.3. [3] Let B be an algebraic Banach A-bimodule. Then {(aα, bα)} is a bounded left approximate identity for A ⊲⊳ B if and only if {aα} is a bounded left approximate identity for A, kbα · ak → 0 for all a ∈ A and aα · b + bαb → b for all b ∈ B. The dual of the space A ⊲⊳ B can be identified with A∗ × B ∗ in the natural way (ϕ, ψ)(a, b) = ϕ(a) + ψ(b). The dual norm on A∗ × B ∗ is of course the maximum norm k(ϕ, ψ)k = max{kϕk, kψk}. The following result identifies the Gelfand space of A ⊲⊳ B. This is a generalization of [1, Theorem 2.2] and [4, Proposition 2.4]. Proposition 2.4. [3] Let B be an algebraic Banach A-bimodule. If E := {(ϕ, 0) : ϕ ∈ ∆(A)} and F := {(ϕ, ψ) : ϕ ∈ ∆(A) ∪ {0}, ψ ∈ ∆(B), and a · ψ = ψ · a = ϕ(a)ψ ∀a ∈ A}, then ∆(A ⊲⊳ B) = E ∪ F . Corollary 2.5. Let A and B be commutative Banach algebras and B is an algebraic Banach A-bimodule which is also symmetric. Then A ⊲⊳ B is semisimple if and only if both A and B are semisimple. Let X be a Banach A-bimodule, for a ∈ A, x ∈ X, x∗ ∈ X ∗, a∗∗ ∈ A∗∗ and x∗∗ ∈ X ∗∗ we define (x∗∗ ◦ a∗∗)(x∗) = x∗∗(a∗∗ ◦ x∗), (a∗∗ ◦ x∗)(x) = a∗∗(x∗ ◦ x), (x∗ ◦ x)(a) = x∗(x · a), (a∗∗ ◦ x∗∗)(x∗) = a∗∗(x∗∗ ◦ x∗) (x∗∗ ◦ x∗)(a) = x∗∗(x∗ ◦ a) (x∗ ◦ a)(x) = x∗(a · x). 4 M. RAMEZANPOUR Clearly, for each a∗∗ ∈ A∗∗ and x∗∗ ∈ X ∗∗ the mappings b∗∗ → b∗∗ ◦x∗∗ : A∗∗ → X ∗∗ and y ∗∗ → y ∗∗ ◦ a∗∗ : X ∗∗ → X ∗∗ are w∗-w∗-continuous. The first topological centres of module actions of A on X may therefore be defined as Z 1 Z 1 A(X ∗∗) = {x∗∗ ∈ X ∗∗ : a∗∗ → x∗∗ ◦ a∗∗ is w∗-w∗-continuous}, X(A∗∗) = {a∗∗ ∈ A∗∗ : x∗∗ → a∗∗ ◦ x∗∗ is w∗-w∗-continuous}. If we consider X = A with the natural A-bimodule structure then we obtain the first Arens product on A∗∗. In this case we write Z 1(A∗∗) instead Z 1 A(A∗∗). The Banach algebra A is called Arens regular if Z 1(A∗∗) = A∗∗. To state our next result we note that if B is an algebraic Banach A-bimodule then B ∗∗ is an algebraic Banach A∗∗-bimodule, when A∗∗ and B ∗∗ are equipped with their first Arens products. Theorem 2.6. [3] Let B be an algebraic Banach A-bimodule. (1) Suppose that A∗∗, B ∗∗, and (A ⊲⊳ B)∗∗ are equipped with their first Arens products. Then (A ⊲⊳ B)∗∗ ∼= A∗∗ ⊲⊳ B ∗∗ (2) If B is Arens regular then (isometric isomorphism). Z 1((A ⊲⊳ B)∗∗) = (cid:0)Z 1(A∗∗) ∩ Z 1 B(A∗∗)(cid:1) × Z 1 A(B ∗∗) References 1. H. Javanshiri and M. Nemati, On a certain product of Banach algebras and some of its properties, Proc. Rom. Acad. Ser. A, 15 (2014), no. 3, 219 -- 227. 2. A. T.-M. Lau, Analysis on a class of Banach algebras with applications to har- monic analysis on locally compact groups and semigroups, Fund. Math. 118 (1983), no. 3, 161 -- 175. 3. M. Ramezanpour, Some cohomological properties of certain product of Banach modules, submitted. 4. M. Sangani Monfared, On certain products of Banach algebras with applications to harmonic analysis, Studia Math., 178 (2007), no. 3, 277 -- 294. 5. Y. Zhang, Weak amenability of module extensions of Banach algebras, Trans. Amer. Math. Soc., 354 (2002), 4131 -- 4151.
1510.03535
1
1510
2015-10-13T05:14:04
Idempotents of small norm
[ "math.FA" ]
Let $\Gamma$ be a locally compact group. We answer two questions left open in [7] and [9]: i) For abelian $\Gamma$, we prove that if $\chi_S \in B(\Gamma)$ is an idempotent with norm $\left\|\chi_S \right\| < \frac{4}{3}$, then $S$ is the union of two cosets of an open subgroup of $\Gamma$. ii) For general $\Gamma$, we prove that if $\chi_S \in M_{cb}A(\Gamma)$ is an idempotent with norm $\left\| \chi_S \right\|_{cb} < \frac{1 + \sqrt{2}}{2}$, then $S$ is an open coset in $\Gamma$.
math.FA
math
IDEMPOTENTS WITH SMALL NORMS JAYDEN MUDGE AND HUNG LE PHAM Abstract. Let Γ be a locally compact group. We answer two questions left open in [7] and [9]: (i) For abelian Γ, we prove that if χS ∈ B(Γ) is an idempotent with norm 3 , then S is the union of two cosets of an open subgroup of Γ. (ii) For general Γ, we prove that if χS ∈ McbA(Γ) is an idempotent with kχSk < 4 norm kχS kcb < 1+√2 2 , then S is an open coset in Γ. 1. Introduction In his 1968 papers, Saeki determined idempotent measures on a locally compact abelian group G with small norms. These are equivalent to determining idempotent functions in the Fourier -- Stieltjes algebras B(Γ) on a locally compact abelian group Γ with small norms (where Γ and G could be taken as Pontryagin duals of each other). The statements of Saeki's results in the Fourier -- Stieltjes setting are: Theorem 1.1 (Saeki). Let Γ be a locally compact abelian group, and let ϕ be an idempotent function in B(Γ) so that ϕ = χS for some nonempty S ⊆ Γ. Then (i) ([6]) If kϕk < 1+√2 (ii) ([7]) If kϕk ∈ (1, 2 4 subgroup of Γ but is not a coset itself. , then S is an open coset of Γ. √17+1 ), then S is the union of two cosets of an open For abelian Γ, it is well-known (see [5, page 73]) that if S is an open coset of Γ, then kχSk = 1, and whereas if S is the union of two cosets of an open subgroup of Γ but is not a coset itself, then 2 q sin(π/2q) if q is odd kχSk =  2 (1.1) if q is even if q = ∞ where q is the "relative order" of the two cosets forming S. The largest value in (1.1) is 4/3 when q = 3 and the smallest one is 1+√2 2 when q = 4. In particular, the number 1+√2 in Theorem 1.1 (i) is sharp. q tan(π/2q) 4 π 2 The paper [7] asked whether or not the interval (1, could be increased to (1, 4 that the interval (1, 4 ) in Theorem 1.1 (ii) 3 ), and we answer this in affirmative in Theorem 3.4. Note 3 ) is sharp because of the discussion in the previous paragraph 4 √17+1 2010 Mathematics Subject Classification. 43A10, 43A30. Key words and phrases. Idempotent, measure, Fourier algebra, Fourier -- Stieltjes algebra, com- pletely bounded multiplier. The first author is supported by the Victoria University of Wellington Masters by thesis Schol- arship 2015. 1 2 JAYDEN MUDGE AND HUNG LE PHAM and also since there are idempotents χS of B(Γ) with kχSk = 4/3 but S is not any union of two cosets of open subgroup of Γ (see the last paragraph of [7]). Lesser is known about idempotents in B(Γ) with small norms for general locally compact group Γ. Ilie and Spronk [3] proved that χS is an idempotent in B(Γ) with kχSk = 1 if and only if S is an open coset of Γ. More generally, Stan proved the following. Theorem 1.2 (Stan [9]). Let Γ be a locally compact group, and let ϕ be an idempo- tent function in McbA(Γ) so that ϕ = χS for some nonempty S ⊆ Γ. If kϕkcb < 2√3 , then S is an open coset of Γ, and in which case kϕkcb = 1. Here McbA(Γ) is the completely bounded multiplier algebra McbA(Γ) of the Fourier algebra A(Γ) and is defined as follows. Since the Fourier algebra A(Γ) is the predual of the group von Neumann algebra VN(Γ), it has the canonical operator space structure, which makes it a completely contractive operator algebra (see the monograph [2] for more details). The completely bounded multiplier algebra McbA(Γ) of A(Γ) consists of those continuous functions ϕ : Γ → C such that the mapping f 7→ ϕ · f, A(Γ) → A(Γ) , is completely bounded, and its completely bounded norm is denoted as kϕkcb (whereas the Fourier -- Stieltjes norm on Γ will be simply denoted as k·k in this paper). In general, we have B(Γ) ⊆ McbA(Γ) , with a decreasing of norms , but, for amenable locally compact groups Γ, B(Γ) = McbA(Γ) , isometrically . Thus an idempotent of B(Γ) with a small norm is always an idempotent of McbA(Γ) with a small(er) norm. In Theorem 2.2, we increase the number 2√3 bound of 1+√2 Saeki, Theorem 1.1 (i), to general locally compact groups. 2 , and so obtaining a generalisation of the first mentioned result of in Stan's Theorem 1.2 to the sharp 2. Idempotents of McbA(Γ) with norm lesser than 1+√2 2 of McbA(Γ) with kχSkcb < 1+√2 In this section, let Γ be any locally compact group, and let χS be an idempotent . Our aim is to show that S is an open coset of Γ. It is obvious that S is open, and so, it remains to show that S is a coset of Γ. By [8, Corollary 6.3 (i)], it is sufficient for us to consider the case where Γ is discrete. 2 We first make a simple observation. Lemma 2.1. For any s ∈ S and t ∈ Γ, if st ∈ S (resp. ts ∈ S), then stn ∈ S for every n ∈ Z (resp. tns ∈ S for every n ∈ Z). Proof. By translation, we may (and shall) suppose that s = e, the identity of Γ. Consider Γ0 be the (abelian) group generated by t, then kχS∩Γ0k = kχS∩Γ0kcb ≤ kχSkcb < 1 + √2 . 2 So by Saeki's Theorem 1.1 (i), we see that S ∩ Γ0 = Γ0. This gives the lemma. (cid:3) IDEMPOTENTS WITH SMALL NORMS 3 To get more information out of the assumption on kχSkcb, we shall follow in the footsteps of [9] and use the connection shown in [1] between the norm k·kcb of McbA(Γ) and the Schur multiplier norms described below. Denote by K0 the space of matrices that have only finitely many nonzero entries whose rows and columns are indexed by Γ. Then K0 is identified with a subspace of B(ℓ2(Γ)). Recall that the Schur multiplication of two matrices A and X, indexed by Γ, is defined as and for each matrix A, indexed by Γ, its Schur multiplier norm is (A • X)(s, t) := A(s, t)X(s, t) kAkSchur := sup(kA • XkB(ℓ2(Γ)) kXkB(ℓ2(Γ)) (s, t ∈ Γ) , : X ∈ K0) . (2.1) Of course, this discussion works for any index set Γ, and a particular matrix that is useful for us is the following 3 × 3 matrix F0 := 1 1 1 1 1  .  1 0 0 √2 √2 2 √2 1 −1 √2 −1  1  2  and the vector ξ := 1  Using the orthogonal matrix U := 1 √2 1 1 we see that  , 1 0 √26 4 > 1 + √2 . 2 kF0kSchur ≥ kA • UkB(ℓ2) ≥ k(A • U )ξkℓ2 = kξkℓ2 As a matter of fact, it is proved in [4, Proposition 5.1(8)] that kF0kSchur = 9 7 , but the above simple calculation is sufficient for our purpose. Hence, any matrix A that has a submatrix of the form F0 in (2.1) must satisfy kAkSchur > 1 + √2 . 2 Returning to our problem on the group Γ, each function ϕ : Γ → C defines a matrix Mϕ, indexed by Γ, by setting An important fact shown in [1] is that Mϕ(s, t) := ϕ(s−1t) (s, t ∈ Γ) . (2.2) kϕkcb = kMϕkSchur . Hence, the previous paragraph implies that MχS cannot have (2.1) as a submatrix. Our main result of this section is the following. Theorem 2.2. Let Γ be a locally compact group, and let ϕ be an idempotent func- , then tion in McbA(Γ) so that ϕ = χS for some nonempty S ⊆ Γ. If kϕkcb < 1+√2 2 S is an open coset of Γ. Proof. As discussed above, we may (and shall) suppose that Γ is discrete. Also, applying a translation if necessary, we suppose that e ∈ S. So it remains to prove that S is a subgroup of Γ. By Lemma 2.1, we see that if u ∈ S, then un ∈ S for every n ∈ Z. Thus it remains to show that S is closed under multiplication. 4 JAYDEN MUDGE AND HUNG LE PHAM We next claim that if u, v ∈ S, then either uv ∈ S or vu ∈ S. Indeed, assume towards a contradiction that both uv /∈ S and vu /∈ S. Then the submatrix of MχS with rows e, u−1, v−1 and colums e, u, v is χS(u) χS(e) χS(v) χS(s) χS(u2) χS(uv)  χS(v) χS(vu) χS(v2)  =   1 1 1 1 1 1 0 0 1  by the previous paragraph. This contradicts the previous discussion. Finally, suppose that u, v ∈ S, the proof is completed if we can show that uv ∈ S. The claim shows that either uv ∈ S or vu ∈ S. Assume the latter holds, then from Lemma 2.1 with s = v and t = u, we obtain that vu−1 ∈ S. Since we must have u−1 ∈ S, this in turn implies, by a similar argument, that v−1u−1 ∈ S. But then, since v−1u−1 = (uv)−1, we must have uv ∈ S. Hence, in any case, uv ∈ S, and the proof is completed. (cid:3) 3. Idempotents of B(Γ) with norm lesser than 4 3 , for abelian Γ In this section, let Γ = (Γ, 0, +) be a locally compact abelian group. We aim √17+1 ) to the to strengthen Saeki's Theorem 1.1 (ii) by enlarging his range of (1, optimal (1, 4 3 ). So let χS be an idempotent function in B(Γ) with kχSk ≤ 4 3 . Actually, in [7], Saeki works with idempotents of the measure algebra M(G) on a locally compact abelian group G, and so, our Γ and his G could be considered as the Pontryagin's duals of each other. Thus B(Γ) ∼= M(G) isometrically, and we denote by µ the idempotent measure in M(G) that corresponds to χS. 4 As in the previous section, we may reduce our problem to the case where Γ is discrete. Thus suppose that Γ is discrete, and so G is compact. Saeki's proof of Theorem 1.1 (ii) in [7] invokes the following lemma several times. 4 . √17+1 Lemma 3.1 (Saeki). Assume as above. Suppose there exists u and v in S and w in Γ such that u + w belongs to S but neither v + w nor v − w belongs to S. Then we have kµk ≥ In the main argument, Saeki uses this lemma to show that if S is not the union of two cosets of a subgroup in Γ, then kµk ≥ . The argument used breaks the problem up into many cases, and in the cases where this lemma is not used, it is always shown that in fact kµk ≥ 4 3 . Thus if we can strengthen this lemma, then Theorem 1.1 (ii) is strengthened also. In fact, we prove the following. Lemma 3.2. Assume as above. Suppose there exists u, v ∈ S, and w ∈ Γ, such that u + w ∈ S, and v + w, v − w 6∈ S. Then kχSk = kµk ≥ 4 3 . Proof. Let us define a function f ∈ C(G) to be √17+1 4 1 2 f (x) = (x, u)[2 + 2(x, w) + (x,−w)] + (x, v)[2 − (x, w) − (x,−w)] . If u − w ∈ S, we get (cid:12)(cid:12)R f (x) dµ(x)(cid:12)(cid:12) = 13 2 , otherwise it will simply be 6. Next, we calculate the uniform norm fG of f , by taking x ∈ G, and set (x, w) =: eiθ. Then we see e−iθ(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)2 − eiθ − e−iθ(cid:12)(cid:12) + 10 cos(θ) + 4 cos2(θ) + 2 − 2 cos(θ) = f (x) ≤(cid:12)(cid:12)(cid:12)(cid:12) =r 25 2 + 2eiθ + 1 2 9 2 4 . IDEMPOTENTS WITH SMALL NORMS Thus fG ≤ 9 2 . Hence This proves our lemma. kµk ≥ (cid:12)(cid:12)RG f (x) dµ(x)(cid:12)(cid:12) fG 6 9/2 ≥ = 4 3 . 5 (cid:3) Now we can get our desired result. Theorem 3.3. Let Γ be a locally compact abelian group, and let ϕ be an idempotent function in B(Γ) so that ϕ = χS for some nonempty S ⊆ Γ. If kϕk ∈ (1, 4 3 ), then S is the union of two cosets of some open subgroup of Γ but is not a coset itself. (cid:3) Proof. This follows from the previous discussion with the argument of [7]. (cid:3) Equivalently, the above translates into the following: Theorem 3.4. Let G be a locally compact abelian group, and let µ be an idempotent measure on G with kµk ∈ (1, 4 3 ). Then dµ(x) = [(−x, γ1) + (−x, γ2)] dm(x) where m is the Haar measure of some compact subgroup H of G, and γ1, γ2 are distinct characters of H. (cid:3) References [1] M. Bozejko, and G. Fendler, 'Herz-Schur multipliers and completely bounded multipliers of the Fourier algebra of a locally compact group', Boll. Un. Mat. Ital. A (6) 3 (1984) 297 -- 302. [2] E.G. Effros, and Z.-J. Ruan, Operator spaces, (London Mathematical Society Monographs. New Series, 23, The Clarendon Press, Oxford University Press, New York).2000 [3] M. Ilie, and N. Spronk, 'Completely bounded homomorphisms of the Fourier algebras', J. Functional Analysis 225 (2005) 480 -- 499. [4] R. H. Levene, 'Norms of idempotent Schur multipliers', New York J. Math. 20 (2014) 325 -- 352. [5] W. Rudin, Fourier analysis on groups, (Interscience Tracts in Pure and Applied Mathematics, No. 12, Interscience Publishers (a division of John Wiley and Sons), New York-London).1962 [6] S. Saeki, 'On norms of idempotent measures', Proc. Amer. Math. Soc. 19 (1968) 600 -- 602. [7] S. Saeki, 'On norms of idempotent measures. II', Proc. Amer. Math. Soc. 19 (1968) 367 -- 371. [8] N. Spronk, 'Measurable Schur multipliers and completely bounded multipliers of the Fourier algebras', Proc. London Math. Soc. (3) 89 (2004) 161 -- 192. [9] A.-M. P. Stan, 'On idempotents of completely bounded multipliers of the Fourier algebra A(G)', Indiana Univ. Math. J. 58 (2009) 523 -- 535. E-mail address: [email protected] School of Mathematics and Statistics, Victoria University of Wellington, Welling- ton, New Zealand E-mail address: [email protected] School of Mathematics and Statistics, Victoria University of Wellington, Welling- ton, New Zealand
1605.09340
3
1605
2017-02-06T15:19:17
Fourier multiplier theorems involving type and cotype
[ "math.FA" ]
In this paper we develop the theory of Fourier multiplier operators $T_{m}:L^{p}(\mathbb{R}^{d};X)\to L^{q}(\mathbb{R}^{d};Y)$, for Banach spaces $X$ and $Y$, $1\leq p\leq q\leq \infty$ and $m:\mathbb{R}^d\to \mathcal{L}(X,Y)$ an operator-valued symbol. The case $p=q$ has been studied extensively since the 1980's, but far less is known for $p<q$. In the scalar setting one can deduce results for $p<q$ from the case $p=q$. However, in the vector-valued setting this leads to restrictions both on the smoothness of the multiplier and on the class of Banach spaces. For example, one often needs that $X$ and $Y$ are UMD spaces and that $m$ satisfies a smoothness condition. We show that for $p<q$ other geometric conditions on $X$ and $Y$, such as the notions of type and cotype, can be used to study Fourier multipliers. Moreover, we obtain boundedness results for $T_m$ without any smoothness properties of $m$. Under smoothness conditions the boundedness results can be extrapolated to other values of $p$ and $q$ as long as $\tfrac{1}{p}-\tfrac{1}{q}$ remains constant.
math.FA
math
FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE JAN ROZENDAAL AND MARK VERAAR Abstract. In this paper we develop the theory of Fourier multiplier operators Tm : Lp(Rd; X) → Lq(Rd; Y ), for Banach spaces X and Y , 1 ≤ p ≤ q ≤ ∞ and m : Rd → L(X, Y ) an operator-valued symbol. The case p = q has been studied extensively since the 1980's, but far less is known for p < q. In the scalar setting one can deduce results for p < q from the case p = q. However, in the vector-valued setting this leads to restrictions both on the smoothness of the multiplier and on the class of Banach spaces. For example, one often needs that X and Y are UMD spaces and that m satisfies a smoothness condition. We show that for p < q other geometric conditions on X and Y , such as the notions of type and cotype, can be used to study Fourier multipliers. Moreover, we obtain boundedness results for Tm without any smoothness properties of m. Under smoothness conditions the boundedness results can be extrapolated to other values of p and q as long as 1 p q remains constant. − 1 7 1 0 2 b e F 6 ] . A F h t a m [ 3 v 0 4 3 9 0 . 5 0 6 1 : v i X r a 1. Introduction Fourier multiplier operators play a major role in analysis and in particular in the theory of partial differential equations. Such operators are of the form Tm(f ) = F−1(mF f ), where F denotes the Fourier transform and m is a function on Rd. Usually one is interested in the boundedness of Tm : Lp(Rd) → Lq(Rd) with 1 ≤ p ≤ q ≤ ∞ (the case p > q is trivial by [28, Theorem 1.1]). The class of Fourier multiplier operators coincides with the class of singular integral operators of convolution type f 7→ K ∗ f , where K is a tempered distribution. The simplest class of examples of Fourier multipliers can be obtained by taking p = q = 2. Then Tm is bounded if and only if m ∈ L∞(Rd), and kTmkL(L2(Rd)) = kmkL∞(Rd). For p = q = 1 and p = q = ∞ one obtains only trivial multipliers, namely Fourier transforms of bounded measures. The case where p = q ∈ (1,∞) \ {2} is highly nontrivial. In general only sufficient conditions on m are known that guarantee that Tm is bounded, although also here it is necessary that m ∈ L∞(Rd). In the classical paper [28] Hormander studied Fourier multipliers and singular integral operators of convolution type. In particular, he showed that if 1 < p ≤ 2 ≤ 2010 Mathematics Subject Classification. Primary: 42B15; Secondary: 42B35, 46B20, 46E40, 47B38. Key words and phrases. Operator-valued Fourier multipliers, type and cotype, Fourier type, Hormander condition, γ-boundedness. The second author is supported by the VIDI subsidy 639.032.427 of the Netherlands Organi- sation for Scientific Research (NWO). 1 2 JAN ROZENDAAL AND MARK VERAAR r = 1 p − 1 q . Tm : Lp(Rd) → Lq(Rd) is bounded if m ∈ Lr,∞(Rd) with 1 q < ∞, then (1.1) Here Lr,∞(Rd) denotes the weak Lr-space. In particular, every m with m(ξ) ≤ Cξ−d/r satisfies m ∈ Lr,∞(Rd). It was also shown that the condition p ≤ 2 ≤ q is necessary here. More precisely, if there exists a function F such that {F > 0} has nonzero measure and for all m : Rd → R with m ≤ F, Tm : Lp(Rd) → Lq(Rd) is bounded, then p ≤ 2 ≤ q. Hormander also introduced an integral/smoothness condition on the kernel K which allows one to extrapolate the boundedness of Tm from Lp0 (Rd) to Lq0(Rd) for some 1 < p0 ≤ q0 < ∞ to boundedness of Tm from Lp(Rd) to Lq(Rd) for all p0 − 1 1 < p ≤ q < ∞ satisfying 1 . This led to extensions of the theory of Calder´on and Zygmund in [13]. In the case p0 = q0 it was shown that the smoothness condition on the kernel K can be translated to a smoothness condition on the multiplier m which is strong enough to deduce the classical Mihlin multiplier theorem. From here the field of harmonic analysis has quickly developed itself and this development is still ongoing. We refer to [24, 25, 36, 54] and references therein for a treatment and the history of the subject. p − 1 q = 1 q0 In the vector-valued setting it was shown in [6] that the extrapolation results of Hormander for p = q still holds. However, there is a catch: m ∈ L∞(Rd) unless X is a Hilbert space. • even for p = q = 2 one does not have Tm ∈ L(L2(Rd; X)) for general In [12] it was shown that Tm ∈ L(Lp(Rd; X)) for m(ξ) := sign(ξ) if X satisfies the so-called UMD condition. In [10] it was realized that this yields a characterization of the UMD property. In [11], [43], [63] versions of the Littlewood–Paley theorem and the Mihlin multiplier theorem were established in the UMD setting. These are very useful for operator theory and evolution equations (see for example [18]). In the vector-valued setting it is rather natural to allow m to take values in the space L(X, Y ) of bounded operators from X to Y . Pisier and Le Merdy showed that the natural analogues of the Mihlin multiplier theorem do not extend to this setting unless X has cotype 2 and Y has type 2 (a proof was published only later on in [4]). On the other hand there was a need for such extensions as it was realized that multiplier theorems with operator-valued symbols are useful in the stability theory and the regularity theory for evolution equations (see [2,27,61]). The missing ingredient for a natural analogue of the Mihlin multiplier theorem turned out to be R-boundedness, which is a strengthening of uniform boundedness (see [9, 14]). In [62] it was shown that Mihlin's theorem holds for m : R → L(X) if the sets {m(ξ) ξ ∈ R \ {0}} and {ξm′(ξ) ξ ∈ R \ {0}} are R-bounded. Conversely, the R-boundedness of {m(ξ) ξ ∈ R \ {0}} is also nec- essary. These results were used to characterize maximal Lp-regularity, and were then used by many authors in evolution equations, partial differential equations, op- erator theory and harmonic analysis (see the surveys and lecture notes [2,16,34,38]). A generalization to multipliers on Rd instead of R was given in [26] and [55], but in some cases one additionally needs the so-called property (α) of the Banach space (which holds for all UMD lattices). Improvements of the multiplier theorems un- der additional geometric assumptions have been studied in [23] and [53] assuming Fourier type and in [32] assuming type and cotype conditions. FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 3 In this article we complement the theory of operator-valued Fourier multipliers by studying the boundedness of Tm from Lp(Rd; X) to Lq(Rd; Y ) for p < q. One r m(ξ) of our main results is formulated under γ-boundedness assumptions on {ξ ξ ∈ Rd \{0}}. We note that R-boundedness implies γ-boundedness (see Subsection 2.4). The result is as follows (see Theorem 3.18 for the proof): Theorem 1.1. Let X be a Banach space with type p0 ∈ (1, 2] and Y a Banach space with cotype q0 ∈ [2,∞), and let p ∈ (1, p0), q ∈ (q0,∞). Let r ∈ [1,∞] be such that 1 q . If m : Rd \ {0} → L(X, Y ) is X-strongly measurable and p − 1 r = 1 d d (1.2) is γ-bounded, then Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. Moreover, if p0 = 2 (or q0 = 2), then one can also take p = 2 (or q = 2). r m(ξ) ξ ∈ Rd \ {0}} ⊆ L(X, Y ) {ξ The condition p ≤ 2 ≤ q cannot be avoided in such results (see below (1.1)). Note that no smoothness on m is required. Theorem 1.1 should be compared to the sufficient condition in (1.1) due to Hormander in the case where X = Y = C. We will give an example which shows that the γ-boundedness condition (1.2) cannot be avoided in general. Moreover, we obtain several converse results stating that type and cotype are necessary. We note that, in case m is scalar-valued and X = Y , the γ-boundedness as- sumption in Theorem 1.1 reduces to the uniform boundedness of (1.2). Even in this setting of scalar multipliers our results appear to be new. In Theorem 3.21 we obtain a variant of Theorem 1.1 for p-convex and q-concave Banach lattices, where one can take p = p0 and q = q0. In [50] we will deduce multiplier results similar to Theorem 1.1 in the Besov scale, where one can let p = p0 and q = q0 for Banach spaces X and Y with type p and cotype q. A vector-valued generalization of (1.1) is presented in Theorem 3.12. We show that if X has Fourier type p0 > p and Y has Fourier type q′0 > q′, then kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr,∞(Rd) , where 1 p − 1 q . We show that in this result the Fourier type assumption is It should be noted that for many spaces (including all Lr-spaces for necessary. r ∈ [1,∞) \ {2}), working with Fourier type yields more restrictive results in terms of the underlying parameters than working with type and cotype (see Subsection 2.2 for a discussion of the differences between Fourier type and (co)type). r = 1 The exponents p and q in Theorem 1.1 are fixed by the geometry of the underlying Banach spaces. However, Corollary 4.2 shows that under smoothness conditions on the multiplier, one can extend the boundedness result to all pairs (p, q) satisfying 1 < p ≤ q < ∞ and 1 r . Here the required smoothness depends on the Fourier type of X and Y and on the number r ∈ (1,∞]. We note that even in the case where X = Y = C, for p < q we require less smoothness for the extrapolation than in the classical results (see Remark 4.4). p − 1 p − 1 q = 1 q = 1 We will mainly consider multiplier theorems on Rd. There are two exceptions. In Remark 3.11 we deduce a result for more general locally compact groups. Moreover, in Proposition 3.4 we show how to transfer our results from Rd to the torus Td. This result appears to be new even in the scalar setting. As an application of the latter we show that certain irregular Schur multipliers with sufficient decay are bounded on the Schatten class C p for p ∈ (1,∞). 4 JAN ROZENDAAL AND MARK VERAAR We have pointed out that questions about operator-valued Fourier multiplier theorems were originally motivated by stability and regularity theory. We have already successfully applied our result to stability theory of C0-semigroups, as will be presented in a forthcoming paper [51]. In [49] the first-named author has also applied the Fourier multiplier theorems in this article to study the H∞-calculus for generators of C0-groups. Other potential applications could be given to the theory of dispersive equations. For instance the classical Strichartz estimates can be viewed as operator-valued Lp- Lq-multiplier theorems. Here the multipliers are often not smooth, as is the case in our theory. More involved applications probably require extensions of our work to oscillatory integral operators, which would be a natural next step in the research on vector-valued singular integrals from Lp to Lq. This article is organized as follows. In Section 2 we discuss some preliminaries on the geometry of Banach spaces and on function space theory. In Section 3 we introduce Fourier multipliers and prove our main results on Lp-Lq-multipliers in the vector-valued setting. In Section 4 we present an extension of the extrapolation result under Hormander–Mihlin conditions to the case p ≤ q. 1.1. Notation and terminology. We write N := {1, 2, 3, . . .} for the natural numbers and N0 := N ∪ {0}. We denote nonzero Banach spaces over the complex numbers by X and Y . The space of bounded linear operators from X to Y is L(X, Y ), and L(X) := L(X, X). The identity operator on X is denoted by IX . For p ∈ [1,∞] and (Ω, µ) a measure space, Lp(Ω; X) denotes the Bochner space of equivalence classes of strongly measurable, p-integrable, X-valued functions on Ω. Moreover, Lp,∞(Ω; X) is the weak Lp-space of all f : Ω → X for which (1.3) αλf (α) 1 kfkLp,∞(Ω;X) := sup α>0 p < ∞, p + 1 The Holder conjugate of p is denoted by p′ and is defined by 1 = 1 where λf (α) := µ({s ∈ Ω kf (s)kX > α}) for α > 0. In the case where Ω ⊆ Rd we implicitly assume that µ is the Lebesgue measure. Often we will use the shorthand notations k · kp and k · kp,∞ for the Lp-norm and Lp,∞-norm. p′ . We write ℓp for the space of p-summable sequences (xk)k∈N0 ⊆ C, and denote by ℓp(Z) the space of p-summable sequences (xk)k∈Z ⊆ C. We say that a function m : Ω → L(X, Y ) is X-strongly measurable if ω 7→ m(ω)x is a strongly measurable Y -valued map for all x ∈ X. We often identify a scalar function m : Rd → C with the operator-valued function em : Rd → L(X) given by em(ξ) := m(ξ)IX for ξ ∈ Rd. The class of X-valued rapidly decreasing smooth functions on Rd (the Schwartz functions) is denoted by S(Rd; X), and the space of X-valued tempered distribu- tions by S′(Rd; X). We write S(Rd) := S(Rd; C) and denote by h·,·i : S′(Rd; X) × S(Rd) → X the X-valued duality between S′(Rd; X) and S(Rd). The Fourier transform of a Φ ∈ S′(Rd; X) is denoted by F Φ or bΦ. If f ∈ L1(Rd; X) then A standard complex Gaussian random variable is a random variable γ : Ω → C , where (Ω, P) is a probability space and γr, γi : Ω → R are of the form γ = γr+iγi√2 bf (ξ) = F f (ξ) :=ZRd e−2πiξ·tf (t) dt (ξ ∈ Rd). FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 5 independent standard real Gaussians. A Gaussian sequence is a (finite or infinite) sequence (γk)k of independent standard complex Gaussian random variables on some probability space. We will use the convention that a constant C which appears multiple times in a chain of inequalities may vary from one occurrence to the next. 2. Preliminaries 2.1. Fourier type. We recall some background on the Fourier type of a Banach space. For these facts and for more on Fourier type see [20, 29, 46]. A Banach space X has Fourier type p ∈ [1, 2] if the Fourier transform F is (Rd; X) for some (in which case it holds for all) bounded from Lp(Rd; X) to Lp′ d ∈ N. We then write Fp,X,d := kFkL(Lp(Rd;X),Lp′ (Rd;X)). Each Banach space X has Fourier type 1 with F1,X,d = 1 for all d ∈ N. If X has Fourier type p ∈ [1, 2] then X has Fourier type r with Fr,X,d ≤ Fp,X,d for all r ∈ [1, p] and d ∈ N. We say that X has nontrivial Fourier type if X has Fourier type p for some p ∈ (1, 2]. In order to make our main results more transparent we will say that X has Fourier cotype p′ whenever X has Fourier type p. Let X be a Banach space, r ∈ [1,∞) and let Ω be a measure space. If X has Fourier type p ∈ [1, 2] then Lr(Ω; X) has Fourier type min(p, r, r′). In particular, Lr(Ω) has Fourier type min(r, r′). 2.2. Type and cotype. We first recall some facts concerning the type and cotype of Banach spaces. For more on these notions and for unexplained results see [1], [17], [30] and [41, Section 9.2]. Let X be a Banach space, (γn)n∈N a Gaussian sequence on a probability space (Ω, P) and let p ∈ [1, 2] and q ∈ [2,∞]. We say that X has (Gaussian) type p if there exists a constant C ≥ 0 such that for all m ∈ N and all x1, . . . , xm ∈ X, (2.1) γnxn(cid:13)(cid:13)(cid:13) We say that X has (Gaussian) cotype q if there exists a constant C ≥ 0 such that for all m ∈ N and all x1, . . . , xm ∈ X, kxnkq(cid:17)1/q (2.2) (cid:16)E(cid:13)(cid:13)(cid:13) (cid:16) mXn=1 kxnkp(cid:17)1/p . 2(cid:17)1/2 , 2(cid:17)1/2 ≤ C(cid:16) mXn=1 mXn=1 ≤ C(cid:16)E(cid:13)(cid:13)(cid:13) mXn=1 γnxn(cid:13)(cid:13)(cid:13) with the obvious modification for q = ∞. The minimal constants C in (2.1) and (2.2) are called the (Gaussian) type p constant and the (Gaussian) cotype q constant and will be denoted by τp,X and cq,X . We say that X has nontrivial type if X has type p ∈ (1, 2], and finite cotype if X has cotype q ∈ [2,∞). Note that it is customary to replace the Gaussian sequence in (2.1) and (2.2) by a Rademacher sequence, i.e. a sequence (rn)n∈N of independent random variables on a probability space (Ω, P) that are uniformly distributed on {z ∈ R z = 1}. This does not change the class of spaces under consideration, only the minimal constants in (2.1) and (2.2) (see [17, Chapter 12]). We choose to work with Gaussian sequences because the Gaussian constants τp,X and cq,X occur naturally here. Each Banach space X has type p = 1 and cotype q = ∞, with τ1,X = c∞,X = 1. If X has type p and cotype q then X has type r with τr,X ≤ τp,X for all r ∈ [1, p] (cid:13)(cid:13)(cid:13)(cid:16) nXk=1 (cid:16) nXk=1 xkp(cid:17)1/p(cid:13)(cid:13)(cid:13)X ≤ C(cid:16) nXk=1 ≤ C(cid:13)(cid:13)(cid:13)(cid:16) nXk=1 X(cid:17)1/q kxkkq kxkkp X(cid:17)1/p xkq(cid:17)1/q(cid:13)(cid:13)(cid:13)X 6 JAN ROZENDAAL AND MARK VERAAR and cotype s with cs,X ≤ cq,X for all s ∈ [q,∞]. A Banach space X is isomorphic to a Hilbert space if and only if X has type p = 2 and cotype q = 2, by Kwapie´n's theorem (see [1, Theorem 7.4.1]). Also, a Banach space X with nontrivial type has finite cotype by the Maurey–Pisier theorem (see [1, Theorem 11.1.14]). Let X be a Banach space, r ∈ [1,∞) and let Ω be a measure space. If X has type p ∈ [1, 2] and cotype q ∈ [2,∞) then Lr(Ω; X) has type min(p, r) and cotype max(q, r) (see [17, Theorem 11.12]). A Banach space with Fourier type p ∈ [1, 2] has type p and cotype p′ (see [30]). By a result of Bourgain a Banach space has nontrivial type if and only if it has nontrivial Fourier type (see [46, 5.6.30]). 2.3. Convexity and concavity. For the theory of Banach lattices we refer the reader to [41]. We repeat some of the definitions which will be used frequently. Let X be a Banach lattice and p, q ∈ [1,∞]. We say that X is p-convex if there exists a constant C ≥ 0 such that for all n ∈ N and all x1, . . . , xn ∈ X, (2.3) , with the obvious modification for p = ∞. We say that X is q-concave if there exists a constant C ≥ 0 such that for all n ∈ N and all x1, . . . , xn ∈ X, (2.4) , with the obvious modification for q = ∞. Every Banach lattice X is 1-convex and ∞-concave. If X is p-convex and q- concave then it is r-convex and s-concave for all r ∈ [1, p] and s ∈ [q,∞]. By [41, Proposition 1.f.3], if X is q-concave then it has cotype max(q, 2), and if X is p- convex and q-concave for some q < ∞ then X has type min(p, 2). If X is p-convex and p′-concave for p ∈ [1, 2] then X has Fourier type p, by [21, Proposition 2.2]. For (Ω, µ) a measure space and r ∈ [1,∞), Lr(Ω, µ) is an r- convex and r-concave Banach lattice. Moreover, if X is p-convex and q-concave and r ∈ [1,∞), then Lr(Ω; X) is min(p, r)-convex and max(q, r)-concave. Specific Banach lattices which we will consider are the Banach function spaces. For the definition and details of these spaces we refer to [40]. If X is a Banach function space over a measure space (Ω, µ) and Y is a Banach space, then X(Y ) consists of all f : Ω → Y such that kf (·)kY ∈ X, with the norm If f ∈ X(Lp(Rd)) for p ∈ [1,∞) and d ∈ N then we write (RRdf (t)p dt)1/p for kfkX(Y ) := kkf (·)kY kX the element of X given by (f ∈ X(Y )). (ω) :=(cid:16)ZRdf (ω)(t)p dt(cid:17)1/p (cid:16)ZRdf (t)p dt(cid:17)1/p k=1 fk ⊗ xk ∈ Lp(Rd) ⊗ X, for n ∈ N, f1, . . . , fn ∈ Lp(Rd) and k=1 fk(t)xk] of k=1 xk(ω)fk] of X(Lp(Rd)). Throughout we will identify these and consider f as an element of both Lp(Rd; X) and X(Lp(Rd)). The following lemma, proved as in [60, Theorem 3.9] by using (2.3) and (2.4) on Note that kfkX(Lp(Rd)) = k(RRdf (t)p dt)1/pkX Let f = Pn x1, . . . , xn ∈ X. Then f determines both an element [t 7→ Pn Lp(Rd; X) and an element [ω 7→ Pn (ω ∈ Ω). FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 7 simple X-valued functions and then approximating, relates the Lp(Rd; X)-norm and the X(Lp(Rd))-norm of such an f and will be used later. Lemma 2.1. Let X be a Banach function space, p ∈ [1,∞) and f ∈ Lp(Rd) ⊗ X. • If X is p-convex then kfkX(Lp(Rd)) ≤ CkfkLp(Rd;X), where C ≥ 0 is as in (2.3). • If X is p-concave then kfkLp(Rd;X) ≤ CkfkX(Lp(Rd)), where C ≥ 0 is as in (2.4) The proof of the following lemma is the same as in [44, Lemma 4] for simple X-valued functions, and the general case follows by approximation. Lemma 2.2. Let X and Y be Banach function spaces, P ∈ L(X, Y ) a positive operator, p ∈ [1,∞) and f ∈ Lp(Rd) ⊗ X. Then (cid:16)ZRdP (f (t))p dt(cid:17)1/p ≤ P(cid:18)(cid:16)ZRdf (t)p dt(cid:17)1/p(cid:19) . 2.4. γ-boundedness. Let X and Y be Banach spaces. A collection T ⊆ L(X, Y ) is γ-bounded if there exists a constant C ≥ 0 such that nXk=1 X(cid:17)1/2 nXk=1 2 Y(cid:17)1/2 (2.5) (cid:16)E(cid:13)(cid:13)(cid:13) γkTkxk(cid:13)(cid:13)(cid:13) ≤ C(cid:16)E(cid:13)(cid:13)(cid:13) 2 γkxk(cid:13)(cid:13)(cid:13) for all n ∈ N, T1, . . . , Tn ∈ T , x1, . . . , xn ∈ X and each Gaussian sequence (γk)n k=1. The smallest such C is the γ-bound of T and is denoted by γ(T ). By the Kahane- Khintchine inequalities, we may replace the L2-norm in (2.5) by an Lp-norm for each p ∈ [1,∞). Every γ-bounded collection is uniformly bounded with supremum bound less than or equal to the γ-bound, and the converse holds if and only if X has cotype 2 and Y has type 2 (see [4]). By the Kahane contraction principle, for each γ- bounded collection T ⊆ L(X, Y ) and each λ ∈ [0,∞), the closure in the strong operator topology of the family {zT z ∈ C,z ≤ λ, T ∈ T } ⊆ L(X, Y ) is γ- bounded with (2.6) γ(cid:16){zT z ∈ C,z ≤ λ, T ∈ T } SOT(cid:17) ≤ λγ(T ). By replacing the Gaussian random variables in (2.5) by Rademacher variables, one obtains the definition of an R-bounded collection T ⊆ L(X, Y ). Each R- bounded collection is γ-bounded. The notions of γ-boundedness and R-boundedness are equivalent if and only if X has finite cotype (see [39, Theorem 1.1]), but the minimal constant C in (2.5) may depend on whether one considers Gaussian or Rademacher variables. In this article we work with γ-boundedness instead of R- boundedness because in our results we will allow spaces which do not have finite cotype. 8 JAN ROZENDAAL AND MARK VERAAR 2.5. Bessel spaces. For details on Bessel spaces and related spaces see e.g. [2, 8, 29, 57]. space H s Lp(Rd; X). Then H s For X a Banach space, s ∈ R and p ∈ [1,∞] the inhomogeneous Bessel potential p (Rd; X) consists of all f ∈ S′(Rd; X) such that F−1((1 + ·)s/2bf (·) ) ∈ kfkHs p (Rd; X) is a Banach space endowed with the norm p(Rd;X) := kF−1((1 + ·2)s/2bf (·))kLp(Rd;X) (f ∈ H s p(Rd; X)), and S(Rd; X) ⊆ H s p(Rd; X) lies dense if p < ∞. In this article we will also deal with homogeneous Bessel spaces. To define these spaces we follow the approach of [57, Chapter 5] (see also [58]). Let X be a Banach space and define S(Rd; X) := {f ∈ S(Rd; X) Dαbf (0) = 0 for all α ∈ Nd 0}. S′(Rd; X) be the space of continuous linear mappings Endow S(Rd; X) with the subspace topology induced by S(Rd; X) and set S(Rd) := S(Rd; C). Let S(Rd) → X. Then each f ∈ S′(Rd; X) yields an f ↾ S(Rd) ∈ S′(Rd; X) by restriction, and f ↾ S(Rd) = g↾ S(Rd) if and only if supp(bf −bg) ⊆ {0}. Conversely, one can check that each f ∈ S′(Rd; X) extends to an element of S′(Rd; X) (see [50] for the tedious details in the vector-valued setting). Hence S′(Rd; X) = S′(Rd; X)/P(Rd; X) for P(Rd; X) := {f ∈ S′(Rd; X) supp(bf ) ⊆ {0}}. As in [24, Proposition 2.4.1] one can show that P(Rd; X) = P(Rd)⊗ X, where P(Rd) is the collection of polynomials on Rd. If F (Rd; X) ⊆ S′(Rd; X) is a linear subspace such that Φ = 0 if supp(bΦ ) ⊆ {0}, then we will identify F (Rd; X) with its image in S′(Rd; X). In particular, this is the case if F (Rd; X) = Lp(Rd; X) for some p ∈ [1,∞]. For s ∈ R and p ∈ [1,∞], the homogeneous Bessel potential space H s p(Rd; X) is the space of all f ∈ S′(Rd; X) such that F−1(·sbf (·)) ∈ Lp(Rd; X), where (ϕ ∈ S(Rd; X)). Then H s p(Rd; X) is a Banach space endowed with the norm hF−1(·sbf (·)), ϕi := hf,F−1(·sbϕ(·))i p(Rd;X) := kF−1(·sbf (·))kLp(Rd;X) p(Rd; X) lies dense if p < ∞. kfk Hs and S(Rd) ⊗ X ⊆ H s (f ∈ H s p (Rd; X)), 3. Fourier multipliers results In this section we introduce operator-valued Fourier multipliers acting on various vector-valued function spaces and discuss some of their properties. We start with some preliminaries and after that in Subsection 3.2 we prove a result that will allow us to transfer boundedness of multipliers on Rd to the torus Td. Then in Subsection 3.3 we present some first simple results under Fourier type conditions. We return to our main multiplier results for spaces with type, cotype, p-convexity and q-concavity in Subsections 3.4 and 3.5. FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 9 3.1. Definitions and basic properties. Fix d ∈ N, let X and Y be Banach spaces, and let m : Rd → L(X, Y ) be X-strongly measurable. We say that m is of moderate growth at infinity if there exist a constant α ∈ (0,∞) and a g ∈ L1(Rd) such that (1 + ξ)−αkm(ξ)kL(X,Y ) ≤ g(ξ) (ξ ∈ Rd). For such an m, let Tm : S(Rd; X) → S′(Rd; Y ) be given by Tm(f ) := F−1(m · bf ) (f ∈ S(Rd; X)). We call Tm the Fourier multiplier operator associated with m and we call m the symbol of Tm. Let p, q ∈ [1,∞]. We say that m is a bounded (Lp(Rd; X), Lq(Rd; Y ))-Fourier multiplier if there exists a constant C ∈ (0,∞) such that Tm(f ) ∈ Lq(Rd; Y ) and kTm(f )kLq(Rd;Y ) ≤ CkfkLp(Rd;X) for all f ∈ S(Rd; X). In the case 1 ≤ p < ∞, Tm extends uniquely to a bounded operator from Lp(Rd; X) to Lq(Rd; Y ) which will be denoted by fTm, and often just by Tm when there is no danger of confusion. If X = Y and p = q then we simply say that m is an Lp(Rd; X)-Fourier multiplier. We will also consider Fourier multipliers on homogeneous function spaces. Let X and Y be Banach spaces and let m : Rd\{0} → L(X, Y ) be X-strongly measurable. We say that m : Rd \ {0} → L(X, Y ) is of moderate growth at zero and infinity if there exist a constant α ∈ (0,∞) and a g ∈ L1(Rd) such that ξα(1 + ξ)−2αkm(ξ)kL(X,Y ) ≤ g(ξ) (ξ ∈ Rd). For such an m, let Tm : S(Rd; X) → S′(Rd; Y ) be given by Tm(f ) := F−1(m · bf ) (f ∈ S(Rd; X)), where Tm(f ) ∈ S′(Rd; Y ) is well-defined by definition of S(Rd; X). We use similar terminology as before to discuss the boundedness of Tm. Often we will simply write Tm = Tm, to simplify notation. In later sections we will use the following lemma about approximation of multi- pliers, which can be proved as in [24, Proposition 2.5.13]. Lemma 3.1. Let X and Y be Banach spaces and q ∈ [1,∞]. For each n ∈ N let mn : Rd → L(X, Y ) be X-strongly measurable, and let m : Rd → L(X, Y ) be such that m(ξ)x = limn→∞ mn(ξ)x for all x ∈ X and almost all ξ ∈ Rd. Suppose that there exist α > 0 and g ∈ L1(Rd) such that (1 + ξ)αkmn(ξ)kL(X,Y ) ≤ g(ξ) for all n ∈ N and ξ ∈ Rd. If f ∈ S(Rd; X) is such that Tmn (f ) ∈ Lq(Rd; Y ) for all n ∈ N, and if lim inf n→∞ kTmn(f )kLq(Rd;Y ) < ∞, then Tm(f ) ∈ Lq(Rd; Y ) with kTm(f )kLq(Rd;Y ) ≤ lim inf n→∞ kTmn(f )kLq(Rd;Y ). The same result holds for f ∈ S(Rd; X) if instead we assume that there exist an α > 0 and g ∈ L1(Rd) such that, for all n ∈ N and ξ ∈ Rd, ξ−α(1 + ξ)2αkmn(ξ)kL(X,Y ) ≤ g(ξ). 10 JAN ROZENDAAL AND MARK VERAAR The case of positive scalar-valued kernels plays a special role. An immediate consequence of [24, Proposition 4.5.10] is: Proposition 3.2 (Positive kernels). Let m : Rd \ {0} → C have moderate growth Tm : Lp(Rd) → Lq(Rd) is bounded for some at zero and infinity. Suppose that p, q ∈ [1,∞] and that F−1m ∈ S′(Rd) is positive. Then, for any Banach space X, the operator Tm ⊗ IX : Lp(Rd; X) → Lq(Rd; X) is bounded of norm kTm ⊗ IXkL(Lp(Rd;X),Lq(Rd;Y )) ≤ kTmkL(Lp(Rd),Lq(Rd)). The Hardy–Littlewood–Sobolev inequality on fractional integration is a typical p − 1 example where Proposition 3.2 can be applied. Example 3.3. Let X be a Banach space and 1 < p ≤ q < ∞. Let m(ξ) := ξ−s q = s for ξ ∈ Rd. Then Tm : Lp(Rd; X) → Lq(Rd; X) is bounded if and only if 1 d . In this case F−1m(·) = Cs · −d+s is positive and therefore the result follows from the scalar case (see [25, Theorem 6.1.3]) and Proposition 3.2. The same holds for q ≤ s the multiplier m(·) := (1 + · 2)−s/2 under the less restrictive condition 1 d . 3.2. Transference from Rd to Td. We will mainly consider Fourier multipliers on Rd. However, we want to present at least one transference result to obtain Fourier multiplier results for the torus Td := [0, 1]d. The transference technique differs slightly from the standard setting of de Leeuw's theorem where p = q (see [15, Theorem 4.5] and [29, Chapter 5]), due to the fact that kTmakL(Lp(Rd),Lq(Rd)) = a−d/rkTmkL(Lp(Rd),Lq(Rd)), where 1 q and ma(ξ) := m(aξ) for a > 0. Let ek : Td → C be given by ek(t) := e2πik·t for k ∈ Z and t ∈ Td. Proposition 3.4 (Transference). Let p, q, r ∈ (1,∞) be such that 1 q . Let m : Rd → L(X, Y ) be such that m(·)x ∈ L1 loc(Rd; Y ) for all x ∈ X. Fix a > 0 and let mkx := a−dR[0,a]d m(t + ka)x dt for k ∈ Zd. If Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded, then for all n ∈ N and (xk)k≤n in X, p − 1 p − 1 p − 1 r = 1 r = 1 ad/r(cid:13)(cid:13)(cid:13) Xk≤n ekmkxk(cid:13)(cid:13)(cid:13)Lq(Td;Y ) ≤ Cd,p,q′kTmk(cid:13)(cid:13)(cid:13) Xk≤n ekxk(cid:13)(cid:13)(cid:13)Lp(Td;X) This result seems to be new even in the scalar case X = Y = C. for some Cd,p,q′ ≥ 0. In particular, the Fourier multiplier operator with symbol (mk)k∈Zd is bounded from Lp(Td; X) to Lq(Td; Y ). Proof. Let P =Pk≤n ekxk. Since Lq′ (Td; Y ∗) is norming for Lq(Td; Y ) and since (Td; Y ∗), it suffices to the Y ∗-valued trigonometric polynomials are dense in Lq′ show that ekmkxk, QE(cid:12)(cid:12)(cid:12) ≤ Cd,p,q′kTmk kPkLp(Td;X)kQkLq′ (Td;Y ∗) ad/r(cid:12)(cid:12)(cid:12)D Xk≤n for Q : Td → Y ∗ an arbitrary Y ∗-valued trigonometric polynomial. Moreover, adding zero vectors xk or y∗k and enlarging n if necessary, we can assume that Q =Pk≤n e−ky∗k. To prove (3.1) observe that for E := Lmin(p,q′)(Rd) and f ∈ E ⊗ X, g ∈ E ⊗ Y ∗, (3.1) the boundedness of Tm is equivalent to (3.2) (cid:12)(cid:12)(cid:12)ZRdhm(ξ)bf (ξ),bg(ξ)i dξ(cid:12)(cid:12)(cid:12) ≤ kTmk kfkLp(Rd;X)kgkLq′ (Rd;Y ∗), bf (ξ) = a−d/p′ Xk≤n 1[0,a]d+ak(ξ)xk, for ξ ∈ Rd. By substitution we find kfkLp(Rd;X) =(cid:16)ZRd h(t)pkP (t)kp =(cid:16)Z[0,1]d H(t)pkP (t)kp X dt(cid:17)1/p X dt(cid:17)1/p 1[0,a]d+ak(ξ)y∗k bg(ξ) = a−d/q Xk≤n =(cid:16) Xj∈ZdZ[0,1]d+j h(t)pkP (t)kp X dt(cid:17)1/p ≤ Cd,pkPkLp(Td;X), FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 11 j=1 sin(πtj ) where we have used that hmbf ,bgi = hTmf, gi. Let h(t) := F−1(1[0,1]d )(t) = eiπ(t1+...+td)Qd Then f ∈ E ⊗ X, g ∈ E ⊗ Y ∗, and for t = (t1, . . . , tn) ∈ Rd, and g(t) := ad/qh(at)Q(−at). f (t) := ad/ph(at)P (at), πtj where we used the standard fact that H(t) =Pj∈Zd h(t + j)p ≤ Cd,p for t ∈ Rd, p ∈ (1,∞) and some Cd,p ≥ 0. Similarly, one checks that kgkLq′ (Rd;Y ∗) ≤ Cd,q′kQkLq′ (Td;Y ∗). Since the left-hand side of (3.2) equals the left-hand side of (3.1), the first statement follows from these estimates. The second statement follows from the first since the X-valued trigonometric (cid:3) polynomials are dense in Lp(Td; X). Remark 3.5. Any Fourier multiplier from Lp(Td; X) to Lq(Td; Y ) with 1 ≤ p ≤ q ≤ ∞, trivially yields a multiplier from Lu(Td; X) into Lv(Td; Y ) for all p ≤ u ≤ v ≤ q. Indeed, this follows from the embedding La(Td; X) ֒→ Lb(Td; X) for a ≥ b. In particular, any boundedness result from Lp(Td; X) to Lq(Td; Y ) implies boundedness from Lu(Td; X) into Lu(Td; Y ). As an application of Proposition 3.4 and Theorem 1.1 we obtain the following: Corollary 3.6. Let X be a Banach space with type p0 ∈ (1, 2] and Y a Banach space with cotype q0 ∈ [2,∞), and let p ∈ (1, p0), q ∈ (q0,∞). Let r ∈ (1,∞] be such that 1 q . If (mk)k∈Zd is a family of operators in L(X, Y ) and p − 1 r = 1 {(kd/r + 1)mk k ∈ Zd} ⊆ L(X, Y ) is γ-bounded, then the Fourier multiplier operator with symbol (mk)k∈Z is bounded from Lp(Td; X) to Lq(Td; Y ). Moreover, if p0 = 2 (or q0 = 2), then one can also take p = 2 (or q = 2). Proof. Let m(ξ) := Pk∈Zd 1[0,1]d(ξ − k)mk for ξ ∈ Rd. Then for k ∈ Zd and ξ ∈ [0, 1]d + k, we have m(ξ) = mk and ξd/r ≤ (k + √d)d/r ≤ Cd,r(kd/r + 1). Therefore, Kahane's contraction principle yields d γ({ξ r m(ξ) ξ ∈ Rd}) ≤ Cd,rγ({(k + 1) r mk k ∈ Zd}), d which is assumed to be finite. By Theorem 1.1, Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. Since mk = R[0,1]d m(t + k) dt for k ∈ Zd, Proposition 3.4 yields the required result. (cid:3) 12 JAN ROZENDAAL AND MARK VERAAR As an application we show how Corollary 3.6 can be used in the study of Schur multipliers. For p ∈ [1,∞) let C p denote the Schatten p-class over a Hilbert space H. For a detailed discussion on these spaces we refer to [17] and [29]. Let (ej)j∈Z be a countable spectral resolution of H. That is, (1) for all j ∈ Z, ej is an orthogonal projection in H; (2) for all j, k ∈ Z, ejek = 0 if j 6= k; (3) for all h ∈ H, Pj∈Z ejh = h. Using the technique of [48, Theorem 4] we deduce the following result from Corollary 3.6. A similar result holds for more general noncommutative Lp-spaces with a similar proof. a − 1 Corollary 3.7. Let a ∈ (1,∞) \ {2} and let r ∈ [1,∞) be such that 1 2. Let m : Z → C be such that Cm := supj∈Z(1 + j1/r)mj < ∞, let f : Z → Z and write mf m,f on C a, given by j,k := mf (j)−f (k). Then the Schur multiplier operator M e r < 1 j,kejvek = lim mf j,kejvek n→∞ Xj,k≤n (3.3) M e mf m,f v := Xj,k∈Z kM e m,fkL(C a) ≤ Ca,rCm for v ∈ C a, is well-defined and satisfies (3.4) for some Ca,r ≥ 0 independent of m. Proof. By duality it suffices to consider a ∈ (1, 2), and by an approximation argu- ment it suffices to consider finite rank operators v ∈ C a. Let p ∈ (1, a) be such that p − 1 r . Since C a has type a and cotype 2 (see [30]) it follows from Theorem 3.6 that the Fourier multiplier Tm associated with (mn)n∈Z is bounded from La(T; C a) to L2(T; C a) with 2 = 1 1 (3.5) kTmkL(La(T;C a),L2(T;C a)) ≤ Cp,aCm. As in the proof of [48, Theorem 4] one sees that = kTm((vn)n∈Z)(t)kC a , kM e m,f vkC a =(cid:13)(cid:13)(cid:13)Xn∈Z mne2πintvn(cid:13)(cid:13)(cid:13)C a kvkC a =(cid:13)(cid:13)(cid:13)Xn∈Z ≤ kTmkL(Lp(T;C a),Lq(T;C a))(cid:13)(cid:13)(cid:13)Xn∈Z where vn :=Pj,k∈Z,f (j)−f (k)=n ejvek for n ∈ Z. Similarly, e2πintvn(cid:13)(cid:13)(cid:13)C a e2πintvn(cid:13)(cid:13)(cid:13)Lp(0,1;C a) Taking Lq and Lp norms over t ∈ [0, 1] in the above identities yields kM e m,f vkC a = kTm(vn)n∈ZkLq(0,1;C a) . where we applied (3.5) in the final step. a − 1 Problem 3.8. Can we take 1 r = 1 2 in Corollary 3.7? = Cp,aCmkvkC a, (cid:3) If the answer to the question in Problem 3.8 is negative, then the limitations of Theorem 1.1 and Corollary 3.6 are natural. Moreover, from the proof of the latter (R; C a) → (see Theorem 3.18 below) it would then follow that the embedding H a− 1 a 1 2 FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 13 γ(R; C a) does not hold for a ∈ (1, 2). Here γ(R; C a) is the C a-valued γ-space used in the proof of Theorem 3.18. 3.3. Fourier type assumptions. Before turning to more advanced multiplier the- orems, we start with the case where we use the Fourier type of the Banach spaces to derive an analogue of the basic estimate kTmkL(L2(Rd)) ≤ kmk∞. Proposition 3.9. Let X be a Banach space with Fourier type p ∈ [1, 2] and Y a Banach space with Fourier cotype q ∈ [2,∞], and let r ∈ [1,∞] be such that r = 1 q . Let m : Rd → L(X, Y ) be an X-strongly measurable map such that km(·)kL(X,Y ) ∈ Lr(Rd). Then Tm extends uniquely to a bounded map from Lp(Rd; X) into Lq(Rd; Y ) with p − 1 1 kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Fp,X,d Fq′,Y,d(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr(Rd) . In Proposition 3.15 we show that this multiplier result characterizes the Fourier type p of X for specific choices of Y , and the Fourier cotype q of Y for specific choices of X. Proof. Let f ∈ S(Rd; X). By Holder's inequality, kmbfkLq′ (Rd;Y ) ≤(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr(Rd) kbfkLp′ (Rd;X) ≤ Fp,X,d(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr(Rd) kfkLp(Rd;X). Since kF−1(g)kLq(Rd;Y ) = kF (g)kLq(Rd;Y ) for g ∈ Lq′ (Rd; Y ), it follows that kTm(f )kLq(Rd;Y ) ≤ Fq′,Y,dkmbfkLq′ (Rd;Y ) ≤ Fp,X,d Fq′,Y,d(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr(Rd) kfkLp(Rd;X), which concludes the proof. (cid:3) Remark 3.10. It follows from Young's inequality (see [24, Exercise 4.5.4] or [3, Proposition 1.3.5]) that Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded with kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ kF−1mkLr′ (Rd;L(X,Y )) (3.6) for all X and Y , 1 ≤ p ≤ q ≤ ∞ and r ∈ [1,∞] such that 1 q , and all X-measurable m : Rd → L(X, Y ) of moderate growth at infinity for which F−1m ∈ Lr′ In certain cases (3.6) is stronger than the result in Proposition 3.9. For instance, if r ∈ [1, 2] and L(X, Y ) has Fourier type r (for r > 1 this implies that either X or Y is finite-dimensional), then (Rd;L(X, Y )). p − 1 r = 1 kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ kF−1mkLr′ (Rd;L(X,Y )) ≤ CkmkLr(Rd;L(X,Y )) for some constant C ≥ 0. Therefore we recover the conclusion of Proposition 3.9 from Young's inequality in a very special case. Remark 3.11. Proposition 3.9 (and Theorem 3.12 below) can also be formulated for general abelian locally compact groups G, not just for Rd. In that case one (bG; X) should assume that the Fourier transform is bounded from Lp(G; X) to Lp′ (bG; Y ) for p ∈ [1, 2] and that the inverse Fourier transform is bounded from Lq′ to Lq(G; Y ) for q ∈ [2,∞]. Here bG is the dual group of G. Then one works with symbols m : bG → L(X, Y ) which are X-strongly measurable and such that 14 JAN ROZENDAAL AND MARK VERAAR [ξ 7→ km(ξ)kL(X,Y )] ∈ Lr(bG), where 1 3.9, one then obtains a constant C ≥ 0 independent of m such that p − 1 r = 1 q . In the same way as in Proposition kTmk ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr( bG) . For G = Td such results can also be deduced from the Rd-case by applying the transference of Proposition 3.4. In the scalar setting we noted in (1.1) that the conclusion of Proposition 3.9 In certain cases we can prove holds under the weaker condition m ∈ Lr,∞(Rd). such a result in the vector-valued setting. Theorem 3.12. Let X be a Banach space with Fourier type p0 ∈ (1, 2] and Y a Banach space with Fourier cotype q0 ∈ [2,∞), and let p ∈ (1, p0) and q ∈ (q0,∞). Let r ∈ [1,∞] be such that 1 q . Let m : Rd → L(X, Y ) be an X-strongly measurable map such that [ξ 7→ km(ξ)kL(X,Y )] ∈ Lr,∞(Rd). Then Tm extends uniquely to a bounded map from Lp(Rd; X) into Lq(Rd; Y ) with p − 1 r = 1 kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr,∞(Rd) , where C ≥ 0 is independent of m. Proof. Observe that by real interpolation (see [56, 1.18.6] and [37, (2.33)]) we obtain F : Lv′,∞(Rd; Y ) → Lv,∞(Rd; Y ) for all v ∈ (q0,∞). Let p1, p2, q1, q2 ∈ (1,∞) be such that 1 p − ε, + ε, 1 p2 1 p1 1 p = = 1 q1 = 1 q + ε, 1 q2 = 1 q − ε for ε > 0 so small that p2 < p0 and q1 > q0. Note that 1 pj − 1 qj = 1 p − 1 q = 1 r . Let f ∈ S(Rd; X). By Holder's inequality (see [24, Exercise 1.4.19] or [45, Theorem 3.5]), for j = 1, 2, kmbfkL q′ j ,∞ (Rd;Y ) ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr,∞(Rd) kbfkL ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr,∞(Rd) kfkLpj (Rd;X) j (Rd;X) p′ for C ≥ 0 independent of m and f , where we used the Fourier type pj of X and . It follows from the first observation and the estimate above that k · kp′ j ,∞ ≤ k · kp′ j kTm(f )kLqj ,∞(Rd;Y ) ≤ CkmbfkL q′ j ,∞ (Rd;Y ) ≤ C(cid:13)(cid:13)km(·)kL(X,Y )(cid:13)(cid:13)Lr,∞(Rd) kfkLpj (Rd;X). Hence Tm : Lpj (Rd; X) → Lqj ,∞(Rd; Y ) is bounded for j ∈ {1, 2}. By real inter- polation (see [56, Theorem 1.18.6.2]) we find that Tm : Lp(Rd; X) → Lq,p(Rd; Y ) and the required result follows from Lq,p(Rd; Y ) ֒→ Lq(Rd; Y ) (see [24, Proposition 1.4.10]). (cid:3) The above result provides an analogue of [28, Theorem 1.12]. In general, we do not know the "right" geometric conditions under which such a result holds. We formulate the latter as an open problem. FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 15 p − 1 Problem 3.13. Let 1 < p ≤ 2 ≤ q < ∞ and let r ∈ [1,∞] be such that 1 q . Classify those Banach spaces X and Y for which Tm ∈ L(Lp(Rd; X), Lq(Rd; Y )) for all X-strongly measurable maps m : Rd → L(X, Y ) such that km(·)kL(X,Y ) ∈ Lr,∞(Rd). r = 1 A similar question can be asked for the case where X = Y and m is scalar-valued. We will now show that the Fourier multiplier result in Proposition 3.9 charac- terizes the Fourier type of the underlying Banach spaces. To this end we need the following lemma. Lemma 3.14. Let X and Y be Banach spaces. Let p ∈ [1, 2], q ∈ [2,∞] and r ∈ (1,∞] be such that 1 q . Assume that for all m ∈ Lr(Rd;L(X, Y )) the operator Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. Then there is a constant C ≥ 0 such that for all f ∈ S(Rd; X) and g ∈ S(Rd; Y ∗) (3.7) (cid:13)(cid:13)kbf (·)kXkbg(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) ≤ CkfkLp(Rd;X)kgkLq′ (Rd;Y ∗). Proof. By the closed graph theorem there exists a constant C ≥ 0 such that p − 1 r = 1 hTmf, gi ≤ CkmkLr(Rd;L(X,Y ))kfkLp(Rd;X)kgkLq′ (Rd;Y ∗) 1 1 r + 1 2 , εkbgk−1 p′ + 1 q ) and kη−bgkq ≤ min(ε (Rd; Y ∗) and m ∈ Lr(Rd;L(X, Y )). It follows that, hmbf ,bgi = hTmf, gi ≤ CkmkLr(Rd;L(X,Y )). for all f ∈ Lp(Rd; X), g ∈ Lq′ for all f ∈ S(Rd; X) with kfkp ≤ 1 and g ∈ S(Rd; Y ∗) with kgkq′ ≤ 1, (3.8) It suffices to show (3.7) for fixed f ∈ S(Rd; X) with kfkp = 1 and g ∈ S(Rd; Y ∗) with kgkq′ = 1. Let ε ∈ (0, 1) and choose simple functions ζ : Rd → X and η : Rd → Y ∗ such that kζ−bfkp′ ≤ min(ε 2 , εkbfk−1 p′ ). Then, by Holder's inequality with 1 hmζ, ηi ≤ hm(ζ − bf ), η −bgi + hm(ζ − bf ),bgi + hmbf , η −bgi + hmbf ,bgi ≤ kmkr(cid:16)kζ − bfkp′kη −bgkq + kζ − bfkp′kbgkq + kbfkp′kη −bgkq + C(cid:17) ≤ kmkr(3ε + C) k=1 1Ak xk and η =Pn for all m ∈ Lr(Rd;L(X, Y )). By considering a common refinement, we may suppose that ζ =Pn k=1 1Ak y∗k for n ∈ N, x1, . . . , xn ∈ X, y∗1, . . . , y∗n ∈ Y ∗ and A1, . . . , An ⊆ Rd disjoint and of finite measure Ak. For 1 ≤ k ≤ n let x∗k ∈ X∗ and yk ∈ Y of norm one be such that hxk, x∗ki = kxkk and hyk, y∗ki ≥ (1 − ε)ky∗kk. Let m : Rd → L(X, Y ) be given by ck1Ak (ξ)hx, x∗kiyk q = 1 and by (3.8), it follows that (ξ ∈ Rd, x ∈ X), m(ξ)x := (3.9) nXk=1 where c1, . . . , cn ∈ R. Then (3.9) implies (1 − ε) nXk=1 ckAkkxkk ky∗kk ≤ (C + 3ε)(cid:16) nXk=1 By taking the supremum over all ck's with Pn (1 − ε)(cid:13)(cid:13)kζ(·)kXkη(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) = (1 − ε)(cid:16) nXk=1 r . ckrAk(cid:17) 1 k=1 ckrAk ≤ 1 we find ky∗kkr′(cid:17) 1 Akkxkkr′ r′ ≤ (C + 3ε). 16 JAN ROZENDAAL AND MARK VERAAR Therefore, using this estimate, the reverse triangle inequality and Holder's inequal- ity (with 1 r′ = 1 p′ + 1 q ), we obtain (cid:13)(cid:13)kbf (·)kXkbg(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) ≤(cid:13)(cid:13)kbf (·)kXkbg(·)kY ∗ − kζ(·)kXkbg(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) +(cid:13)(cid:13)kζ(·)kXkbg(·)kY ∗ − kζ(·)kXkη(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) +(cid:13)(cid:13)kζ(·)kXkη(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) ≤(cid:13)(cid:13)kbf (·) − ζ(·)kXkbg(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) +(cid:13)(cid:13)kζ(·)kXkη(·) −bg(·)kY ∗(cid:13)(cid:13)Lr′ (Rd) + ≤ kbf − ζkp′kbgkq + kζkp′kη −bgkq + ≤ ε + (kbf − ζkp′ + kbfkp′)kη −bgkq + Letting ǫ tend to zero yields (3.7) for kfkp = 1 = kgkq′, as was to be shown. C + 3ε 1 − ε C + 3ε 1 − ε ≤ 3ε + C + 3ε 1 − ε . C + 3ε 1 − ε (cid:3) Now we are ready to show that, by letting Y vary, the Fourier multiplier result in Proposition 3.9 characterizes the Fourier type of X, and vice versa. r = 1 p − 1 Proposition 3.15. Let X and Y be Banach spaces. Let 1 q with p ∈ [1, 2], q ∈ [2,∞) and r ∈ (1,∞]. Assume that for all m ∈ Lr(Rd;L(X, Y )) the operator Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. (1) If Y = C and q = 2, then X has Fourier type p. (2) If X = C and p = 2, then Y has Fourier type q′. (3) If Y = X∗ and q = p′, then X has Fourier type p. Proof. By Lemma 3.14, (3.7) holds for some C ≥ 0. Therefore in case (1) we obtain, for fixed f ∈ S(Rd; X) and for all ϕ ∈ S(Rd), (cid:13)(cid:13)kbf (·)kXϕ(·)(cid:13)(cid:13)Lr′ (Rd) ≤ CkfkLp(Rd;X)kϕkL2(Rd), where we used the fact that F : L2(Rd) → L2(Rd) is an isometry. Taking the supremum over all kϕkL2(Rd) ≤ 1 we see that and the result follows. (cid:3) Remark 3.16. An alternative proof of Proposition 3.15 can be given using the transference of Proposition 3.4. However, this yields worse bounds and it seems that the analogue in the type-cotype setting requires the same technique as in Proposition 3.15. The estimate which can be proved under the assumption of Lemma 3.14 is as follows. There is a constant C ≥ 0 such that for all (xk)k≤n in X and (y∗k)k≤n in Y ∗, Y(cid:17) 1 Xky∗kkr′ kxkkr′ r′ . ekxk(cid:13)(cid:13)(cid:13)Lp(Td;X)(cid:13)(cid:13)(cid:13) Xk≤n eky∗k(cid:13)(cid:13)(cid:13)Lq′ (Td;Y ∗) ≤ C(cid:13)(cid:13)(cid:13) Xk≤n (cid:16) Xk≤n and hence X has Fourier type p. In case (2) we deduce in the same way that Y ∗ has Fourier type q′ and thus also that Y has Fourier type q′, by duality. Finally, for (3) note that 1 p′ . Thus, taking f = g ∈ S(Rd; X) in (3.7) yields kbfkLp′ (Rd;X) ≤ CkfkLp(Rd;X), r′ = 2 Lp′ (Rd;X) ≤ Ckfk2 kbfk2 Lp(Rd;X), FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 17 We end this section with a simple example which shows that the geometric limitation in Theorem 3.9 is also natural in the case X = Y = ℓu. We will come back to this in Example 3.30, where type and cotype will be used to derive different results. Example 3.17. Let p ∈ (1, 2], and for q ∈ [2,∞) let r ∈ (1,∞] be such that r = 1 q . Let u ∈ [1,∞) and let X := ℓu. Let (ej)j∈N0 ⊆ X be the standard basis of X, and for k ∈ N let Sk ∈ L(X) be such that Sk(ej) := ej+k for j ∈ N0. Let m : R → L(ℓu) be given by m(ξ) := P∞k=1 ck1(k−1,k](ξ)Sk for ξ ∈ R, where ck = k− 1 p − 1 1 r log(k + 1)−2 for k ∈ N. Observe that ∞Xk=1 L(X) dξ = ZR km(ξ)kr cr k < ∞, u 1 with the obvious modification for r = ∞. If u ∈ [p, p′], then X has Fourier type p and Fourier cotype q = p′. Thus by Proposition 3.9, in this case Tm : Lp(R; X) → Lq(R; X) is bounded. We show that this result is sharp in the sense that for u /∈ [p, p′] the conclusion is false. This shows that Proposition 3.9 is optimal in the exponent of the Fourier type of the space for X = Y = ℓu. Let q ∈ [2,∞) and assume that Tm ∈ L(Lp(R; X), Lq(R; X)). Let, for k ∈ N, k=n+1 ϕke0. Then 2nXk=n+1 =(cid:16) 2nXk=n+1 kTm(f )(t)kX =(cid:13)(cid:13)(cid:13) ϕk : R → C be such that cϕk = 1(k−1,k] and let, for n ∈ N, f := P2n ckuϕk(t)u(cid:17) 1 ckϕk(t)ek(cid:13)(cid:13)(cid:13)ℓu (cid:12)(cid:12) for all t ∈ R and k ∈ N0, for each t ∈ R. Since ϕk(t) =(cid:12)(cid:12) sin(πt) for some C1 ∈ (0,∞). On the other hand, kfkLp(R;X) =(cid:13)(cid:13)P2n (cid:12)(cid:12)P2n k=n+1 ϕk(t)(cid:12)(cid:12) = (cid:12)(cid:12) sin(πnt) there exists a constant C2 ∈ (0,∞) such that kfkLp(R;ℓu) = C2n1− 1 r log(n)−2 ≤ kTmkL(Lp(R;X),Lq(R;X))C2n1− 1 (cid:12)(cid:12) for all t ∈ R, since P2n uc2nkϕ1kLq(R) ≥ C1n kTm(f )kLq(R;X) ≥ n r log(n)−2. u− 1 u − 1 C1n p . πt 1 k=n+1 ϕk(cid:13)(cid:13)Lp(R). Now, k=n+1 bϕk = 1(n,2n]. Therefore p . It follows that πt 1 u ≤ 1 − 1 Letting n → ∞ we deduce that 1 q = p′, we obtain u ≥ p. By a duality argument one sees that also u ≤ p′. 3.4. Type and cotype assumptions. In Proposition 3.9 and Theorem 3.12 we obtained Fourier multiplier results under Fourier type assumptions on the spaces X and Y . In this section we will present multiplier results under the less restrictive geometric assumptions of type p and cotype q on the underlying spaces X and Y . q′ . Thus, in the special case r = 1 p + 1 First we prove Theorem 1.1 from the Introduction. Theorem 3.18. Let X be a Banach space with type p0 ∈ (1, 2] and Y a Banach space with cotype q0 ∈ [2,∞), and let p ∈ (1, p0) and q ∈ (q0,∞), r ∈ (1,∞) be r = 1 such that 1 q . Let m : Rd \ {0} → L(X, Y ) be an X-strongly measurable r m(ξ) ξ ∈ Rd \{0}} ⊆ L(X, Y ) is γ-bounded. Then Tm extends map such that {ξ uniquely to a bounded map fTm ∈ L(Lp(Rd; X), Lq(Rd; Y )) with p − 1 d d r m(ξ) ξ ∈ Rd \ {0}}), kfTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Cγ({ξ 18 JAN ROZENDAAL AND MARK VERAAR where C ≥ 0 is independent of m. Moreover, if p0 = 2 (or q0 = 2), then one can also take p = 2 (resp. q = 2). It is unknown whether Theorem 3.18 holds with p = p0 and q = q0 (see Problem 3.19 below). Proof. We will prove the result under the condition: (3.10) H d p− d p 2 (Rd; X) ֒→ γ(Rd; X) and γ(Rd; Y ) ֒→ H d q − d q 2 (Rd; Y ). Here γ(Rd; X) is the X-valued γ-space (for more on these spaces see [59]). Note that the assumptions imply (3.10). Indeed, this follows from the homogeneous versions of [60, Proposition 3.5] and of [33, Theorem 1.1] (proved in exactly the same way, here we use the assumption that X has type p0 and p < p0). Moreover, if p0 = 2, then H 0 2 (Rd; X) = L2(Rd; X) ֒→ γ(Rd; X) (see [59, Theorem 11.6]), hence in this case one can in fact take p = 2. The embedding for Y follows in a similar way. r m(ξ)m1(ξ) for ξ ∈ Rd. Let f ∈ S(Rd; X). p and m2(ξ) := ξ Let m1(ξ) := ξ 2 − d d d It follows from (3.10) that kTm(f )kLq(Rd;Y ) = kTm2(f )k H d q q − d 2 (Rd;Y ) d d ≤ CkTm2(f )kγ(Rd;Y ) ≤ C1km2bfkγ(Rd;Y ) ≤ Cγ({ξ ≤ Cγ({ξ ≤ Cγ({ξ = Cγ({ξ r m(ξ) ξ ∈ Rd \ {0}})km1bfkγ(Rd;X) r m(ξ) ξ ∈ Rd \ {0}})kTm1fkγ(Rd;X) r m(ξ) ξ ∈ Rd \ {0}})kTm1fk H r m(ξ) ξ ∈ Rd \ {0}})kfkLp(Rd;X), − d 2 d p p d d (Rd;X) where we have used kfkγ(Rd;X) = kbfkγ(Rd;X) (see [30]), the γ-multiplier Theorem (see [35, Proposition 4.11] and [59, Theorem 5.2]) and the fact that γ(Rd; X) = γ∞(Rd; Y ) because Y does not contain a copy of c0 (see [59, Theorem 4.3]). Since S(Rd; X) ⊆ Lp(Rd; X) is dense, this concludes the proof. (cid:3) In Theorem 3.21 we provide conditions under which one can take p = p0 and q = q0. The general case we state as an open problem: Problem 3.19. Let 1 ≤ p ≤ 2 ≤ q ≤ ∞ and r ∈ (1,∞] be such that 1 p − 1 q . Classify those Banach spaces X and Y for which Tm ∈ L(Lp(Rd; X), Lq(Rd; Y )) for all X-strongly measurable maps m : Rd → L(X, Y ) such that {ξd/rm(ξ) : ξ ∈ Rd \ {0}} is γ-bounded. r = 1 The same problem can be formulated in case m is scalar-valued, in which case the γ-boundedness reduces to uniform boundedness. Remark 3.20. Assume X and Y have property (α) as introduced in [47]. (This implies that X has finite cotype, and if X and Y are Banach lattices then property (α) is in fact equivalent to finite cotype.) In the multiplier theorems in this paper where γ-boundedness is an assumption, one can deduce a certain γ-boundedness result for the Fourier multiplier operators as well. Indeed, assume for example the conditions of Theorem 3.18. Let {mj : Rd \ {0} → L(X, Y ) j ∈ J } be a set of X- strongly measurable mappings for which there exists a constant C ≥ 0 such that for FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 19 d r mj(ξ) ξ ∈ Rd} ⊆ L(X, Y ) is γ-bounded by C. Note that, since each j ∈ J , {ξ X and Y have finite cotype, γ-boundedness and R-boundedness are equivalent. Now we claim that {gTmj j ∈ J } ⊆ L(Lp(Rd; X), Lq(Rd; Y )) is γ-bounded as well. To prove this claim one can use the method of [22, Theorem 3.2]. Indeed, using their notation, it follows from the Kahane-Khintchine inequalities that Rad(X) has the same type as X and Rad(Y ) has the same cotype as Y . Therefore, given j1, . . . , jn ∈ J and the corresponding mj1 , . . . , mjn , one can apply Theorem 3.18 to the multiplier M : Rd \ {0} → L(Rad(X), Rad(Y )) given as the diagonal operator with diagonal (mj1 , . . . , mjn ). In order to check the γ-boundedness one now applies property (α) as in [22, Estimate (3.2)]. 3.5. Convexity, concavity and Lp-Lq results in lattices. In this section we will prove certain sharp results in p-convex and q-concave Banach lattices. First of all, from the proof of Theorem 3.18 we obtain the following result with the sharp exponents p and q. Theorem 3.21. Let p ∈ [1, 2], q ∈ [2,∞), and let r ∈ [1,∞] be such that 1 p − 1 q . Let X be a complemented subspace of a p-convex Banach lattice with finite cotype and Y a Banach space that is continuously embedded in a q-concave Banach lattice. Let m : Rd → L(X, Y ) be an X-strongly measurable map such that {ξ r m(ξ) ξ ∈ Rd \ {0}} ⊆ L(X, Y ) is γ-bounded. Then Tm extends uniquely to a bounded map fTm ∈ L(Lp(Rd; X), Lq(Rd; Y )) with r = 1 (3.11) d d where C is a constant depending on X, Y , p, q and d. kfTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Cγ({ξ r m(ξ) ξ ∈ Rd \ {0}}), Proof. In the case where X is a p-convex and Y is a q-concave Banach lattice, the embeddings in (3.10) can be proved in the same way as in [60, Theorem 3.9], where the inhomogeneous case was considered. Therefore, the result in this case follows from the proof of Theorem 3.18. Now let X0 be a p-convex Banach lattice with finite cotype such that X ⊆ X0, let P ∈ L(X0) be a projection with range X and let Y0 be a q-concave Banach lattice with a continuous embedding ι : Y ֒→ Y0. Let m0 : Rd → L(X0, Y0) be given by m0(ξ) := ι ◦ m(ξ) ◦ P ∈ L(X0, Y0) for ξ ∈ Rd. It is easily checked that {m0(ξ) ξ ∈ Rd} ⊆ L(X0, Y0) is γ-bounded, with (3.12) γ({m0(ξ) ξ ∈ Rd \ {0}}) ≤ kιkL(Y,Y0)kPkL(X0)γ({ξ r m(ξ) ξ ∈ Rd \ {0}}). d (3.13) As we have shown above, there exists a constant C ∈ (0,∞) that depends only on X0, Y0, p, q and d such that Tm0 extends uniquely to a bounded operator gTm0 ∈ L(Lp(Rd; X0), L(Rd; Y0)) with kTm0kL(Lp(Rd;X0),Lq(Rd;Y0)) ≤ Cγ({m0(ξ) ξ ∈ Rd}). Since Tm = gTm0↾S(R;X), the result follows from (3.12) and (3.13). Remark 3.22. Note from (3.12) and (3.13) that the constant C in (3.11) depends on X and Y as C = kPkL(X0) kιkL(Y,Y0) C1, where P ∈ L(X0) is a projection with range X on a p-convex Banach lattice X0 with finite cotype, ι ∈ L(Y, Y0) is a continuous embedding of Y in a q-concave Banach lattice Y0 and C1 is a constant that depends only on X0, Y0, p, q and d. (cid:3) 20 JAN ROZENDAAL AND MARK VERAAR Remark 3.23. By using Theorems 3.18 and 3.21 and by multiplying in the Fourier domain by appropriate powers of ξ, versions of these theorems for multipliers from H α p (Rd; X) to H β q (Rd; Y ) can be derived. Similar results can be derived for the inhomogeneous spaces as well. So far, in all our results about (Lp, Lq)-multipliers the indices p and q have been restricted to the range p ≤ 2 ≤ q, which is necessary when considering general mul- tipliers (see (1.1)). However, we have also seen in Example 3.3 that for the scalar multiplier m(ξ) = ξ−s such a restriction is not necessary, as follows from Proposi- tion 3.2 since the kernel associated with m is positive. We now show that also for operator-valued multipliers with positive kernels on p-convex and q-concave Banach lattices, the restriction p ≤ 2 ≤ q is not necessary and moreover γ-boundedness can be avoided. First we state the result for multipliers between Bessel spaces. Theorem 3.24. Let p, q ∈ [1,∞) with p ≤ q, and let r ∈ (1,∞] be such that 1 r = p− 1 q . Let X be a p-convex Banach lattice with finite cotype, and let Y be a q-concave Banach lattice. Suppose that K : Rd → L(X, Y ) is such that K(·)x ∈ L1(Rd; Y ) for all x ∈ X, K(s) is a positive operator for all s ∈ Rd, and m : Rd → L(X, Y ) is such that F (Kx) = mx for all x ∈ X. Then Tm ∈ L( H d/r (Rd; X), Lq(Rd; Y )) and p 1 (3.14) kTmkL( Hd/r p (Rd;X),Lq(Rd;Y )) ≤ Ckm(0)kL(X,Y ) ≤ C sup ξ∈Rd\{0} km(ξ)kL(X,Y ) for some C ≥ 0 independent of K. By further approximation arguments one can often avoid the assumptions that K(·)x ∈ L1(Rd; Y ) for all x ∈ X. It follows from [51] that the bound in Theorem 3.24 is optimal in a certain sense. Proof. The second estimate in (3.14) follows from the continuity of mx = F (Kx). Since S(Rd) ⊗ X is dense in H d/r(Rd; X), for the first estimate in (3.14) it suffices to fix an f ∈ S(Rd) ⊗ X and to show that Tm(f ) ∈ Lq(Rd; X) with (3.15) kTm(f )kLq(Rd;Y ) ≤ Ckm(0)k kfk Hd/r (Rd;X). p Since X has finite cotype, it does not contain a copy of c0. Hence, by [41, Theorem 1.a.5 and Proposition 1.a.7], X is order continuous. Moreover, the range of f is contained in a separable subspace X0 of X. By [41, Proposition 1.a.9], X0 has a weak order unit. Now [41, Theorem 1.b.14] implies that X0 is order isometric to a Banach function space. Similarly, Y is order continuous, and the range of Tm(f ) is contained in a separable subspace Y0 which is order isometric to a Banach function space. So henceforth we may assume without loss of generality that X and Y are Banach function spaces. It follows by approximation from Lemma 2.1 that kK ∗ fkLq(Rd;Y ) ≤ C1kK ∗ fkY (Lq(Rd)) = C(cid:13)(cid:13)(cid:13)(cid:16)ZRdK ∗ f (t)q dt(cid:17)1/q(cid:13)(cid:13)(cid:13)Y q = C(cid:13)(cid:13)(cid:13)(cid:16)ZRd(cid:12)(cid:12)(cid:12)ZRd K(s)f (t − s) ds(cid:12)(cid:12)(cid:12) ≤ C(cid:13)(cid:13)(cid:13)ZRd(cid:16)ZRdK(s)f (t − s)q dt(cid:17)1/q dt(cid:17)1/q(cid:13)(cid:13)(cid:13)Y ds(cid:13)(cid:13)(cid:13)Y FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 21 for some constant C ≥ 0, where we used Minkowski's integral inequality in the final step. Lemma 2.2, applied to the positive operator K(s) ∈ L(X, Y ) and the function f (· − s) ∈ Lp(Rd) ⊗ X for each s ∈ Rd, yields ZRd(cid:16)ZRdK(s)f (t − s)q dt(cid:17)1/q ds ≤ZRd =ZRd K(s)(cid:16)ZRdf (t − s)q dt(cid:17)1/q K(s)(cid:16)ZRdf (t)q dt(cid:17)1/q ds = m(0)x0, ds ∈ X. Therefore, ds(cid:13)(cid:13)(cid:13)Y ≤ km(0)k(cid:13)(cid:13)(cid:13)(cid:16)ZRdf (t)q dt(cid:17)1/q(cid:13)(cid:13)(cid:13)X . (Rd) ֒→ Lq(Rd) yields = kkf (·)kLq(Rd)kX ≤ Ckkf (·)k Hd/r p (Rd)kX. where x0 :=(cid:16)RRdf (t)q dt(cid:17)1/q The Sobolev embedding H d/r (cid:13)(cid:13)(cid:13)ZRd(cid:16)ZRdK(s)f (t − s)q dt(cid:17)1/q (cid:13)(cid:13)(cid:13)(cid:16)ZRdf (t)q dt(cid:17)1/q(cid:13)(cid:13)(cid:13)X p Finally, Lemma 2.1 yields that, for n(ξ) := ξd/rIX ∈ L(X), kkf (·)k Hd/r Combining all these estimates yields (3.15) and concludes the proof. (Rd)kX = kkTn(f )(·)kLp(Rd)kX ≤ CkTn(f )kLp(Rd;X) = Ckfk Hd/r p p (Rd;X). (cid:3) 1 In terms of Lp-Lq-multipliers we obtain the following result. Note that below we require that the kernel associated with the multiplicative perturbation ξd/rm(ξ) of m is positive, unlike in Proposition 3.2 where this positivity was required of the kernel associated with m. Corollary 3.25. Let p, q ∈ [1,∞) with p ≤ q, and let r ∈ (1,∞] be such that 1 r = p− 1 q . Let X be a p-convex Banach lattice with finite cotype, and let Y be a q-concave Banach lattice. Suppose that K : Rd → L(X, Y ) is such that K(·)x ∈ L1(Rd; Y ) for all x ∈ X, K(s) is a positive operator for all s ∈ Rd, and m : Rd \ {0} → L(X, Y ) is such that F (Kx)(·) = ·d/rm(·)x for all x ∈ X. Then Tm extends uniquely to a bounded map eTm ∈ L(Lp(Rd; X), Lq(Rd; Y )) with keTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ C sup ξd/rkm(ξ)kL(X,Y ) for some C ≥ 0 independent of m. Proof. First note that m is of moderate growth at infinity, where we use that r > 1. Hence Tm : S(Rd) ⊗ X → S′(Rd; Y ) is well-defined. Now the result follows by applying Theorem 3.24 to the symbol ξ 7→ ξd/rm(ξ) ∈ L(X, Y ), since f 7→ Tξ−d/r(f ) is an isometric isomorphism Lp(Rd; X) → H d/r (Rd; X) and Tm(f ) = Tξd/rm(ξ)(Tξ−d/r(f )) for f ∈ S(Rd; X). (cid:3) 3.6. Converse results and comparison. In the next result we show that in certain situations the type p of X (or cotype q of Y ) is necessary in Theorems 1.1, 3.18 and 3.21. The technique is a variation of the argument of Proposition 3.15 and in particular Lemma 3.14. ξ∈Rd\{0} p 22 JAN ROZENDAAL AND MARK VERAAR Lemma 3.26. Let X be a Banach space with cotype 2 and let Y be a Banach space p− 1 with type 2. Let p ∈ (1, 2], q ∈ [2,∞) and r ∈ (1,∞] be such that 1 q . Assume that for all strongly measurable m : Rd → L(X, Y ) for which {ξ r m(ξ) ξ ∈ Rd} is γ-bounded, the operator Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. Then r kbf (ξ)kXkbg(ξ)kY ∗ dξ ≤ CkfkLp(Rd;X)kgkLq′ (Rd;Y ∗) (3.16) ZRd ξ− d for some C ≥ 0 and all f ∈ S(Rd; X) and g ∈ S(Rd; Y ∗). r = 1 d At first glance it might seem surprising that we use that X has cotype 2 and Y r m(ξ) ξ ∈ Rd} in a has type 2. This is to be able to handle the γ-bound of {ξ simple way. d Proof. Since X has cotype 2 and Y has type 2, the γ-boundedness and uniform r m(ξ) ξ ∈ Rd} are equivalent. Therefore, by the closed graph boundedness of {ξ theorem there is a constant C such that for all f ∈ Lp(Rd; X) and g ∈ Lq′ (Rd; Y ∗) d Hence, S(Rd; Y ∗), d hTmf, gi ≤ C sup{ξ letting M (ξ) := ξ r km(ξ)kL(X,Y ) ξ ∈ Rd}kfkLp(Rd;X)kgkLq′ (Rd;Y ∗). r m(ξ), we see that for all f ∈ S(Rd; X) and g ∈ d (cid:12)(cid:12)(cid:12)ZRdhM (ξ)ξ− d r bf (ξ),bg(ξ)i dξ(cid:12)(cid:12)(cid:12) ≤ C sup{kM (ξ)k ξ ∈ Rd}kfkLp(Rd;X)kgkLq′ (Rd;Y ∗). Taking the supremum over all strongly measurable M which are uniformly bounded by 1, a similar approximation argument as in Lemma 3.14 yields the desired result. (cid:3) d r = 1 Proposition 3.27. Let X and Y be Banach spaces. Let p ∈ (1, 2], q ∈ [2,∞) and p − 1 r ∈ (1,∞] be such that 1 q . Assume that for all X-strongly measurable m : Rd → L(X, Y ) such that {ξ r m(ξ) ξ ∈ Rd} is γ-bounded, the operator Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded. Then the following assertions hold: (1) If X has cotype 2, Y = C, and q = 2, then X has type p. (2) If Y has type 2, X = C, and p = 2, then Y has cotype q. (3) If Y = X∗ has type 2, and q = p′, then X has type p. Proof. First consider (1). From (3.16) we find that for all f ∈ S(Rd; X) and g ∈ S(Rd), r kbf (ξ)kXbg(ξ) dξ ≤ CkfkLp(Rd;X)kbgkL2(Rd). kξ 7→ ξ− d Taking the supremum over all g with kgkL2(Rd) = kbgkL2(Rd) = 1, we obtain (3.17) By an approximation argument this estimate extends to all f ∈ Lp(Rd; X). particular, let f (t) := Pk≤n 1[− 1 t ∈ Rd. Then r bf (ξ)kL2(Rd;X) ≤ CkfkLp(Rd;X). kbf (ξ)k = ζ(ξ)(cid:13)(cid:13)(cid:13) Xk≤n In 2 )d (t + k)xk for n ∈ N, x1, . . . , xn ∈ X and ek(ξ)xk(cid:13)(cid:13)(cid:13), 2 , 1 ZRd ξ− d FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 23 where ζ(ξ) :=Qd j=1 sin(πξj ) for some cd > 0 and all ξ ∈ [ 1 πξj and ek(ξ) := e2πik·ξ for ξ ∈ Rd. Since ζ(ξ)ξ−d/r ≥ cd 2 , 1 2 ]d, it follows from (3.17) that Let (γk)k≤n be a Gaussian sequence. Replacing xk by γkxk, and taking L2(Ω)- norms, we find that 2 ]d;X) ≤ Cc−1 2 , 1 ekxk(cid:13)(cid:13)(cid:13)L2([− 1 (cid:13)(cid:13)(cid:13) Xk≤n γkxk(cid:13)(cid:13)(cid:13)L2(Ω;X) ≤ Cc−1 (cid:13)(cid:13)(cid:13) Xk≤n d (cid:16) Xk≤n kxkkp(cid:17) 1 p . d (cid:16) Xk≤n kxkkp(cid:17) 1 p . 2 ]d, (γkek(t))k≤n is identically dis- Case (2) can be proved in a similar way by reversing the roles of f and g. Indeed, Here we used the fact that for each t ∈ [− 1 tributed as (γk)k≤n. This implies that X has type p. this gives that Y ∗ has type q′ and hence Y has cotype q. 2 , 1 In case (3) we let f = g ∈ S(Rd; X) in (3.16) and argue as below (3.17). Here we use that X ⊆ X∗∗ has cotype 2 (see [17, Proposition 11.10]). (cid:3) If X = C, then (3.17) is a special case of Pitt's inequality (see [5] and [7]): kξ 7→ ξ−αbf (ξ)kLq(Rd;X) ≤ Cks 7→ sβf (s)kLp(Rd;X), q + β − α = d. (3.18) where 1 < p ≤ q < ∞, 0 ≤ α < d Note that Theorem 3.18 and the proof of Proposition 3.27 show that (3.18) holds if X has type p0 > p and cotype 2. Moreover, by the proof above one sees that Pitt's inequality with β = 0 and q = 2 implies that X has type p and X∗ has type p. Moreover, in the case α = β = 0 and q = p′, Pitt's inequality is equivalent to X having Fourier type p. It seems that a vector-valued analogue of Pitt's inequality has never been studied in detail. This leads to the following natural open problem: q , 0 ≤ β < d p′ and d p + d Problem 3.28. Characterize those Banach spaces X for which Pitt's inequality (3.18) holds. For p-convex and q-concave Banach lattices, (3.18) can be proved by reducing to the scalar case using the technique of [21, Proposition 2.2]. Next we show that a γ-boundedness assumption cannot be avoided in general. In the case where p = q such a result is due Cl´ement and Pruss (see [29, Chapter 5]). In Proposition 3.9 and Theorem 3.12 we have seen that γ-boundedness is not needed for certain Lp-Lq-multiplier theorems. In the following result we derive the necessity of the γ-boundedness of {m(ξ) ξ ∈ Rd} under special conditions on m. Proposition 3.29. Assume 1 < p ≤ q < ∞ and let 1 q . Assume m : Rd → L(X, Y ) is such that there is a constant a > 0 such that m takes the constant value mk on each of the cubes Qa,k = a([0, 1]d + k) with k ∈ Zd. If Tm : Lp(Rd; X) → Lq(Rd; X) is bounded, then p − 1 r = 1 γ({mk k ∈ Z}) ≤ R({mk k ∈ Z}) ≤ Cd,p,q′ a−d/rkTmk for some Cd,p,q′ ≥ 0. In Example 3.30 we will provide an example where even the γ-boundedness of {ξd/rm(ξ) ξ ∈ Rd} is necessary. However, in general such a result does not hold (see Remark 3.31). 24 JAN ROZENDAAL AND MARK VERAAR Proof. From Proposition 3.4 and Remark 3.5 we obtain that (3.19) (cid:13)(cid:13)(cid:13) Xk≤n ekmkxk(cid:13)(cid:13)(cid:13)Lp(Td;Y ) ≤ a−d/rCd,p,q′kTmk(cid:13)(cid:13)(cid:13) Xk≤n ekxk(cid:13)(cid:13)(cid:13)Lp(Td;X) . Now the R-boundedness follows from [4]. For convenience we include a short ar- gument below. Let (εk)k≤n be a sequence of independent random variables which are uniformly distributed on Ω := [0, 1]d. Replacing xk by εkxk in (3.19) and integrating over Ω yields that (cid:13)(cid:13)(cid:13) Xk≤n εkmkxk(cid:13)(cid:13)(cid:13)Lp(Ω;Y ) εkekmkxk(cid:13)(cid:13)(cid:13)Lp(Ω×Td;Y ) =(cid:13)(cid:13)(cid:13) Xk≤n ≤ a−d/rCd,p,q′kTmk(cid:13)(cid:13)(cid:13) Xk≤n ≤ a−d/rCd,p,q′kTmk(cid:13)(cid:13)(cid:13) Xk≤n εkekxk(cid:13)(cid:13)(cid:13)Lp(Ω×Td;X) εkxk(cid:13)(cid:13)(cid:13)Lp(Ω;X) . Here we used the fact that for each t ∈ Td, (εkek(t))k≤n and (εk)k≤n are identi- cally distributed. Finally, the estimate for the γ-bound is well-known and follows from a random- (cid:3) ization argument. 1 p − 1 The following example, which is similar to Example 3.17, shows that Theorem In particular, it shows that the γ-boundedness 3.21 is sharp in a certain sense. condition is necessary in certain cases. Example 3.30. Let p ∈ [1, 2], and for q ∈ [2,∞) let r ∈ (1,∞] be such that r = 1 q . Let X := ℓu for u ∈ [1,∞). Let (ej)j∈N0 ⊆ X be the standard basis of X, and for k ∈ N0 let Sk ∈ L(X) be such that Sk(ej) := ej+k for j ∈ N0. Let m : R → L(ℓu) be given by m(ξ) := P∞k=1 ck1(k−1,k](ξ)Sk for ξ ∈ R, with ck := k−α log(k + 1)−2 for α ≥ 0 arbitrary but fixed for the moment. 2(cid:12)(cid:12). By (2.6) and [19, Theorem 3.1] we find v =(cid:12)(cid:12) 1 Let v ∈ [2,∞] be such that 1 u − 1 a constant C ≥ 0 such that r −α log(k + 1)−2Sk k ∈ N}) r −α log(k + 1)−2kSkkL(X))∞k=1kℓv r −α)v log(k + 1)−2v(cid:17) 1 k( 1 γ({ξ v 1 1 q + 1 v . r − α ≤ − 1 p − 1 q + 1 (with the obvious modification for v = ∞), and the latter expression is finite if and only if 1 If u ∈ [p, 2] then X is a p-convex and q-concave Banach lattice for all q ≥ p, hence by Theorem 3.21 we find that with α = 1 2 , Tm : Lp(R; X) → Lq(R; X) is bounded for all q ≥ 2. Note that for q = 2 and u > p, m is more singular than in Example 3.17, where we used Proposition 3.9 to obtain the boundedness of Tm : Lp(R; X) → Lp′ u − 1. In the special case where u = p, both results can be combined using complex interpolation to obtain that Tm : Lp(R; X) → Lq(R; X) is bounded for all q ∈ [2, p′] if α = 2 pronounced when p = u = 1. Note also that the difference between Proposition 3.9 and Theorem 3.21 is most In this case X = ℓ1 has trivial type and trivial (R; X) for α = 1 p − 1 p − 1. p′ > 1 p + 1 u − 1 1 r m(ξ) ξ ∈ R}) ≤ γ({k ≤ Ck(k ≤ C(cid:16) ∞Xk=1 v , i.e. if and only if α ≥ 1 p − 1 FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 25 p − 1 u − 1 2(cid:12)(cid:12)(cid:12) for Fourier type, but cotype q = 2. Hence Proposition 3.9 only yields the boundedness of Tm : L1(R; X) → L∞(R; X) for α ≥ 1, which can also be obtained trivially since in this case m is integrable. On the other hand, Theorem 3.21 yields the nontrivial statement that Tm : L1(R; X) → L2(R; X) is bounded for α ≥ 1. u the operator Tm : L2(R; X) → Lq(R; X) is bounded. In the special case that u = q, combined with Example 3.17 we find that Tm : Lp(R; X) → Lq(R; X) is bounded for all p ∈ [q′, 2] with α = 2 Now fix q ∈ [2,∞) and let u ∈ [2, q]. Then, similarly, with α = 1 − 1 q′ − 1. q − 1 We now show that in certain cases the condition α ≥ 1 uc2nkϕ1kLq(R) = Cn p for p > 1. Therefore, α ≥ 1 the γ-boundedness of {ξ1/rm(ξ) ξ ∈ R} from above is sharp in order for Tm : Lp(R; X) → Lq(R; X) to be bounded. First suppose that u ∈ [1, 2]. For k ∈ N k=n+1 ϕke0. Then, as in Example 3.17, we find that q + (cid:12)(cid:12)(cid:12) 1 let ϕk : R → C be such that cϕk = 1(k−1,k], and for n ∈ N let f := P2n and kfkLp(R;X) ≤ C2n1− 1 p − 1. This shows that for q = 2 and u ∈ [1, 2], the condition on α which guarantees γ-boundedness is necessary. In the case u ∈ [2,∞), a duality argument shows that α ≥ 1 q′ − 1 = 1 − 1 q , which shows that the γ-boundedness condition is also necessary if p = 2 and u ∈ [2,∞). and Tm : Lp(R; X) → L2(R; X) is bounded, then 1 u ≥ p. Similarly, if Tm : L2(R; X) → Lq(R; X) is bounded with α = 1 − 2 u′ ≤ 1 − 1 By considering mn(ξ) :=Pn X = ℓp with p ∈ [1, 2] and 1 u − 1 Recall from the last part of Example 3.17 that if u ∈ [1,∞) and α = 2 q′ , and thus u ≤ q. k=1 1(k−1,k](ξ)Sk a similar argument yields that for p − 1 2 , one has p − 1 p and thus q , then kTm(f )kLq(R;X) ≥ n u ≤ 1 − 1 p + α = 1 1 q′ + α = 1 u′ + 1 r = 1 1 1 u n−α log(n)−2 u + 1 kTmnkL(Lp(R;X),L2(R;X)) hp γ({ξ r mn(ξ) ξ ∈ R}). 1 In particular this shows that the γ-bound provides the right factor in certain cases. In the following remark we show that one cannot prove the γ-boundedness, or even the uniform boundedness, of {ξd/rm(ξ) ξ ∈ Rd} in general. Remark 3.31. Let m : Rd \ {0} → L(X, Y ) be X-strongly measurable. If r < ∞, then one cannot prove sup{ξσkm(ξ)k ξ ∈ Rd \ {0}} ≤ CkTmk for any σ ∈ R. Indeed, σ ≤ 0 is not possible for the multiplier m(ξ) := ξ−d/r which is unbounded near zero. For σ > 0, one can use the same multiplier and a translation argument to deduce a contradiction. Moreover, for any nonzero multiplier m one can consider mh = m(· − h) for h ∈ Rd. Then kTmk = kTmhk and it follows that ξ0 + hσkm(ξ0)k = sup{ξσkm(ξ − h)k ξ ∈ Rd \ {0}} ≤ CkTmhk = CkTmk for all ξ0 ∈ Rd. Letting h → ∞ yields a contradiction whenever m(ξ0) 6= 0. In the next remark we compare the results obtained in this section with the ones obtained by Fourier type methods. Remark 3.32. 26 JAN ROZENDAAL AND MARK VERAAR (i) Consider the case of scalar-valued multipliers m. If X = Y has Fourier type (Rd); X)) for all p′ . This class of multipliers is larger than the p0 > p, then Theorem 3.12 states that Tm ∈ L(Lp(Rd; X), Lp′ m ∈ Lr,∞(Rd), where 1 one obtained in Theorem 3.18 since p − 1 r = 1 sup{ξ d r m(ξ) ξ ∈ Rd} ≤ CdkmkLr,∞(Rd). On the other hand, the geometric conditions in Theorem 3.18 are less restric- tive. Indeed, Fourier type p0 implies that X has type p0 and cotype p′0, but the converse is false. (ii) An important difference between Proposition 3.9 and Theorem 3.12 and the results obtained in Subsections 3.4 and 3.5 is that the former do not require any γ-boundedness condition. Of course the assumptions on type and cotype are less restrictive, and furthermore by [31] the γ-boundedness can be avoided w,1(Rd;L(X, Y )) for if X has cotype u and Y has type v and · w = 1 r m(·) ∈ B d w 1 d u − 1 v . In this case γ({ξ d r m(ξ) ξ ∈ Rd}) ≤ k · d r m(·)kB d w . w,1(Rd;L(X,Y )) 4. Extrapolation In this section we briefly discuss an extension of the extrapolation results of Hormander in [28]. Let m : Rd \ {0} → L(X, Y ) be a strongly measurable map of moderate growth at zero and infinity. For r ∈ [1,∞), ∈ [1,∞) and n ∈ N, consider the following variants of the Mihlin–Hormander condition: r − d Rα+ d (M1)r,,n There exists a constant M1 ≥ 0 such that for all multi-indices α ≤ n, (x ∈ X, R > 0). (M2)r,,n There exists a constant M2 ≥ 0 such that for all multi-indices α ≤ n (y∗ ∈ Y ∗, R > 0). (cid:16)ZR≤ξ<2R k∂αm(ξ)xk dξ(cid:17)1/ (cid:16)ZR≤ξ<2R k∂αm(ξ)∗y∗k dξ(cid:17)1/ ≤ M2ky∗k ≤ M1kxk Rα+ d r − d In the case = 2, r = 1, X = Y = R, condition (M1)r,,n reduces to the classical Hormander condition in [28, Theorem 2.5] (see also [24, Theorem 5.2.7]). Now we can formulate the main result of this section. It extends [28, Theorem 1 q0 = 1 2.5] to the vector-valued setting and to general exponents p, q ∈ (1,∞). Theorem 4.1 (Extrapolation). Let p0, q0, r ∈ [1,∞] with r 6= 1 be such that 1 p0 − r . Let m : Rd \ {0} → L(X, Y ) be a strongly measurable map of moderate growth at zero and infinity. Suppose that Tm : Lp0 (Rd; X) → Lq0 (Rd; Y ) is bounded of norm B. (1) Suppose that p0 ∈ (1,∞], Y has Fourier type ∈ [1, 2] with ≤ r, and (M1)r,,n holds for n := ⌊ d r⌋ + 1. Then Tm ∈ L(Lp(Rd; X), Lq(Rd; Y )) and − d kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Cp0,q0,p,d(M1 + B) (4.1) for all (p, q) such that p ∈ (1, p0] and 1 as p ↓ 1. p − 1 q = 1 r , where Cp0,q0,p,d ∼ (p − 1)−1 FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 27 (2) Suppose that q0 ∈ (1,∞), X has Fourier type ∈ [1, 2] with ≤ r, and (M2)r,,n holds for n := ⌊ d r⌋ + 1. Then Tm ∈ L(Lp(Rd; X), Lq(Rd; Y )) and − d kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Cp0,q0,q,d(M2 + B), (4.2) for all (p, q) satisfying q ∈ [q0,∞) and 1 r , where Cp0,q0,q,d ∼ q as q ↑ ∞. The proof will be presented in [50]. It is based on the classical argument in the case p = q (see [24, Theorem 5.2.7]). One of the other ingredients is an operator- valued analogue of [28, Theorem 2.2]. p − 1 q = 1 As a consequence we obtain the following extrapolation result: Corollary 4.2. Let p0, q0, r ∈ [1,∞] with q0 6= 1 and r 6= 1 be such that 1 − d Let X and Y both have Fourier type ∈ [1, 2] ≤ r and let n := ⌊ d m : Rd \ {0} → L(X, Y ) be such that, for all multi-indices α ≤ n, (4.3) Suppose that Tm : Lp0 (Rd; X) → Lq0 (Rd; Y ) is bounded of norm B. Then, for all exponents p and q satisfying 1 < p ≤ q < ∞ and 1 q , Tm : Lp(Rd; X) → Lq(Rd; Y ) is bounded and k∂αm(ξ)k ≤ Cξ−α− d p0 − 1 = 1 r . r⌋ + 1. Let ξ ∈ Rd \ {0}. p − 1 q = 1 r , q0 kTmkL(Lp(Rd;X),Lq(Rd;Y )) ≤ Cp,q,d(B + C) for some constant Cp,q,d ≥ 0. In particular, one can always take = 1 and n = ⌊ d r′⌋ + 1 in the above results. Proof. Note that, for ξ ∈ Rd, x ∈ X and y∗ ∈ Y ∗, km(ξ)xkY ≤ km(ξ)kL(X,Y ) kxkX and km∗y∗kX ∗ ≤ km(ξ)kL(X,Y ) ky∗kY ∗ , and similarly for the derivatives of m. Therefore, the result follows from Theorem 4.1 (1) and (2). Indeed, (i) p0, q0 ∈ (1,∞): apply (1) and (2). (ii) p0 ∈ (1,∞], q0 = ∞: apply (1). (iii) p0 = 1, q0 ∈ (1,∞): apply (2). (iv) p0 = 1, q0 = ∞ is not possible, since r 6= 1. (v) p0 = 1, q0 = 1 is not possible, since q0 6= 1. (cid:3) If p0 = q0 = 1, then Theorem 4.1 and Corollary 4.2 are true with = 1 (see [50]). Next we consider several applications of these extrapolation results. In [42] an Lp-Lq-Fourier multiplier result was proved assuming differentiability up to order d. Moreover, in [52] an extension is discussed in the case d = 1. We prove a similar result in the Hilbert space case in arbitrary dimensions assuming less differentiability. Example 4.3. Let X and Y be Hilbert spaces. First consider r ∈ (2,∞] and let n := ⌊d( 1 r )⌋ + 1 and assume that m : Rd \ {0} → C is such that for all α ≤ n (4.4) Then Tm : Lp(Rd; X) → Lq(Rd; X) is bounded for all 1 < p ≤ q < ∞ such that p − 1 r . Indeed, we first prove the boundedness of Tm in special cases. If r = ∞, then one can take p0 = q0 = 2 and the boundedness of Tm from L2(Rd; X) into L2(Rd; Y ) follows from Plancherel's isometry and the uniform boundedness of m. If ∂αm(ξ) ≤ Cξ−α− d (ξ ∈ Rd \ {0}). 2 − 1 q = 1 1 r 28 JAN ROZENDAAL AND MARK VERAAR r < ∞, then we can find p0 ∈ (1, 2) and q0 ∈ (2,∞) such that 1 r . Since X and Y have Fourier type 2 the boundedness of Tm from Lp0(Rd; X) into Lq0 (Rd; Y ) follows from Theorem 3.12. Now Corollary 4.2 can be applied to extrapolate the boundedness to the remaining cases. p0 − 1 = 1 q0 q = 1 p − 1 Next let r ∈ (1, 2]. Then all p, q ∈ (1,∞) satisfying 1 r are such that p ∈ (1, 2) and q ∈ (2,∞). Hence each m satisfying (4.4) for α = 0 yields a bounded operator Tm : Lp(Rd; X) → Lq(Rd; Y ) for all such p, q by Theorem 3.12. Remark 4.4. Even in the case where X = Y = C (or X and Y are Hilbert spaces) the result in Corollary 4.2 with ρ = 2 was only known for r = ∞. The point is that we only need derivatives up to order ⌊d( 1 r )⌋ + 1 if r > 2, whereas the classical condition requires derivatives up to ⌊d/2⌋ + 1. However, if m would have derivatives up to order n := ⌊d/2⌋ + 1 for which (4.3) holds, then the multiplier M (ξ) := ξd/rm(ξ) would satisfy the classical Mihlin condition: for all α ≤ n 2 − 1 k∂αM (ξ)k ≤ Cξ−α (ξ ∈ Rd \ {0}). Therefore, TM ∈ L(Lp(Rd), Lp(Rd)) for all p ∈ (1,∞). Consequently we find that, for any 1 < p ≤ q < ∞ with 1 p − 1 q = 1 r , kTmkL(Lp(Rd),Lq(Rd)) ≤ kTMkL(Lp(Rd),Lp(Rd))kTξ−d/rkL(Lp(Rd),Lq(Rd)) < ∞, where we used the Hardy-Littlewood-Sobolev inequality (see Example 3.3). For r ≤ 2 we have already observed in Example 4.3 that in the Hilbertian setting no derivatives are required. Thus in the scalar or Hilbertian setting we emphasize that the only new point is that less derivatives are required of the multiplier for p < q. In the case where X and Y are general Banach spaces, the assertion about Tξ−d/r remains true. However, the boundedness of TM is not as simple to obtain and in general requires geometric conditions on X (even if m is scalar-valued) and an R-boundedness version of the Mihlin condition (see [38]). Another application of Corollary 4.2 is that we can extrapolate the result of Theorem 3.18 to other values of p and q. A similar result holds for Theorem 3.21. Corollary 4.5. Let X be a Banach space with type p0 ∈ (1, 2] and Y a Banach space with cotype q0 ∈ [2,∞), and let p1 ∈ (1, p0) and q1 ∈ (q0,∞), r ∈ [1,∞] be such that 1 r m(ξ) ξ ∈ Rd \ {0}} ⊆ L(X, Y ) is γ-bounded. n := ⌊d( 1 (4.5) Assume that X and Y both have Fourier type ∈ [1, 2] with ≤ r and let . Let m : Rd \ {0} → L(X, Y ) be such that {ξ r )⌋ + 1. Assume for all multi-indices α ≤ n p1 − 1 − 1 r = 1 q1 d k∂αm(ξ)k ≤ Cξ−α− d r (ξ ∈ Rd \ {0}). Then Tm extends uniquely to a bounded map fTm ∈ L(Lp(Rd; X), Lq(Rd; Y )) for all 1 < p ≤ q < ∞ satisfying 1 Proof. The case where p = p1 and q = q1 follows from Theorem 3.18. The result for the remaining values of p and q follows from Corollary 4.2. (cid:3) p − 1 q = 1 r . [1] F. Albiac and N. J. Kalton. Topics in Banach space theory, volume 233 of Graduate Texts in Mathematics. Springer, New York, 2006. References FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 29 [2] H. Amann. Operator-valued Fourier multipliers, vector-valued Besov spaces, and applications. Math. Nachr., 186:5–56, 1997. [3] W. Arendt, C. J. K. Batty, M. Hieber, and F. Neubrander. Vector-valued Laplace transforms and Cauchy problems, volume 96 of Monographs in Mathematics. Birkhauser/Springer Basel AG, Basel, second edition, 2011. [4] W. Arendt and S. Bu. The operator-valued Marcinkiewicz multiplier theorem and maximal regularity. Math. Z., 240(2):311–343, 2002. [5] W. Beckner. Pitt's inequality with sharp convolution estimates. Proc. Amer. Math. Soc., 136(5):1871–1885, 2008. [6] A. Benedek, A.-P. Calder´on, and R. Panzone. Convolution operators on Banach space valued functions. Proc. Nat. Acad. Sci. U.S.A., 48:356–365, 1962. [7] J.J. Benedetto and H.P. Heinig. Weighted Fourier inequalities: new proofs and generaliza- tions. J. Fourier Anal. Appl., 9(1):1–37, 2003. [8] J. Bergh and J. Lofstrom. Interpolation spaces. An introduction. Springer-Verlag, Berlin, 1976. Grundlehren der Mathematischen Wissenschaften, No. 223. [9] E. Berkson and T. A. Gillespie. Spectral decompositions and harmonic analysis on UMD spaces. Studia Math., 112(1):13–49, 1994. [10] J. Bourgain. Some remarks on Banach spaces in which martingale difference sequences are unconditional. Ark. Mat., 21(2):163–168, 1983. [11] J. Bourgain. Vector-valued singular integrals and the H 1-BMO duality. Probability theory and harmonic analysis, Pap. Mini-Conf., Cleveland/Ohio 1983, Pure Appl. Math., Marcel Dekker 98, 1-19 (1986)., 1986. [12] D. L. Burkholder. A geometric condition that implies the existence of certain singular integrals of Banach-space-valued functions. In Conference on harmonic analysis in honor of Antoni Zygmund, Vol. I, II (Chicago, Ill., 1981), Wadsworth Math. Ser., pages 270–286. Wadsworth, Belmont, CA, 1983. [13] A. P. Calderon and A. Zygmund. On the existence of certain singular integrals. Acta Math., 88:85–139, 1952. [14] P. Cl´ement, B. de Pagter, F. A. Sukochev, and H. Witvliet. Schauder decompositions and multiplier theorems. Studia Math., 138(2):135–163, 2000. [15] K. de Leeuw. On Lp multipliers. Ann. of Math. (2), 81:364–379, 1965. [16] R. Denk, M. Hieber, and J. Pruss. R-boundedness, Fourier multipliers and problems of elliptic and parabolic type. Mem. Amer. Math. Soc., 166(788):viii+114, 2003. [17] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. [18] G. Dore and A. Venni. On the closedness of the sum of two closed operators. Math. Z., 196:189–201, 1987. [19] O. van Gaans. On R-boundedness of unions of sets of operators. In Partial differential equations and functional analysis, volume 168 of Oper. Theory Adv. Appl., pages 97–111. Birkhauser, Basel, 2006. [20] J. Garc´ıa-Cuerva, K. S. Kazaryan, V. I. Kolyada, and J. L. Torrea. The Hausdorff-Young inequality with vector-valued coefficients and applications. Uspekhi Mat. Nauk, 53(3(321)):3– 84, 1998. [21] J. Garc´ıa-Cuerva, J. L. Torrea, and K. S. Kazarian. On the Fourier type of Banach lattices. In Interaction between functional analysis, harmonic analysis, and probability (Columbia, MO, 1994), volume 175 of Lecture Notes in Pure and Appl. Math., pages 169–179. Dekker, New York, 1996. [22] M. Girardi and L. Weis. Criteria for R-boundedness of operator families. In Evolution equa- tions, volume 234 of Lecture Notes in Pure and Appl. Math., pages 203–221. Dekker, New York, 2003. [23] M. Girardi and L. Weis. Operator-valued Fourier multiplier theorems on Lp(X) and geometry of Banach spaces. J. Funct. Anal., 204(2):320–354, 2003. [24] L. Grafakos. Classical Fourier analysis, volume 249 of Graduate Texts in Mathematics. Springer, New York, second edition, 2008. [25] L. Grafakos. Modern Fourier analysis, volume 250 of Graduate Texts in Mathematics. Springer, New York, second edition, 2009. [26] R. Haller, H. Heck, and A. Noll. Mikhlin's theorem for operator-valued Fourier multipliers in n variables. Math. Nachr., 244:110–130, 2002. 30 JAN ROZENDAAL AND MARK VERAAR [27] M. Hieber. Operator valued Fourier multipliers. In Topics in nonlinear analysis, volume 35 of Progr. Nonlinear Differential Equations Appl., pages 363–380. Birkhauser, Basel, 1999. [28] L. Hormander. Estimates for translation invariant operators in Lp spaces. Acta Math., 104:93– 140, 1960. [29] T. Hytonen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach Spaces. Volume I: Martingales and Littlewood-Paley Theory, volume 63 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3). Springer, 2016. [30] T. Hytonen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach Spaces. Volume II. Probabilistic Methods and Operator Theory. 2016. Preliminary version at http://fa.its.tudelft.nl/~neerven/ . [31] T. Hytonen and M. Veraar. R-boundedness of smooth operator-valued functions. Integral Equations Operator Theory, 63(3):373–402, 2009. [32] T.P. Hytonen. New thoughts on the vector-valued Mihlin-Hormander multiplier theorem. Proc. Amer. Math. Soc., 138(7):2553–2560, 2010. [33] N. Kalton, J. van Neerven, M. Veraar, and L. Weis. Embedding vector-valued Besov spaces into spaces of γ-radonifying operators. Math. Nachr., 281(2):238–252, 2008. [34] N. J. Kalton and L. Weis. The H∞-calculus and sums of closed operators. Math. Ann., 321(2):319–345, 2001. [35] N.J. Kalton and L.W. Weis. The H∞-calculus and square function estimates. In Nigel J. Kalton Selecta, Volume 1, pages 715–764. Springer, 2016. [36] Y. Katznelson. An introduction to harmonic analysis. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2004. [37] S. G. Kreın, Yu. ¯I. Petun¯ın, and E. M. Semenov. Interpolation of linear operators, volume 54 of Translations of Mathematical Monographs. American Mathematical Society, Providence, R.I., 1982. Translated from the Russian by J. Szucs. [38] P. C. Kunstmann and L. Weis. Maximal Lp-Regularity for Parabolic Equations, Fourier Multiplier Theorems and H∞-functional Calculus. In Functional Analytic Methods for Evo- lution Equations (Levico Terme 2001), volume 1855 of Lecture Notes in Math., pages 65–312. Springer, Berlin, 2004. [39] S. Kwapie´n, M. Veraar, and L. Weis. R-boundedness versus γ-boundedness. Ark. Mat., 54(1):125–145, 2016. [40] P.-K. Lin. Kothe-Bochner function spaces. Birkhauser Boston, Inc., Boston, MA, 2004. [41] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. II, volume 97 of Ergebnisse der Mathematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas]. Springer- Verlag, Berlin-New York, 1979. Function spaces. [42] P. I. Lizorkin. Multipliers of Fourier integrals in the spaces Lp, θ. Trudy Mat. Inst. Steklov, 89:231–248, 1967. [43] T. R. McConnell. On Fourier multiplier transformations of Banach-valued functions. Trans. Amer. Math. Soc., 285(2):739–757, 1984. [44] S. Montgomery-Smith. Stability and dichotomy of positive semigroups on Lp. Proc. Amer. Math. Soc., 124(8):2433–2437, 1996. [45] Richard O'Neil. Convolution operators and L(p, q) spaces. Duke Math. J., 30:129–142, 1963. [46] A. Pietsch and J. Wenzel. Orthonormal systems and Banach space geometry, volume 70 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1998. [47] G. Pisier. Some results on Banach spaces without local unconditional structure. Compositio Math., 37(1):3–19, 1978. [48] D. Potapov and F. Sukochev. Operator-Lipschitz functions in Schatten-von Neumann classes. Acta Math., 207(2):375–389, 2011. [49] J. Rozendaal. Functional calculus for C0-groups using (co)type. Online at http://arxiv.org/abs/1508.02036, 2015. [50] J. Rozendaal and M. Veraar. Fourier multiplier theorems on Besov spaces under type and cotype conditions. To appear in Banach Journal of Mathematical Analysis. Online at http://arxiv.org/abs/1606.03272, 2016. [51] J. Rozendaal and M. Veraar. Stability theory for semigroups using (Lp, Lq) Fourier multipli- ers. In preparation, 2017. [52] L. O. Sarybekova, T. V. Tararykova, and N. T. Tleukhanova. On a generalization of the Lizorkin theorem on Fourier multipliers. Math. Inequal. Appl., 13(3):613–624, 2010. FOURIER MULTIPLIER THEOREMS INVOLVING TYPE AND COTYPE 31 [53] R. Shahmurov. On integral operators with operator-valued kernels. J. Inequal. Appl., page 12, 2010. [54] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory inte- grals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III. [55] Z. Strkalj and L. Weis. On operator-valued Fourier multiplier theorems. Trans. Amer. Math. Soc., 359(8):3529–3547 (electronic), 2007. [56] H. Triebel. Interpolation theory, function spaces, differential operators. Johann Ambrosius Barth, Heidelberg, second edition, 1995. [57] H. Triebel. Theory of function spaces. Modern Birkhauser Classics. Birkhauser/Springer Basel AG, Basel, 2010. Reprint of 1983 edition. [58] H. Triebel. Tempered homogeneous function spaces. Zurich: European Mathematical Society (EMS), 2015. [59] J. van Neerven. γ-radonifying operators-a survey. In The AMSI-ANU Workshop on Spectral Theory and Harmonic Analysis, volume 44 of Proc. Centre Math. Appl. Austral. Nat. Univ., pages 1–61. Austral. Nat. Univ., Canberra, 2010. [60] M. Veraar. Embedding results for γ-spaces. In Recent trends in analysis. Proceedings of the conference in honor of Nikolai Nikolski on the occasion of his 70th birthday, Bordeaux, France, August 31 – September 2, 2011, pages 209–219. Bucharest: The Theta Foundation, 2013. [61] L. Weis. Stability theorems for semi-groups via multiplier theorems. In Differential equations, asymptotic analysis, and mathematical physics (Potsdam, 1996), volume 100 of Math. Res., pages 407–411. Akademie Verlag, Berlin, 1997. [62] L. Weis. Operator-valued Fourier multiplier theorems and maximal Lp-regularity. Math. Ann., 319(4):735–758, 2001. [63] F. Zimmermann. On vector-valued Fourier multiplier theorems. Studia Math., 93(3):201–222, 1989. Institute of Mathematics Polish Academy of Sciences, ul. ´Sniadeckich 8, 00-656 War- saw, Poland E-mail address: [email protected] Delft Institute of Applied Mathematics, Delft University of Technology, P.O. Box 5031, 2628 CD Delft, The Netherlands E-mail address: [email protected]
1207.5375
2
1207
2013-05-14T14:47:09
Pisier's inequality revisited
[ "math.FA" ]
Given a Banach space $X$, for $n\in \mathbb N$ and $p\in (1,\infty)$ we investigate the smallest constant $\mathfrak P\in (0,\infty)$ for which every $f_1,...,f_n:{-1,1}^n\to X$ satisfy \int_{{-1,1}^n}\Bigg|\sum_{j=1}^n \partial_jf_j(\varepsilon)\Bigg|^pd\mu(\varepsilon) \leq \mathfrak{P}^p\int_{{-1,1}^n}\int_{{-1,1}^n}\Bigg\|\sum_{j=1}^n \d_j\Delta f_j(\varepsilon)\Bigg\|^pd\mu(\varepsilon) d\mu(\delta), where $\mu$ is the uniform probability measure on the discrete hypercube ${-1,1}^n$ and ${\partial_j}_{j=1}^n$ and $\Delta=\sum_{j=1}^n\partial_j$ are the hypercube partial derivatives and the hypercube Laplacian, respectively. Denoting this constant by $\mathfrak{P}_p^n(X)$, we show that $\mathfrak{P}_p^n(X)\le \sum_{k=1}^{n}\frac{1}{k}$ for every Banach space $(X,|\cdot|)$. This extends the classical Pisier inequality, which corresponds to the special case $f_j=\Delta^{-1}\partial_j f$ for some $f:{-1,1}^n\to X$. We show that $\sup_{n\in \N}\mathfrak{P}_p^n(X)<\infty$ if either the dual $X^*$ is a $\mathrm{UMD}^+$ Banach space, or for some $\theta\in (0,1)$ we have $X=[H,Y]_\theta$, where $H$ is a Hilbert space and $Y$ is an arbitrary Banach space. It follows that $\sup_{n\in \N}\mathfrak{P}_p^n(X)<\infty$ if $X$ is a Banach lattice of finite cotype.
math.FA
math
Pisier's inequality revisited Tuomas Hytonen Department of Mathematics and Statistics University of Helsinki P.O.B. 68, FI-00014 Helsinki, Finland E-mail: [email protected] Assaf Naor Courant Institute of Mathematical Sciences New York University 251 Mercer Street, New York NY, 10012, USA E-mail: [email protected] Abstract Given a Banach space X, for n ∈ N and p ∈ (1,∞) we investigate the smallest constant P ∈ (0,∞) for which every n-tuple of functions f1, . . . , fn : {−1, 1}n → X satisfies p dµ(ε)dµ(δ), nXj=1 δj∆fj(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) where µ is the uniform probability measure on the discrete hypercube {−1, 1}n and {∂j}n j=1 ∂j are the hypercube partial derivatives and the hypercube Laplacian, respectively. Denoting this constant by Pn p (X), we show that Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 p dµ(ε) ∂j fj(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 6 PpZ{−1,1}nZ{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) j=1 and ∆ = Pn nXk=1 p (X) 6 Pn 1 k for every Banach space (X,k · k). This extends the classical Pisier inequality, which corresponds to the special case fj = ∆−1∂j f for some f : {−1, 1}n → X. We show that supn∈N Pn p (X) < ∞ if either the dual X∗ is a UMD+ Banach space, or for some θ ∈ (0, 1) we have X = [H, Y ]θ, where H is a Hilbert space and Y is an arbitrary Banach space. It follows that supn∈N Pn p (X) < ∞ if X is a Banach lattice of nontrivial type. 1 2 1 Introduction T. Hytonen and A. Naor Fix a Banach space (X,k · k) and n ∈ N. For every f : {−1, 1}n → X and j ∈ {1, . . . , n} the hypercube jth partial derivative of f , which is denoted ∂jf : {−1, 1}n → X, is defined as (1.1) ∂jf (ε) def= f (ε) − f (ε1, . . . , εj−1,−εj, εj+1, . . . , εn) 2 . The hypercube Laplacian of f , denoted ∆f : {−1, 1}n → X, is (1.2) ∆f (ε) def= ∂jf (ε). nXj=1 It is immediate to check that ∆ is invertible on the space of all mean zero functions f : {−1, 1}n → X. Below ∆−1 is understood to be defined for every f : {−1, 1}n → X by setting ∆−1f = ∆−1f . Here f = f−R{−1,1}n f (δ)dµ(δ), where µ denotes the uniform probability measure on {−1, 1}n. The following inequality is due to Pisier [28]. Throughout this paper the asymptotic notation ., & indicates the corresponding inequalities up to universal constant factors. We will also denote equivalence up to universal constant factors by ≍, i.e., A ≍ B is the same as (A . B) ∧ (A & B). Theorem 1.1 (Pisier's inequality). For every Banach space (X,k·k), every n ∈ N, every p ∈ [1,∞] and every f : {−1, 1}n → X, we have (1.3) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p f (δ)dµ(δ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dµ(ε)!1/p f (ε) −Z{−1,1}n δj∂jf (ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) . log n Z{−1,1}nZ{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 p dµ(ε)dµ(δ)!1/p . Due to the application of Pisier's inequality to the theory of nonlinear type (see [28, 25, 12, 23]), it is of great interest to understand when (1.3) holds true with the log n term replaced by a constant that may depend on the geometry of X but is independent of n. Talagrand proved [30] that the log n term in (1.3) is asymptotically optimal for general Banach spaces X, Wagner proved [31] that the log n term in (1.3) can be replaced by a universal constant if p = ∞ and X is a general Banach space, and in [25] it is shown that the log n term in (1.3) can be replaced by a constant that is independent of n if X is a UMD Banach space. It remains an intriguing open question whether every Banach space of nontrivial type satisfies (1.3) with Pisier's inequality revisited 3 the log n term replaced by a constant that is independent of n. If true, this would resolve a 1976 question of Enflo [9] by establishing that Rademacher type p and Enflo type p coincide (see [25, 23] and Section 6 below). Here we obtain a new class of Banach spaces that satisfies a dimension- independent Pisier inequality. Our starting point is the following extension of Pisier's inequality. Definition 1.2 (Pisier constant of X). The n-dimensional Pisier constant of X (with exponent p), denoted Pn p (X), is the infimum over those P ∈ (0,∞) such that every f1, . . . , fn : {−1, 1}n → X satisfy (1.4) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) We also set dµ(ε)dµ(δ)!1/p . nXj=1 p δj∆fj(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p ∂jfj(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dµ(ε)!1/p nXj=1 6 P Z{−1,1}nZ{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Pp(X) def= sup Pn p (X). n∈N Inequality (1.4) reduces to Pisier's inequality if we choose fj = ∆−1∂jf for some f : {−1, 1}n → X. The generalized inequality (1.4) has the ad- vantage of being well-behaved under duality, as explained in Section 2. The following theorem yields a logarithmic bound on Pp n(X), thus extending Pisier's inequality. Theorem 1.3. For every Banach space X, every p ∈ [1,∞] and every n ∈ N, Pn p (X) 6 1 k . nXk=1 Our approach yields a quantitative improvement over Pisier's inequality only in lower order terms: an optimization of Pisier's argument (as carried out in [23]) shows that the O(log n) term in (1.3) can be taken to be at most log n + O(log log n), while Theorem 1.3 shows that this term can be taken to be log n + O(1). In [25] it was shown that the logarithmic term in (1.3) can be replaced by a constant that is independent of n if X is a UMD Banach space. Recall that X is a UMD Banach space if for every p ∈ (1,∞) there exists a constant β ∈ (0,∞) such that if {Mj}n j=0 is a p-integrable X-valued martingale 4 T. Hytonen and A. Naor defined on some probability space (Ω, P), then for every ε1, . . . , εn ∈ {−1, 1} we have (1.5) M0 + ZΩ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 εj(Mj − Mj−1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p dP 6 βpZΩ kMnkpdP. The infimum over those β ∈ (0,∞) for which (1.5) holds true is denoted βp(X). It can be shown (see [7]) that βp(X) . p2 p−1β2(X), so in order to define the UMD property it suffices to require the validity of (1.5) for p = 2. UMD Banach spaces are known to be superreflexive [20, 1], and one also has βq(X ∗) = βp(X), where q = p/(p − 1) (see e.g. [7]). In [10] Garling investigated the natural weakening of (1.5) in which the desired inequality is required to hold true in expectation over ε1, . . . , εn ∈ {−1, 1} rather than for every ε1, . . . , εn ∈ {−1, 1}. Specifically, say that X is a UMD+ Banach space if for every p ∈ (1,∞) there exists a constant β ∈ (0,∞) such that if {Mj}n j=0 is a p-integrable X-valued martingale defined on some probability space (Ω, P) then (1.6) Z{−1,1}nZΩ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) M0 + nXj=1 εj(Mj − Mj−1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p dPdµ(ε) 6 βpZΩ kMnkpdP. The infimum over those β for which (1.6) holds true is denoted β+ p (X). Theorem 1.4. If X is a Banach space such that X ∗ is UMD+ then the following inequality holds true. Fix p ∈ (1,∞) and n ∈ N. For every function F : {−1, 1}n × {−1, 1}n → X and j ∈ {1, . . . , n} denote δjF (ε, δ)dµ(δ). Fj(ε) def= Z{−1,1}n ∆−1∂jFj(ε)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) q (X ∗)(cid:18)Z{−1,1}nZ{−1,1}n kF (ε, δ)kpdµ(ε)dµ(δ)(cid:19)1/p dµ(ε)!1/p p nXj=1 6 β+ , Then (1.7) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) where q = p/(p − 1). For every f1, . . . , fn : {−1, 1}n → X, an application of Theorem 1.4 to j=1 δjfj(ε) yields the following estimate on the the function F (ε, δ) = Pn Pisier constant of a UMD+ Banach space. Pisier's inequality revisited 5 Corollary 1.5. Pp(X) 6 β+ q (X ∗). It is unknown if a UMD+ Banach space must also be a UMD Banach space, though it seems reasonable to conjecture that there are UMD+ spaces that are not UMD. Regardless of this, Theorem 1.4 and Corollary 1.5 are conceptually different from the result of [25], which relies on the full force of the UMD condition, i.e. it requires the validity of (1.5) for every choice of signs ε1, . . . , εn, while our argument needs such estimates to hold true only for an average choice of signs. We also have a quantitative improvement: in [25] it was shown that Pisier's inequality holds true with the O(log n) term in (1.3) replaced by βp(X) = βq(X ∗), while we obtain the same esti- q (X ∗) 6 βq(X ∗). Geiss mate with the O(log n) term in (1.3) replaced by β+ proved [11] that for every η ∈ (0, 1) there is Cη ∈ (0,∞) such that for every M > 1 there is a Banach space X that satisfies ∞ > βq(X ∗) > Cηβ+ q (X ∗)2−η > M. Remark 1.1. Inequality (1.7) is an extension of the generalized Pisier in- equality (1.4), but for general Banach spaces it behaves very differently: unlike the logarithmic behavior of Theorem 1.3, the best constant appearing in the right hand side of (1.7) for a general Banach space X must be at least i=1(1 + εiδiηi). a constant multiple of √n, as exhibited by the case X = L1(({−1, 1}n, µ), R) and F : {−1, 1}n × {−1, 1}n → X given by F (ε, δ)(η) =Qn Suppose that θ ∈ (0, 1) and X = [H, Y ]θ, where H is a Hilbert space and Y is an arbitrary Banach space. Here [·,·]θ denotes complex interpola- tion (see [3]). Theorem 1.6 below shows that in this case Pp(X) < ∞, and therefore Pisier's inequality holds true with the log n term in (1.3) replaced by a constant that is independent of n. Pisier proved [27] that every Banach lattice of nontrivial type (see [19]) is of the form [H, Y ]θ for some θ ∈ (0, 1), so we thus obtain the desired dimension independence in Pisier's inequality for Banach lattices of nontrivial type. This result does not follow from pre- viously known cases in which a dimension-independent Pisier inequality has been proved, since, as shown by Bourgain [4, 5], there exist Banach lattices of nontrivial type which are not UMD. Note, however, that we are still far from proving the conjectured dimension-independent Pisier inequality for Banach spaces with nontrivial type: any space of the form [H, Y ]θ admits an equivalent norm whose modulus of smoothness has power type 2/(1 + θ) (see [27, 8]), while there exist Banach spaces with nontrivial type that do not admit such an equivalent norm (see [14, 16, 15, 29]). 6 T. Hytonen and A. Naor Theorem 1.6. Let X, Y be Banach spaces and let H be a Hilbert space. Suppose that for some θ ∈ (0, 1) we have X = [H, Y ]θ. Then for every p ∈ (1,∞), Pp(X) 6 2 max{p, p/(p − 1)} . Remark 1.2. If r ∈ (2,∞) then the O(log n) term in Pisier's inequal- ity (1.3), when p = 2 and X = ℓr, can be replaced by O(r), due to the fact that β+ 2 (ℓr) ≍ r (this follows from Hitczenko's work [13], as explained to us by Mark Veraar). This bound also follows from Theorem 1.6. At the same time, an inspection of Talagrand's example in [30] shows that this term must be at least a constant multiple of log r. Determining the correct order of magnitude as r → ∞ of the constant in Pisier's inequality when X = ℓr remains an interesting open problem. 1 − θ 2 Duality The dimension n ∈ N will be fixed from now on. For p ∈ [1,∞] and a Banach space X, let Lp(X) denote the vector-valued Lebesgue space Lp(({−1, 1}n, µ), X). Thus Lp(Lp(X)) can be naturally identified with the space Lp(({−1, 1}n × {−1, 1}n, µ × µ), X). For f ∈ Lp(X) we denote its Fourier expansion by f = XA⊆{1,...,n}bf (A)WA, and where the Walsh function WA : {−1, 1}n → {−1, 1} corresponding to A ⊆ {1, . . . , n} is given by WA(ε1, . . . , εn) = Qi∈A εi, and the Fourier co- efficient bf (A) ∈ X is given by bf (A) =R{−1,1}n f (x)WA(x)dµ(x). Using this (standard) notation, we have ∀ i ∈ {1, . . . , n}, ∀ f ∈ Lp(X), bf (A)WA, i∈A ∂if = XA⊆{1,...,n} Abf (A)WA, Abf (A)WA. ∀ f ∈ Lp(X), ∆f = XA⊆{1,...,n} ∀ f ∈ Lp(X), ∆−1f def= XA⊆{1,...,n} nXi=1 bf ({i})W{i}. Rad(f ) def= A6=∅ 1 The Rademacher projection of f ∈ Lp(X) is defined as usual by Pisier's inequality revisited 7 def= (I − Rad)(Lp(X)). We denote below RadX The dual of (RadX,k · kLp(X)) is naturally identified with the quotient Lq(X ∗)/Rad⊥ def= Rad(Lp(X)) and Rad⊥ X X ∗, where q = p/(p − 1). Define an operator S : Lp(Lp(X)) → Lp(X) by (2.1) ∀ F ∈ Lp(Lp(X)), S(F ) def= nXj=1 ∆−1∂jbF ({j}). Using this notation, Theorem 1.4 is nothing more than the following oper- ator norm bound. kSkLp(Lp(X))→Lp(X) 6 β+ q (X ∗). The adjoint operator S∗ : Lq(X ∗) → Lq(Lq(X ∗)) is given by ∀ g ∈ Lq(X ∗), ∀ δ ∈ {−1, 1}n, S∗(g)(δ) = δj∆−1∂jg. nXj=1 Therefore Theorem 1.4 has the following equivalent dual formulation. Theorem 2.1 (Dual formulation of Theorem 1.4). Let Z be a UMD+ Ba- nach space. Then for every q ∈ (1,∞) and every g ∈ Lq(Z) we have Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 q δj∆−1∂jg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Lq(Z) 1/q dµ(δ) 6 β+ q (Z)kgkLq(Z). Theorem 2.1, and consequently also Theorem 1.4, will be proven in Sec- tion 3. Let T be the restriction of S to RadLp(X). Thus Pn p (X) = kTkRad Lp (X)→Lp(X) = kT ∗kLq(X ∗)→Lq(Lq(X ∗))/Rad⊥ Lq(X ∗) is given by . Lq (X ∗) The adjoint T ∗ : Lq(X ∗) → Lq(Lq(X ∗))/Rad⊥ nXj=1 ∀ g ∈ Lq(X ∗), ∀ δ ∈ {−1, 1}n, T ∗(g) = δj∆−1∂jg + Rad⊥ Lq(X ∗). Therefore Theorem 1.3 has the following equivalent dual formulation. Theorem 2.2 (Dual formulation of Theorem 1.3). Let Z be a Banach space and q ∈ [1,∞]. Then for every g ∈ Lq(Z) we have inf Φ∈Rad⊥ Lq (Z)Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Φ(δ) + nXj=1 q δj∆−1∂jg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Lq(Z) 1/q dµ(δ) 6 nXk=1 1 k!kgkLq(Z). 8 T. Hytonen and A. Naor Theorem 2.2, and consequently also Theorem 1.3, will be proven in Sec- θ = [H, Y ∗]θ (see [3]), we also have the following equiva- tion 3. Since [H, Y ]∗ lent dual formulation of Theorem 1.6. Theorem 2.3 (Dual formulation of Theorem 1.6). Let H be a Hilbert space, W a Banach space, and θ ∈ (0, 1). Set Z = [H, W ]θ. Then for every q ∈ (1,∞) and g ∈ Lq(Z), inf Ψ∈Rad⊥ Lq (Z)Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Ψ(δ) + nXj=1 q δj∆−1∂jg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 6 Lq(Z) 1/q dµ(δ) kgkLq(Z). Theorem 2.3, and consequently also Theorem 1.6, will be proven in Sec- 1 − θ 2 max{q, q/(q − 1)} tion 5. 3 Proof of Theorem 2.1 Fix q ∈ (1,∞) and g ∈ Lq(Z). Let Sn denote the symmetric group on {1, . . . , n}. For σ ∈ Sn and k ∈ {0, . . . , n} define gσ (3.1) gσ k ∈ Lq(Z) by k (ε) def= XA⊆{σ−1(1),...,σ−1(k)}bg(A)WA(ε) X g kXi=1 δσ−1(k+1),...,δσ−1(n)∈{−1,1} = 1 2n−k εσ−1(i)eσ−1(i) + δσ−1(i)eσ−1(i)! , nXi=k+1 where here, and in what follows, e1, . . . , en denotes the standard basis of Rn. Then {gσ implying that k=0 is a Z-valued martingale with gσ n = g and gσ k}n (3.2) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXk=1 δk(cid:0)gσ k − gσ q k−1(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Lq(Z) 1/q dµ(δ) In (3.2) we may replace {δk}n k=1, since these two sequences of signs have the same joint distribution. Then we make the change of variable j = σ−1(k), so that k = σ(j). Averaging the resulting inequality over σ ∈ Sn, and using the convexity of the norm, we see that k=1 by {δσ−1(k)}n 6 β+ q (Z)kgkLq(Z) . 0 = bg(∅), (3.3) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 n! Xσ∈Sn nXj=1 δj(cid:0)gσ σ(j) − gσ q σ(j)−1(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Lq(Z) 1/q dµ(δ) 6 β+ q (Z)kgkLq(Z) . Pisier's inequality revisited 9 It remains to note that for each δ ∈ {−1, 1}n we have δj {σ ∈ Sn : max σ(A) = σ(j)} 1 1 = nXj=1 n! Xσ∈Sn δj(cid:0)gσ σ(j)−1(cid:1) σ(j) − gσ nXj=1 δj X∅(A⊆{1,...,n} n! Xσ∈Sn max σ(A)=σ(j)bg(A)WA = XA⊆{1,...,n} A6=∅ Xj∈A = XA⊆{1,...,n} nXj=1 bg(A)WA δj∆−1∂jg. Pj∈A δj A A6=∅ n! = bg(A)WA (3.4) Due to (3.3) and (3.4) the proof of Theorem 2.1 is complete. 4 Proof of Theorem 2.2 The following lemma introduces an auxiliary function which is a variant of a similar function that was used by Pisier in [28]. Lemma 4.1. Let Z be a Banach space. Fix n ∈ N, q ∈ [1,∞] and t ∈ (0, 1). For g ∈ Lq(Z) define Gt ∈ Lq(Lq(Z)) by (t + (1 − t)δi) − tn 1 − t g. Gt(δ) def= 1 1 − t XA⊆{1,...,n}bg(A)WAYi∈A Rad(Gt)(δ) = XA⊆{1,...,n} tA−1Xj∈A A6=∅ δjbg(A)WA, (4.1) Then (4.2) and (4.3) kGtkLq(Lq(Z)) 6 1 − tn 1 − t kgkLq(Z). Proof. Identity (4.2) follows from (4.1) since for every A ⊆ {1, . . . , n}, Rad Yi∈A (t + (1 − t)δi)! = tA−1(1 − t)Xj∈A δj. 10 T. Hytonen and A. Naor To prove (4.3) observe that for every ε, δ ∈ {−1, 1}n, (1 − t)Gt(δ)(ε) (4.4) (4.5) tB(1 − t)n−BWArB(δ) − tng(ε) i nYi=1(cid:16)t + (1 − t)δ1A(i) = XA⊆{1,...,n}bg(A)WA(ε) = XA⊆{1,...,n}bg(A)WA(ε) XB⊆{1,...,n} = XB({1,...,n} = XB({1,...,n} (cid:17) − tng(ε) tB(1 − t)n−B XA⊆{1,...,n}bg(A)WA∩B(ε)WArB(εδ) gB(ε, δ) def= gXj∈B εjej + Xj∈{1,...,n}rB εjδjej . tB(1 − t)n−BgB(ε, δ), where in (4.4) we use (4.1) and in (4.5) for every B ⊆ {1, . . . , n} we set Proof of Theorem 2.2. Observe that for every δ ∈ {−1, 1}n we have tB(1 − t)n−B = 1 − tn 1 − t . Since gB is equidistributed with g, it follows from (4.5) that (4.2) 0 1 0 (4.6) A6=∅ A6=∅ 1 6 nXj=1 kgkLq(Z) kGtkLq(Lq(Z)) It follows that if we set 1 − t XB({1,...,n} AXj∈A δj∆−1∂jg = XA⊆{1,...,n} δjbg(A)WA (cid:18)Z 1 tA−1dt(cid:19)Xj∈A = XA⊆{1,...,n} = Rad(cid:18)Z 1 Gt(δ)dt(cid:19) . Gtdt − Rad(cid:18)Z 1 Φ def= Z 1 dµ(δ) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) δj∆−1∂jg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 Gtdt(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lq(Lq(Z)) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 6 (cid:18)Z 1 Z 1 k=0R 1 It remains to note thatR 1 0 tkdt =Pn 1−t dt =Pn−1 then Φ ∈ Rad⊥ Gtdt(cid:19) . Lq(Z) and Φ(δ) + 1−tn 0 0 0 q 1/q (4.6) 0 Lq(Z) (4.3) (4.7) δjbg(A)WA dt(cid:19)kgkLq(Z). 0 1 − tn 1 − t 1 k . k=1 Pisier's inequality revisited 11 5 Proof of Theorem 2.3 For t ∈ (0, 1) define a linear operator Vt : Lq(Z) → Lq(Lq(Z)) by (5.1) Vt(g)(δ) def= Gt(δ) −cGt(∅) (4.1) = 1 1 − t XA⊆{1,...,n}bg(A)WA Yi∈A (t + (1 − t)δi) − tA! . Lemma 5.1. Let H be a Hilbert space. Then for every t ∈ (0, 1), (5.2) kVtkL2(H)→L2(L2(H))6 √1 − t Proof. Observe that for every A ⊆ {1, . . . , n} we have √1 − t2 6 . 1 1 (5.3) Z{−1,1}n Yi∈A (t + (1 − t)δi) − tA!2 = XB(A dµ(δ) t2B(1 − t)2(A−B) =(cid:0)t2 + (1 − t)2(cid:1)A − t2A. It follows from (5.1), (5.3), and the orthogonality of {WA}A⊆{1,...,n}, that (5.4) kVtkL2(H)→L2(L2(H)) = max a∈{1,...,n}p(t2 + (1 − t)2)a − t2a 1 − t . Now, for every a ∈ {1, . . . , n} and t ∈ (0, 1) we have (5.5) (cid:0)t2 + (1 − t)2(cid:1)a − t2a = (1 − t)2 6 (1 − t)2 a−1Xk=0(cid:0)t2 + (1 − t)2(cid:1)a−1−k t2k a−1Xk=0 t2k = (1 − t)2 1 − t2a 1 − t2 6 1 − t 1 + t , where in the first inequality of (5.5) we used the estimate t2 + (1 − t)2 6 1, which holds for every t ∈ [0, 1]. The desired estimate (5.2) now follows from a substitution of (5.5) into (5.4). Lemma 5.2. Let H be a Hilbert space and let W be a Banach space. Fix θ ∈ (0, 1) and q ∈ (1,∞). Set Z = [H, W ]θ. Then for every t ∈ (0, 1) we have (5.6) kVtkLq(Z)→Lq(Lq(Z)) 6 2 (1 − t)1−(1−θ) min{1/q,1−1/q} . 12 T. Hytonen and A. Naor Proof. For every r ∈ [1,∞] we have kVt(g)kLr(Lr(W )) Consequently, (5.1) = (cid:13)(cid:13)(cid:13)Gt −cGt(∅)(cid:13)(cid:13)(cid:13)Lr(Lr(W )) 6 2kGtkLr(Lr(W )) (4.3) 6 2kgkLr(W ) 1 − t . (5.7) ∀ r ∈ [1,∞], kVtkLr(W )→Lr(Lr(W )) 6 2 1 − t . If q ∈ [2,∞) then we interpolate (see [3]) between (5.2) and (5.7) with W = H and r = ∞. If q ∈ (1, 2] then we interpolate between (5.2) and (5.7) with W = H and r = 1. The norm bound thus obtained implies the estimate (5.8) ∀ q ∈ (1,∞), kVtkLq(H)→Lq(Lq(H)) 6 2 (1 − t)max{1/q,1−1/q} . Finally, interpolation between (5.8) and (5.7) with r = q gives the desired norm bound (5.6). Proof of Theorem 2.3. By (5.1) we have Rad(Vt(g)) = Rad(Gt). There- fore, analogously to (4.7), if we set Ψ def= Z 1 then Ψ ∈ Rad⊥ 0 Vt(g)dt − Rad(cid:18)Z 1 0 Gtdt(cid:19) =Z 1 0 Vt(g)dt − Rad(cid:18)Z 1 0 Vt(g)dt(cid:19) , Lq(Z) and by (4.6) for every δ ∈ {−1, 1}n we have (5.9) Ψ(δ) + Vt(g)(δ)dt. Hence, Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 Ψ(δ) + (5.9)∧(5.6) 0 δj∆−1∂jg =Z 1 nXj=1 dµ(δ) δj∆−1∂jg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Z 1 2kgkLq(Z) Lq(Z) 0 q 1/q 6 = (1 − t)1−(1−θ) min{1/q,1−1/q} dt 2kgkLq(Z) . (1 − θ) min{1/q, 1 − 1/q} This is precisely the assertion of Theorem 2.3. Pisier's inequality revisited 13 6 Enflo type in uniformly smooth Banach spaces A Banach space X has Rademacher type p ∈ [1, 2] (see e.g. [21]) if there exists TR ∈ (0,∞) such that for all n ∈ N and all x1, . . . , xn ∈ X, nXj=1 p εjxj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dµ(ε) 6 T p R nXj=1 kxjkp. (6.1) (6.2) Z{−1,1}n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Z{−1,1}n X has Enflo type p (see [9, 6, 28, 25]) if there exists TE ∈ (0,∞) such that for all n ∈ N and all f : {−1, 1}n → X, kf (ε) − f (−ε)kp 2p dµ(ε) 6 T p E nXj=1 k∂jfkp Lp(X). By considering the function f (ε) = Pn j=1 εjxj one sees that (6.1) is a special case of (6.2). It is a long-standing open problem [9] whether, conversely, (6.1) implies (6.2). A crucial feature of (6.2) is that it is a purely metric condition (thus one can define when a metric space has Enflo type p), while (6.1) is a linear condition. See [22] for a purely metric condition (which is more complicated than, but inspired by, Enflo type) that is known to be equivalent to Rademacher type. Observe that if (6.1) holds then it follows from (1.4) that for every f1, . . . , fn : {−1, 1}n → X, (6.3) nXj=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∆−1∂jfj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(X) 6 TRPn p (X) nXj=1 Lp(X)!1/p kfjkp . The special case fj = ∂jf shows that (6.3) implies (6.2) with TE 6 TRPn p (X). For this reason it is worthwhile to investigate (6.3) on its own right. Let Qn p (X) be the infimum over those Q ∈ (0,∞) such that every f1, . . . , fn : {−1, 1}n → X satisfy (6.4) We also set (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 ∆−1∂jfj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(X) 6 Q nXj=1 Lp(X)!1/p kfjkp . Qp(X) def= sup Qn p (X). n∈N 14 T. Hytonen and A. Naor By duality, Qn every g ∈ Lq(X ∗) satisfies p (X) equals the infimum over those Q ∈ (0,∞) for which (6.5) nXj=1(cid:13)(cid:13)∆−1∂jg(cid:13)(cid:13)q Lq(X ∗)!1/q 6 QkgkLq(X ∗). Letting SX = {x ∈ X : kxk = 1} denote the unit sphere of X, recall that the modulus of uniform convexity of X is defined for ε ∈ [0, 2] as δX(ε) = inf(cid:26)1 − kx + yk 2 : x, y ∈ SX , kx − yk = ε(cid:27) . The modulus of uniform smoothness of X is defined for τ ∈ (0,∞) as ρX (τ ) def= sup(cid:26)kx + τ yk+kx − τ yk 2 − 1 : x, y ∈ SX(cid:27) . These moduli relate to each other via the following classical duality formula of Lindenstrauss [18]. (6.6) δX ∗(ε) = supnτ ε 2 − ρX (τ ) : τ ∈ [0, 1]o . Theorem 6.1. For every K, p ∈ (1,∞) there exists C(K, p) ∈ (0,∞) such that if X is a Banach space that satisfies ρX (τ ) 6 Kτ p for all τ ∈ (0,∞), then Qp(X) 6 C(K, p). Proof. We shall use here the notation introduced in the proof of Theorem 2.1 (Section 3). It follows from (6.6) that δX ∗(ε)&K,pεq for every ε ∈ [0, 2] (here, and it what follows, the notation .K,p suppresses constant factors that may depend only on K and p). Hence, for g ∈ Lq(X ∗) and σ ∈ Sn, k=0, as defined in (3.1), is an X ∗-valued martingale, it follows since {gσ from Pisier's martingale inequality [26] that k}n (6.7) nXk=1 Lq(X ∗)!1/q k−1kq kgσ k − gσ .K,p kgkLq(X ∗). By reindexing (6.7) with k = σ(j), averaging over σ ∈ Sn, and using the convexity of the norm, we obtain the estimate (6.8)  nXj=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 n! Xσ∈Sn(cid:0)gσ σ(j) − gσ σ(j)−1(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) q Lq(X ∗) 1/q  .K,p kgkLq(X ∗). (6.9) 1 n! Xσ∈Sn(cid:0)gσ σ(j) − gσ = XA⊆{1,...,n} j∈A σ(j)−1(cid:1) = 1 n! Xσ∈Sn X∅(A⊆{1,...,n} max σ(A)=σ(j)bg(A)WA {σ ∈ Sn : max σ(A) = σ(j)} n! bg(A)WA = ∆−1∂jg. Pisier's inequality revisited 15 Arguing as in (3.4), for every j ∈ {1, . . . , n} we have the identity Consequently, (6.8) combined with (6.9) imply that (6.5) holds true with Q .K,p 1. This concludes the proof of Theorem 6.1. Remark 6.1. It follows from [17, Sec. 6] that a Banach space X satisfying the assumption of Theorem 6.1 has Enflo type p. Theorem 6.1 can be viewed as a generalization of this fact to yield the inequality (6.4). In [24] it was shown that any Banach space satisfying the assumption of Theorem 6.1 actually has K. Ball's Markov type p property [2], a property which is a useful strengthening of Enflo type p. Acknowledgements T. H. is supported in part by the European Union through the ERC Starting Grant "Analytic probabilistic methods for borderline singular integrals", and by the Academy of Finland through projects 130166 and 133264. A. N. is supported in part by NSF grant CCF-0832795, BSF grant 2010021, the Packard Foundation and the Simons Foundation. Part of this work was completed while A. N. was visiting Universit´e Pierre et Marie Curie, Paris, France. References [1] D. J. Aldous. Unconditional bases and martingales in Lp(F ). Math. Proc. Cambridge Philos. Soc., 85(1):117–123, 1979. [2] K. Ball. Markov chains, Riesz transforms and Lipschitz maps. Geom. Funct. Anal., 2(2):137–172, 1992. [3] J. Bergh and J. Lofstrom. Interpolation spaces. An introduction. Springer-Verlag, Berlin, 1976. Grundlehren der Mathematischen Wis- senschaften, No. 223. [4] J. Bourgain. Some remarks on Banach spaces in which martingale dif- ference sequences are unconditional. Ark. Mat., 21(2):163–168, 1983. 16 T. Hytonen and A. Naor [5] J. Bourgain. On martingales transforms in finite-dimensional lattices with an appendix on the K-convexity constant. Math. Nachr., 119:41– 53, 1984. [6] J. Bourgain, V. Milman, and H. Wolfson. On type of metric spaces. Trans. Amer. Math. Soc., 294(1):295–317, 1986. [7] D. L. Burkholder. Martingales and Fourier analysis in Banach spaces. In Probability and analysis (Varenna, 1985), volume 1206 of Lecture Notes in Math., pages 61–108. Springer, Berlin, 1986. [8] M. Cwikel and S. Reisner. Interpolation of uniformly convex Banach spaces. Proc. Amer. Math. Soc., 84(4):555–559, 1982. [9] P. Enflo. Uniform homeomorphisms between Banach spaces. In S´eminaire Maurey-Schwartz (1975–1976), Espaces, Lp, applications radonifiantes et g´eom´etrie des espaces de Banach, Exp. No. 18, page 7. Centre Math., ´Ecole Polytech., Palaiseau, 1976. [10] D. J. H. Garling. Random martingale transform inequalities. In Prob- ability in Banach spaces 6 (Sandbjerg, 1986), volume 20 of Progr. Probab., pages 101–119. Birkhauser Boston, Boston, MA, 1990. [11] S. Geiss. A counterexample concerning the relation between de- coupling constants and UMD-constants. Trans. Amer. Math. Soc., 351(4):1355–1375, 1999. [12] O. Giladi and A. Naor. Improved bounds in the scaled Enflo type inequality for Banach spaces. Extracta Math., 25(2):151–164, 2010. [13] P. Hitczenko. Domination inequality for martingale transforms of a Rademacher sequence. Israel J. Math., 84(1-2):161–178, 1993. [14] R. C. James. A nonreflexive Banach space that is uniformly nonoc- tahedral. Israel J. Math., 18:145–155, 1974. [15] R. C. James. Nonreflexive spaces of type 2. Israel J. Math., 30(1- 2):1–13, 1978. [16] R. C. James and J. Lindenstrauss. The octahedral problem for Banach spaces. In Proceedings of the Seminar on Random Series, Convex Sets and Geometry of Banach Spaces (Mat. Inst., Aarhus Univ., Aarhus, 1974; dedicated to the memory of E. Asplund), pages 100–120. Various Publ. Ser., No. 24, Aarhus Univ., Aarhus, 1975. Mat. Inst. Pisier's inequality revisited 17 [17] S. Khot and A. Naor. Nonembeddability theorems via Fourier anal- ysis. Math. Ann., 334(4):821–852, 2006. [18] J. Lindenstrauss. On the modulus of smoothness and divergent series in Banach spaces. Michigan Math. J., 10:241–252, 1963. [19] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. II, vol- ume 97 of Ergebnisse der Mathematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas]. Springer-Verlag, Berlin, 1979. Function spaces. [20] B. Maurey. Syst`eme de Haar. In S´eminaire Maurey-Schwartz 1974– 1975: Espaces Lp, applications radonifiantes et g´eom´etrie des espaces de Banach, Exp. Nos. I et II, pages 26 pp. (erratum, p. 1). Centre Math., ´Ecole Polytech., Paris, 1975. [21] B. Maurey. Type, cotype and K-convexity. In Handbook of the ge- ometry of Banach spaces, Vol. 2, pages 1299–1332. North-Holland, Amsterdam, 2003. [22] M. Mendel and A. Naor. Scaled Enflo type is equivalent to Rademacher type. Bull. Lond. Math. Soc., 39(3):493–498, 2007. [23] A. Naor. An introduction to the Ribe program. Jpn. J. Math., 7(2):167–233, 2012. [24] A. Naor, Y. Peres, O. Schramm, and S. Sheffield. Markov chains in smooth Banach spaces and Gromov-hyperbolic metric spaces. Duke Math. J., 134(1):165–197, 2006. [25] A. Naor and G. Schechtman. Remarks on non linear type and Pisier's inequality. J. Reine Angew. Math., 552:213–236, 2002. [26] G. Pisier. Un exemple concernant la super-r´eflexivit´e. In S´eminaire Maurey-Schwartz 1974–1975: Espaces Lp applications radonifiantes et g´eom´etrie des espaces de Banach, Annexe No. 2, page 12. Centre Math. ´Ecole Polytech., Paris, 1975. [27] G. Pisier. La m´ethode d'interpolation complexe: applications aux treillis de Banach. In S´eminaire d'Analyse Fonctionnelle (1978–1979), pages Exp. No. 17, 18. ´Ecole Polytech., Palaiseau, 1979. 18 T. Hytonen and A. Naor [28] G. Pisier. Probabilistic methods in the geometry of Banach spaces. In Probability and analysis (Varenna, 1985), volume 1206 of Lecture Notes in Math., pages 167–241. Springer, Berlin, 1986. [29] G. Pisier and Q. H. Xu. Random series in the real interpolation spaces between the spaces vp. In Geometrical aspects of functional analysis (1985/86), volume 1267 of Lecture Notes in Math., pages 185–209. Springer, Berlin, 1987. [30] M. Talagrand. Isoperimetry, logarithmic Sobolev inequalities on the discrete cube, and Margulis' graph connectivity theorem. Geom. Funct. Anal., 3(3):295–314, 1993. [31] R. Wagner. Notes on an inequality by Pisier for functions on the In Geometric aspects of functional analysis, volume discrete cube. 1745 of Lecture Notes in Math., pages 263–268. Springer, Berlin, 2000.
1902.08372
1
1902
2019-02-22T06:16:35
Integration with respect to deficient topological measures on locally compact spaces
[ "math.FA" ]
Topological measures and deficient topological measures generalize Borel measures and correspond to certain non-linear functionals. We study integration with respect to deficient topological measures on locally compact spaces. Such an integration over sets yields a new deficient topological measure if we integrate a nonnegative vanishing at infinity function; and it produces a signed deficient topological measure if we use a continuous function on a compact space. We present many properties of these resulting deficient topological measures and of signed deficient topological measures. In particular, they are absolutely continuous with respect to the original deficient topological measure and Lipschitz continuous. Deficient topological measures obtained by integration over sets can also be obtained from non-linear functionals. We show that for a deficient topological measure $ \mu$ that assumes finitely many values, there is a function $ f $ such that $\int_X f \, d \mu = 0$, but $\int_X (-f )\, d \mu \neq 0$. We present different criteria for $\int_X f \, d \mu = 0$. We also prove some convergence results, including a Monotone convergence theorem.
math.FA
math
INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES S. V. BUTLER, UCSB ABSTRACT. Topological measures and deficient topological measures general- ize Borel measures and correspond to certain non-linear functionals. We study integration with respect to deficient topological measures on locally compact spaces. Such an integration over sets yields a new deficient topological measure if we integrate a nonnegative vanishing at infinity function; and it produces a signed deficient topological measure if we use a continuous function on a com- pact space. We present many properties of these resulting deficient topological measures and of signed deficient topological measures. In particular, they are absolutely continuous with respect to the original deficient topological measure and Lipschitz continuous. Deficient topological measures obtained by integra- tion over sets can also be obtained from non-linear functionals. We show that for a deficient topological measure µ that assumes finitely many values, there is a function f such that RX f dµ = 0, but RX (−f ) dµ 6= 0. We present different criteria for RX f dµ = 0. We also prove some convergence results, including a Monotone convergence theorem. 1. INTRODUCTION Topological measures (initially called quasi-measures) were introduced by J. F. Aarnes in [1], [2], and [3]. These generalizations of measures are defined on open and closed subsets of a topological space. Despite the lack of an algebraic structure on the domain, absence of subadditivity, and unavailability of many stan- dard techniques of measure theory and functional analysis, topological measures still possess many features of regular Borel measures. Via integration, topologi- cal measures correspond to functionals that are not linear, but are linear on singly Date: February 21, 2019. 2010 Mathematics Subject Classification. 28A25, 28C05, 28C15, 46T99, 46F99. Key words and phrases. deficient topological measure, signed deficient topological measure, non- linear functionals, quasi-integration, absolute continuity, Lipschitz continuous functional. 1 2 S. V. BUTLER, UCSB generated subalgebras. Topological measures and corresponding quasi-linear func- tionals are connected to the problem of linearity of the expectational functional on the algebra of observables in quantum mechanics. Initial papers by Aarnes were followed by many papers devoted to the subject. Applications of topological mea- sures and corresponding non-linear functionals to symplectic topology have been studied in numerous papers beginning with [13] (which has been cited over 100 times), and in a monograph [15]. Topological measures are a subclass of deficient topological measures, which also correspond via integration to certain non-linear functionals. (See, for example, [14], [17], [18], [7], and [9] for more information). Deficient topological measures are not only interesting by themselves, but also provide an essential framework for studying topological measures and various non-linear functionals. This provided a motivation for our study of integration with respect to deficient topological mea- sures on locally compact spaces, especially given the fact that the vast majority of results so far has been proven for compact spaces (which has impeded the develop- ment of the area and its applications). We demonstrate that integration with respect to a deficient topological measure sometimes gives the same results as integration with respect to a measure would give; in other instances, the results are very dif- ferent. Some of our results are new, and some are generalizations of known results about deficient topological measure on compact spaces to signed deficient topolog- ical measures on compact spaces and /or deficient topological measures on locally compact spaces. (See [10] and [11] for more information about signed deficient topological measures.) We begin (Section 2) with the study of integration with respect to a deficient topological measure over a set. It is done by using "restricted" deficient topolog- ical measures. Integration over sets yields a new deficient topological measure if we integrate a nonnegative vanishing at infinity function; and it produces a signed deficient topological measure if we use a continuous function on a compact space. We present many properties of these resulting deficient topological measures and of signed deficient topological measures. In particular, these resulting (signed) deficient topological measures are absolutely continuous with respect to the orig- inal deficient topological measure, and they are Lipschitz continuous. In Section INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES3 3 we show that the deficient topological measures obtained in Section 2 from fi- nite deficient topological measures by integrating over sets can also be obtained from non-linear functionals. We present more properties of such deficient topo- logical measures. In Section 4 we show that for a deficient topological measure µ that assumes finitely many values, there is f ≥ 0 such that RX f dµ = 0, but RX(−f ) dµ 6= 0. Using integration over zero and cozero sets, we present different criteria for RX f dµ = 0. We conclude the paper (Section 5) with some conver- gence results, including a Monotone convergence theorem. In this paper X is a locally compact, connected space. By C(X) we denote the set of all real-valued continuous functions on X with the uniform norm, by Cb(X) the set of bounded continuous functions on X, by C0(X) the set of continuous functions on X vanishing at infinity, and by Cc(X) the set of continuous functions with compact support. By C + c (X) we denote the collection of all nonnegative functions from C0(X) and Cc(X), respec- 0 (X) and C + tively. Z(f ) and Coz(f ) stand for the zero and the cozero sets of the function f , i.e. Z(f ) = {x ∈ X : f (x) = 0}, Coz(f ) = X \ Z(f ). When we consider maps into [−∞, ∞] we assume that any such map attains at most one of ∞, −∞, and is not identically ∞ or −∞. We denote by E the closure of a set E, and byF a union of disjoint sets. We de- note by 1 the constant function 1(x) = 1, by id the identity function id(x) = x, and by 1K the characteristic function of a set K. By supp f we mean {x : f (x) 6= 0}. Several collections of sets are used often. They include: O(X); C (X); and K (X) -- the collection of open subsets of X; the collection of closed subsets of X; and the collection of compact subsets of X, respectively. Definition 1. Let X be a topological space and ν be a set function on a family of subsets of X that contains O(X) ∪ C (X). We say that • ν is compact-finite if ν(K) < ∞ for any K ∈ K (X); • ν is simple if it only assumes values 0 and 1; • a nonnegative set-function ν is finite if ν(X) < ∞. Definition 2. A Radon measure m on X is a Borel measure that is finite on all compact sets, outer regular on all Borel sets, and inner regular on all open sets, i.e. 4 S. V. BUTLER, UCSB m(E) = inf{m(U ) : E ⊆ U, U open } for every Borel set E, and m(U ) = sup{m(K) : K ⊆ U, K compact } for every open set U . Recall the following fact (see, for example, [12, Chapter XI, 6.2]): Lemma 3. Let K ⊆ U, K ∈ K (X), U ∈ O(X) in a locally compact space X. Then there exists a set V ∈ O(X) with compact closure such that K ⊆ V ⊆ V ⊆ U. Definition 4. A deficient topological measure on a locally compact space X is a set function ν : C (X) ∪ O(X) −→ [0, ∞] which is finitely additive on compact sets, inner compact regular, and outer regular, i.e. : (DTM1) if C ∩ K = ∅, C, K ∈ K (X) then ν(C ⊔ K) = ν(C) + ν(K); (DTM2) ν(U ) = sup{ν(C) : C ⊆ U, C ∈ K (X)} for U ∈ O(X); (DTM3) ν(F ) = inf{ν(U ) : F ⊆ U, U ∈ O(X)} for F ∈ C (X). For a closed set F , ν(F ) = ∞ iff ν(U ) = ∞ for every open set U containing F . Definition 5. A topological measure on X is a set function µ : C (X) ∪ O(X) −→ [0, ∞] satisfying the following conditions: (TM1) if A, B, A ⊔ B ∈ K (X) ∪ O(X) then µ(A ⊔ B) = µ(A) + µ(B); (TM2) µ(U ) = sup{µ(K) : K ∈ K (X), K ⊆ U } for U ∈ O(X); (TM3) µ(F ) = inf{µ(U ) : U ∈ O(X), F ⊆ U } for F ∈ C (X). The following two theorems from [7, Section 4] give criteria for a deficient topological measure to be a topological measure or a measure. Theorem 6. (I) Let X be compact, and ν a deficient topological measure. The following are equivalent: (a) ν is a topological measure (b) ν(X) = ν(C) + ν(X \ C), C ∈ C (X) (c) ν(X) ≤ ν(C) + ν(X \ C), C ∈ C (X) (II) Let X be locally compact, and ν a deficient topological measure. The following are equivalent: (a) ν is a topological measure (b) ν(U ) = ν(C) + ν(U \ C), C ∈ K (X), U ∈ O(X) INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES5 (c) ν(U ) ≤ ν(C) + ν(U \ C), C ∈ K (X), U ∈ O(X) Theorem 7. Let µ be a deficient topological measure on a locally compact space X. The following are equivalent: (a) If C, K are compact subsets of X, then µ(C ∪ K) ≤ µ(C) + µ(K). (b) If U, V are open subsets of X, then µ(U ∪ V ) ≤ µ(U ) + µ(V ). (c) µ admits a unique extension to an inner regular on open sets, outer regular Borel measure m on the Borel σ-algebra of subsets of X. m is a Radon measure iff µ is compact-finite. If µ is finite then m is a regular Borel measure. Remark 8. In [7, Section 3] we show that a deficient topological measure ν is τ -smooth on compact sets (i.e. if a net Kα ց K , where Kα, K ∈ K (X) then µ(Kα) → µ(K)), and also τ -smooth on open sets (i.e. if a net Uα ր U , where Uα, U ∈ O(X) then µ(Uα) → µ(U )). where At, A ∈ O(X) ∪ C (X), and at most one of the closed sets (if there are any) A deficient topological measure ν is also superadditive, i.e. if Ft∈T At ⊆ A, is not compact, then ν(A) ≥Pt∈T ν(At). If F ∈ C (X) and C ∈ K (X) are disjoint, then ν(F ) + ν(C) = ν(F ⊔ C). One may consult [7] for more properties of deficient topological measures on locally compact spaces. Definition 9. A signed deficient topological measure on a locally compact space X is a set function ν : C (X) ∪ O(X) −→ [−∞, ∞] that assumes at most one of ∞, −∞ and that is finitely additive on compact sets, inner compact regular on open sets, and outer regular on closed sets, i.e. (SDTM1) If C ∩ K = ∅, C, K ∈ K (X) then ν(C ⊔ K) = ν(C) + ν(K); (SDTM2) µ(U ) = lim{µ(K) : K ∈ K (X), K ⊆ U } for U ∈ O(X); (SDTM3) µ(F ) = lim{µ(U ) : U ∈ O(X), F ⊆ U } for F ∈ C (X). Remark 10. In condition (SDTM2) we mean the limit of the net ν(C) with the index set {C ∈ K (X) : C ⊆ U } ordered by inclusion. The limit exists and is equal to ν(U ). Condition (SDTM3) is interpreted in a similar way, with the index set being {U ∈ O(X) : U ⊇ C} ordered by reverse inclusion. 6 S. V. BUTLER, UCSB Remark 11. Since we consider set-functions that are not identically ∞ or −∞, we see that for a signed deficient topological measure ν(∅) = 0. If ν and µ are (signed) deficient topological measures that agree on K (X) (or on O(X)) then ν = µ; if ν ≤ µ on K (X) (or on O(X)) then ν ≤ µ. Remark 12. For a signed deficient topological measure ν we may define its total variation, a deficient topological measure ν, by ν(U ) = sup{ nXi=1 ν(Ki) : nGi=1 Ki ⊆ U, Ki ⊆ K (X), n ∈ N}; for an open set U , and for a closed subset F ⊆ X ν(F ) = inf{ν(U ) : F ⊆ U, U ∈ O(X)}. See [7, Sections 2,3] for detail. Definition 13. We define kνk = sup{ν(K) : K ∈ K (X)} for a signed deficient topological measure ν. If µ is a deficient topological measure then kµk = µ(X). Definition 14. A signed deficient topological measure ν is called proper if from m ≤ ν, where m is a Radon measure and ν is a total variation of ν, it follows that m = 0. Let µ be a deficient topological measure, and ν be a signed deficient topological measure. We say that ν is absolutely continuous with respect to µ (and we write ν ≪ µ) if µ(A) = 0 implies ν(A) = 0. Definition 15. A signed topological measure on a locally compact space X is a set function µ : O(X) ∪ C (X) −→ [−∞, ∞] that assumes at most one of ∞, −∞ and satisfies the following conditions: (STM1) if A, B, A ⊔ B ∈ K (X) ∪ O(X) then µ(A ⊔ B) = µ(A) + µ(B); (STM2) µ(U ) = lim{µ(K) : K ∈ K (X), K ⊆ U } for U ∈ O(X); (STM3) µ(F ) = lim{µ(U ) : U ∈ O(X), F ⊆ U } for F ∈ C (X). Remark 16. Definition 9 was first introduced in [10]. Any deficient topological measure (topological measure, signed topological measure) is a signed deficient topological measure. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES7 Remark 17. There is a correspondence between deficient topological measures and certain non-linear functionals, see [9, Section 8]. In particular, there is an order-preserving isomorphism between compact-finite topological measures on X and quasi-integrals on Cc(X), and µ is a measure iff the corresponding functional is linear (see Theorem 42 in Section 4 of [8] and Theorem 3.9 in [16] for the first version of the representation theorem.) We outline the correspondence. (I) Given a finite deficient topological measure µ on a locally compact space X and f ∈ Cb(X), define functions on R: R1(t) = R1,µ,f (t) = µ(f −1((t, ∞))), R2(t) = R2,µ,f (t) = µ(f −1([t, ∞))). Let r be the Lebesque-Stieltjes measure associated with −R1, a regular Borel measure on R. We define a functional on C0(X): (1) R(f ) =ZR id dr =Z b a id dr =Z b a R1(t)dt + aµ(X) =Z b a R2(t)dt + aµ(X). where [a, b] is any interval containing f (X). If f is nonnegative with f (X) ⊆ [0, b] we have: R(f ) =Z b 0 id dr =Z b 0 R1(t)dt =Z b 0 R2(t)dt. We call the functional R a quasi-integral (with respect to a deficient topo- logical measure µ) and write: ZX f dµ = R(f ) = Rµ(f ) =ZR id dr. (2) (3) (II) Functional R is non-linear. By [9, Lemma 7.7, Lemma 7.10, Lemma 3.6, Lemma 7.12] we have: (a) R(f ) is positive-homogeneous, i.e. R(cf ) = cR(f ) for c ≥ 0, and R(0) = 0. (b) if g h = 0, where g, h ≥ 0 or g ≥ 0, h ≤ 0, then R(gh) = R(g) + R(h). (c) R is monotone, i.e. if f ≤ g then R(f ) ≤ R(g). (d) µ(X) · inf x∈X f (x) ≤ R(f ) ≤ µ(X) · supx∈X f (x) for any f ∈ C0(X). 8 S. V. BUTLER, UCSB (III) A functional ρ with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called a d-functional if on nonnegative functions it is positive-homogeneous, monotone, and orthogonally additive, i.e. for f, g ∈ D(ρ) (the domain of ρ) we have: (d1) f ≥ 0, a > 0 =⇒ ρ(af ) = aρ(f ); (d2) 0 ≤ g ≤ f =⇒ ρ(g) ≤ ρ(f ); (d3) f · g = 0, f, g ≥ 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). Let ρ be a d-functional with C + we may take functional R on C + c (X) ⊆ D(ρ) ⊆ Cb(X). In particular, 0 (X). The corresponding deficient topo- logical measure µ = µρ is given as follows: If U is open, µρ(U ) = sup{ρ(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U }, if F is closed, µρ(F ) = inf{µρ(U ) : F ⊆ U, U ∈ O(X)}. If K is compact, µρ(K) = inf{ρ(g) : g ∈ Cc(X), g ≥ 1K } = inf{ρ(g) : g ∈ Cc(X), 1K ≤ g ≤ 1}. (See Definition 33 and Lemma 35 in Section 4 of [8].) If given a finite deficient topological measure µ, we obtain R, and then µR, then µ = µR. Definition 18. Let X be locally compact. A functional ρ on C0(X) is Lipschitz continuous if for every compact K ⊆ X there exists a number NK such that ρ(f ) − ρ(g) ≤ NKkf − gk for all f, g ∈ Cc(X) with support in K. The proof of the following theorem can be found, for example, in [5], §7.2.6. Theorem 19. Let l be a regular τ -smooth measure on a topological space X and let {fα} be an increasing net of nonnegative lower semicontinuous functions such that the function f = limα fα is bounded. Then α ZX lim fα(x) l(dx) =ZX f (x) l(dx). Remark 20. It is easy to see that a left-continuous nondecreasing function (that does not assume −∞) or right-continuous nonincreasing function (that does not assume −∞) is lower semicontinuous. Corollary 21. Let l be a regular τ -smooth measure on a topological space X. Let {gα} be a decreasing net of nonnegative nonincreasing left-continuous functions INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES9 converging to g. Suppose one of the functions {gα} is bounded. Then α Z b lim a gα(x) l(dx) =Z b a g(x) l(dx). Proof. Suppose gγ is bounded above by M > 0. Then gα ≤ M for all α ≥ γ. Consider functions fα = M − gα and f = M − g. By Remark 20, functions fα are lower semi-continuous. Apply Theorem 19 to fα. (cid:3) 2. INTEGRALS OVER A SET VIA "RESTRICTED" DEFICIENT TOPOLOGICAL MEASURES We would like to begin with integration with respect to a deficient topological measure over a set. Integration of a continuous bounded function with respect to a deficient topological measure is described in part (I) of Remark 17, but it will not work for a function restricted to a set, as such a function need not be continuous anymore. Another approach to obtain integration over a set would be to integrate the function with respect to a deficient measure that is restricted to a set. But if µ is a deficient topological measure on X, one can not obtain a deficient topological measure on A ∈ O(X) ∪ C (X) by simply restricting µ to A, i.e., considering µ(A ∩ B), B ∈ O(X) ∪ C (X). One simple reason for this is that the intersection of two arbitrary sets from O(X) ∪ C (X) does not in general belong to O(X) ∪ C (X). However, there is still a way to obtain new deficient topological measures by "restriction". The next two results are from [7, Section 5]. Theorem 22. Let µ be a deficient topological measure on a locally compact space X, V ∈ O(X). Define a set function µV on O(X) ∪ C (X) by letting µV (U ) = µ(U ∩ V ), U ∈ O(X), µV (F ) = inf{µV (U ) : F ⊆ U, U ∈ O(X)}, F ∈ C (X). Then µV is a deficient topological measure on X. Theorem 23. Let X be locally compact, and let F ⊆ C (X). There exists a deficient topological measure µF on X such that µF (K) = µ(F ∩ K) for K ∈ K (X), µF (U ) = sup{µF (K) : K ⊆ U, K ∈ K (X)} for U ∈ O(X). 10 S. V. BUTLER, UCSB If F ∈ K (X) then µF (C) = µ(F ∩ C) for every C ∈ C (X). The next theorem states some properties of "restricted" deficient topological measures µV and µF , given by Theorem 22 and Theorem 23. Theorem 24. Let X be locally compact, V ∈ O(X), F ∈ C (X), and µ, ν be defi- cient topological measures on X. Let µV and µF be deficient topological measures given by Theorem 22 and Theorem 23. Then (v1) µV (U ) = sup{µK(U ) : K ⊆ U, K ∈ K (X)} for every U ∈ O(X). (v2) µF (C) = inf{µV (C) : F ⊆ V, V ∈ O(X)} for every C ∈ K (X). If F is compact then the equality holds for every C ∈ C (X). (v3) If A, B are disjoint compact sets, or disjoint open sets, then µA⊔B = µA + µB. (v4) If µ ≤ ν then µA ≤ νA for any set A ∈ O(X) ∪ C (X). (v5) If A, B ⊆ O(X) ∪ C (X), A ⊆ B then µA ≤ µB. (v6) If a, b ≥ 0 then (aµ + bν)A = aµA + bνA for any set A ∈ O(X) ∪ C (X). (v7) If A ∈ O(X)∪ K (X) then µA(X) = µ(A). If F ∈ C (X) then µF (X) ≤ µ(F ). Proof. (v1) Assume first that µV (U ) = µ(U ∩ V ) < ∞. Given ǫ > 0, let K ∈ K (X) be such that K ⊆ U ∩ V, µ(U ∩ V ) − µ(K) < ǫ. Note that for any C ⊆ U we have µK (C) = µ(K∩C) ≤ µ(U ∩V ) = µV (U ), which gives µK(U ) ≤ µV (U ). Since µV (U ) − µK(U ) ≤ µV (U ) − µK(K) = µ(U ∩ V ) − µ(K) < ǫ, we have: µV (U ) = sup{µK(U ) : K ⊆ U, K ∈ K (X)}. Now assume that µV (U ) = µ(U ∩ V ) = ∞. For each n ∈ N choose Kn ∈ K (X) such that Kn ⊆ U ∩V, µ(Kn) ≥ n. For any compact C such that Kn ⊆ C ⊆ U we have µKn(C) = µ(Kn ∩ C) = µ(Kn) ≥ n, and we see that µKn(U ) ≥ n. Then sup{µK (U ) : K ⊆ U, K ∈ K (X)} = ∞ = µV (U ). (v2) Follows from Theorem 23 and the following result from [7, Section 5]. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES11 Lemma 25. Let X be locally compact, µ be a deficient topological mea- sure on X, and F ∈ C (X). Then for any C ∈ K (X) (4) µ(F ∩ C) = inf{µV (C) : F ⊆ V, V ∈ O(X)}, where µV is a deficient topological measure on X from Theorem 22. If F ∈ K (X) then (4) holds for any C ∈ C (X). (v3) By Remark 11 it is enough to check the equality on compact sets (or open sets). The argument is easy and left to the reader. (v4) Take A ∈ C (X) (respectively, A ∈ O(X)). Then µA(K) ≤ νA(K) for every K ∈ K (X) (respectively, µA(U ) ≤ νA(U ) for every U ∈ O(X)). The statement then follows from Remark 11. (v5) Arguing as in part (v4) we see that µA ≤ µB if A, B ∈ C (X) or if A, B ∈ O(X). If F ⊆ U, F ⊆ C (X), U ∈ O(X), then by part (v2) µF (K) ≤ µU (K) for every K ∈ K (X). By Remark 11 µF ≤ µU . Consider the last case: U ⊆ F, U ∈ O(X), F ∈ C (X). Let V ∈ O(X) be such that F ⊆ V . So U ⊆ F ⊆ V . Then µU (K) ≤ µV (K), and by part (v2) µU (K) ≤ µF (K) for every K ∈ K (X). By Remark 11 µU ≤ µF . (v6) The argument is similar to one for part (v4). (v7) Easy to see from Theorem 22 and Theorem 23. (cid:3) Definition 26. Suppose X is locally compact, µ is a deficient topological measure on X, and A ∈ O(X) ∪ C (X) is such that µ(A) < ∞. In particular, for a finite deficient topological measure µ on X, we take A ∈ O(X) ∪ C (X). Let µA be a deficient topological measure given by Theorem 22 or Theorem 23. We define the the quasi-integral (with respect to a deficient topological measure µ) of a function g ∈ Cb(X) over a set A with µ(A) < ∞ to be the functional RµA(g) as in formula (3) (where we use a deficient topological measure µA) and we write: (5) ZA g dµ =ZX g dµA = RµA(g). Remark 27. If A = X we have the usual Rµ(g). If µ is a measure, we obtain the standard integralRA g dµ over a set A . To evaluate RµA(g) we may use formulas 12 S. V. BUTLER, UCSB (1), (2), and part (v7) of Theorem 24. For example, if f ∈ Cb(X), f (X) ⊆ [a, b], E ∈ O(X) ∪ K (X), and µ(E) < ∞, we may use: (6) ZE f dµ = aµ(E) +Z b a R1,µE ,f (t) dt = aµ(E) +Z b a R2,µE ,f (t) dt, where, for example, R1,µV ,f (t) = µV (f −1((t, ∞))) = ν(V ∩ f −1((t, ∞))) for E = V ∈ O(X), (7) R2,µC ,f (t) = µC(f −1([t, ∞))) = µ(C ∩ f −1([t, ∞))) for E = C ∈ K (X). If f ∈ C + 0 (X) and f (X) ⊆ [0, b] then for A ∈ O(X) ∪ C (X) with µ(A) < ∞ we may use: (8) ZA f dµ =Z b 0 R1,µA,f (t) dt =Z b 0 R2,µA,f (t) dt. Let a′ = supE f, b′ = inf E f . Since Ri,µE ,f (t) = µ(E) for t ∈ [a, a′), and Ri,µE ,f (t) = ∅ for t ∈ (b′, b] (where i = 1, 2), we may rewrite ( 6 ) as: (9) ZE f dµ = a′µ(E) +Z b′ a′ R1,µE ,f (t) dt = a′µ(E) +Z b′ a′ R2,µE ,f (t) dt Lemma 28. Let X be locally compact, µ be a deficient topological measure on X, and g ∈ Cb(X). We have: (i) If V ∈ O(X), µ(V ) < ∞ then ZV g dµ = limnZK g dµ : K ⊆ U, K ∈ K (X)o. (ii) Suppose g ∈ C(X) if X is compact and g ∈ C + 0 (X) if X is locally compact. If F ∈ C (X), µ(F ) < ∞ then ZF g dµ = limnZV g dµ : F ⊆ V, V ∈ O(X)o. Proof. Let g(X) ⊆ [a, b]. If g ∈ C + 0 (X) we take a = 0. (i) For any K ⊆ V, K ∈ K (X) the function R1,µK ,g(t) is right-continuous (see [9, Lemma 6.3]), nonincreasing, so by Remark 20 it is lower-semicontinuous. For each t ∈ [a, b] by part (v5) of Theorem 24 the net {R1,µK ,g(t) : K ⊆ INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES13 V, K ∈ K (X)} (ordered by inclusion) is nondecreasing, and by part (v1) of Theorem 24 lim K⊆V,K∈K (X) R1,µK ,g(t) = R1,µV ,g(t). Note that 0 ≤ R1,µV ,g(t) ≤ µ(V ). Apply formula (6) and Theorem 19 to obtain the statement. (ii) Since µ(F ) < ∞, there exists U ∈ O(X) such that F ⊆ U, µ(U ) < ∞, and then µ(V ) < ∞ for all open sets V such that F ⊆ V ⊆ U . For any V ∈ O(X) containing F the function R2,µV ,g is left-continuous on [a, b] (use [9, Lemma 6.3]), nonincreasing, and bounded. The assumptions as- sure that each set g−1([t, ∞)) is compact for t ∈ (a, b]. For t ∈ [a, b] by part (v5) of Theorem 24 the net {R2,µV ,g(t) : V ∈ O(X), F ⊆ V } (or- dered by reverse inclusion) is nonincreasing, and by part (v2) of Theorem 24 lim V ∈O(X),F ⊆V R2,µV ,g(t) = R2,µF ,g(t). Applying formula (6) and Corollary 21 we obtain the statement. (cid:3) Theorem 29. Let X be locally compact, µ be a compact-finite deficient topological measure on X. (a) Let g ∈ C + 0 (X). Then there exists a compact-finite deficient topological measure µg on X such that µg(A) =RA g dµ for every A ∈ O(X)∪ C (X) with µ(A) < ∞, and µg(K) ≤ kgkµ(K) for any K ∈ K (X). Also, if E ⊔ D = F , where E, D, F ∈ C (X) and µ(F ) < ∞, then µg(E) + µg(D) = µg(F ). (b) Let X be compact, g ∈ C(X). Then a set function νg on A ∈ O(X) ∪ C (X) given by νg(A) = RA g dµ is a signed deficient topological mea- sure with finite norm kνgk ≤ kgkµ(X). If g ≥ 0 then νg is a deficient topological measure, and νg = µg. Proof. (a) Consider a set function λ on K (X) given by λ(K) =RK g dµ. By part (v3) of Lemma 24 λ is finitely additive on K (X). We take µg to be the positive variation λ+ of λ, so µg is a deficient topological measure (see 14 S. V. BUTLER, UCSB [7, Section 3]). By definition of λ+ and Lemma 28 we see that µg(A) = RA g dµ for every A ∈ O(X) ∪ C (X) with µ(A) < ∞. For any K ∈ K (X) using part (II) of Remark 17 and part (v7) of Theorem 24 we have: (10) µg(K) = RµK (g) ≤ kgkµK (X) = kgkµ(K), and we see that µg is compact finite. Now suppose E ⊔ D = F , where E, D, F ∈ C (X) and µ(F ) < ∞. Let g(X) ⊆ [0, b]. For each t > 0 g−1([t, ∞)) ∩ F = (g−1([t, ∞)) ∩ E) ⊔ (g−1([t, ∞)) ∩ G), and all sets are compact. Since µ is finitely additive on compact sets, we see that Z b 0 R2,µF ,g(t) dt =Z b 0 R2,µE ,g(t) dt +Z b 0 R2,µD ,g(t) dt, which (since µ(F ), µ(E), µ(D) < ∞) means that µg(F ) = µg(E) + µg(D). (b) From part (v3) of Theorem 24 and Lemma 28 we see that νg is a signed de- ficient topological measure. As in (10), νg(K) ≤ kgkµ(K) ≤ kgkµ(X) for each K ∈ K (X), so by Definition 13 kνgk ≤ kgkµ(X). If g ≥ 0 then νg = µg on K (X), so by Remark 11 νg = µg. (cid:3) The next example shows that by Theorem 29 we can obtain a signed deficient topological measure. Example 30. In this example we shall use a "parliamentry" topological measure µ on X = [−4, 0] × [0, 4]: given an odd number of distinct points of a compact space X, we assign a closed or open set which is also solid (i.e. which is connected and has a connected complement) µ-value 1 if the set contains more than half of the points, and µ-value 0 otherwise. (For more information, see [2, Remark 6.1] or [4, Introduction, Example b].) Let µ be the simple topological measure based on points p = (−4, 0), q = (−4, 4), and s = (0, 4). Let g(x, y) = x, so g(X) = [−4, 0]. Let K ∈ K (X) be the closed triangle with vertices (−4, 0), (−4, 4), and (0, 0), so K is a closed solid set with µ(K) = 1. Note that for any t > −4 the INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES15 set K ∩ g−1([t, ∞)) is a closed solid set with µ(K ∩ g−1([t, ∞))) = 0, and so R2,µK ,g(t) = µ(K ∩ g−1([t, ∞))) = 0. By formula (6) we see that νg(K) = −4µ(K) = −4. Thus, νg is a signed deficient topological measure. Note also that νg(K) = −4 = inf K g · µ(K). Now we shall show that νg is not a signed topological measure. Let C ∈ K (X) be the closed triangle with vertices (−4, 0), (−2, 4), and (0, 0), and let U, V be the connected components of X \ C. Triangles U, V are open solid sets. Since µ(U ) = µ(V ) = 0, µ(X) = 1, the sets U ∩ g−1((t, ∞)), V ∩ g−1((t, ∞)), and g−1((t, ∞)) are open solid sets with at most one of points p, q, s for each t > 0, by formula (6) we calculate νg(U ) = νg(V ) = 0, νg(X) = −4. For each t > −4 the set C ∩ g−1([t, ∞)) is not solid, but its complement consists of two disjoint open solid sets which we call Ut and Vt, and µ(Ut) = 1, µ(Vt) = 0. Then R2,µC ,g(t) = µ(C ∩ g−1([t, ∞))) = µ(X) − µ(Ut) − µ(Vt) = 0 for each t > −4. Since µ(C) = 0, we obtain νg(C) = 0. Since νg(X) = −4, while νg(V ) = νg(W ) = νg(C) = 0, we see that νg is not a signed topological measure. Remark 31. If in Theorem 29 µ is a measure, then it is easy to see that νg is a signed measure, which is a measure if g ≥ 0. However, the next example shows that if µ is a topological measure, then µg may not be a a topological measure. Example 32. Let X = R2, and D be the disk of radius 2 centered at the ori- gin. Let g ∈ Cc(X) be the function with supp g = D, whose graph is the cone with the vertex (0, 2) and the base D. Thus, g(X) = [0, 2]. Let µ be the sim- ple topological measure based on points (0, 0), (1, 0) and (4, 0) as in [6, Exam- ple 2]. Since g−1((t, ∞)) is a bounded open solid set, by [6, Remark 4] we see µg(X) = RX g dµ = R 4 that R1,µ,g(t) = 1 for 0 < t < 1, and R1,µ,g(t) = 0 for 1 ≤ t ≤ 2. Then 0 R1,µ,g(t) dt = 1. Let compact K = {(x, 0) : 1 ≤ x ≤ 4} and V = X \ K. For each t ≥ 0 we see that K ∩ g−1([t, ∞)) is a compact solid set, and V ∩ g−1((t, ∞)) is a bounded open solid set. Note that R2,µK ,g(t) = µ(K ∩ g−1([t, ∞))) = 0, and R1,µV ,g(t) = 0 for each t > 0. Thus, µg(K) = µg(V ) = 0. It follows that µg is not a topological measure. When µ is a topological measure, µg may not be a topological measure even for a strictly positive function g. Let X = [1, 3] × [0, 1]. Let µ be the "parliamentry" 16 S. V. BUTLER, UCSB simple topological measure based on points p = (2, 0), q = (1, 1), and s = (3, 1). Let g(x, y) = x, so g > 0 and g(X) = [1, 3]. Let K be the triangle with vertices (1, 1), (1, 0), (3, 0) and let V = X \ K. So K ∈ K (X), V ∈ O(X), X = K ⊔ V , and µ(K) = 1, µ(V ) = 0, µ(X) = 1. Note that for t ∈ (1, 3] the set K ∩ g−1([t, ∞)) is a closed solid set containing at most one of the points p, q, s, and so R2,µK ,g(t) = µ(K ∩ g−1([t, ∞))) = 0. Similarly, R1,µV ,g(t) = 0 for t ∈ (1, 3]. Using formula (6) we see that µg(V ) = 0 and µg(K) = 1. By formula (1) we also have µg(X) = 2, since R2,µ,g(t) = 1 if t ∈ [1, 2], and R2,µ,g(t) = 0 if t > 2. Thus, µg can not be a topological measure. Theorem 33. Let X be locally compact, µ be a compact-finite deficient topological measure on X. Let µg (for g ∈ C + 0 (X)) and νg (for g ∈ C(X) and X compact) be as in Theorem 29. Then (b1) If a ≥ 0 then µag = aµg. If g = 0 then µg = 0. The same holds for νg. (b2) If g h = 0, where g, h ≥ 0 then µg+h = µg + µh. If g h = 0, g ≥ 0, h ≤ 0 then νg+h = νg + νh. (b3) If g ≤ h then µg ≤ µh (respectively, νg ≤ νh). (b4) If A ∈ O(X) ∪ C (X) and µ(A) < ∞ then for λg = µg or λg = νg (11) µ(A) · inf A g ≤ λg(A) ≤ µ(A) · sup A g. In particular, (12) µg(A) ≤ µ(A) · sup A g ≤ kgkµ(A), νg(A) ≤ µ(A) sup A g ≤ kgkµ(A). (b5) If A ∈ O(X) ∪ C (X), µ(A) < ∞ and g = h on A then µg(A) = µh(A) (respectively, νg(A) = νh(A)). (b6) (Lipschitz continuity) If K ∈ K (X), f, g ∈ C + c (X) and supp f, supp g ⊆ K then µg(K) − µf (K) = RµK (g) − RµK (f ) ≤ kf − gkµ(K). (b7) If A ∈ O(X) ∪ K (X) and g = 1 on A then µg(A) = µ(A) (respectively, νg(A) = µ(A)). (b8) For any open set U ⊆ X we have µ(U ) = sup{µg(K) : K ⊆ U, K ∈ K (X), g = 1 on K} (respectively, µ(U ) = lim{νg(K) : K ⊆ U, K ∈ K (X), g = 1 on K}). INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES17 (b9) If µ = cλ + dδ, where c, d ≥ 0 and λ, δ are compact-finite deficient topological measures on X, then µg = cλg+dδg. (Here λg, δg are deficient topological measures obtained from λ, δ by Theorem 29.) Respectively, νg = cλg + dδg, where λg, δg are signed deficient topological measures obtained from λ, δ by Theorem 29. (b10) If µ ≤ λ where λ is a compact-finite deficient topological measures on X, then µg ≤ λg (respectively, νg ≤ λg) where λg is a deficient topological measure (respectively, a signed deficient topological measure) obtained from λ by Theorem 29. (b11) If µ is a proper deficient topological measure, then so is µg (respectivey, νg). If g > 0 then µ is a proper deficient topological measure iff µg is a proper deficient topological measure. (b12) If µ is a measure then νg is a signed measure, which is a measure if g ≥ 0. If g > 0 then µ is a Radon measure iff µg is a Radon measure. (b13) If µ is finite and g ≥ 0 then µg(X) = µg(Coz(g)). (b14) (Absolute continuity) For any ǫ > 0 there exists δ > 0 such that if A ∈ O(X) ∪ C (X) and µ(A) < δ, then µg(A) < ǫ (respectively, νg(A) < ǫ). Hence, µg ≪ µ (respectively, νg ≪ µ). If X is compact, g, h ∈ C(X) we also have: (i) If c is a constant then νc = cµ and νg+c = νg + cµ. (ii) If g ≤ c, g = c on supp h, h ∈ C(X) then νg+h = νg + νh. (iii) (Lipschitz continuity): νg(A)−νh(A) = RµK (g)−RµK (h) ≤ µ(A)kg− hk for A ∈ O(X) ∪ C (X). Proof. (b1) Since ag ∈ C + 0 (X), we may define the deficient topological mea- sure µag. By Remark 11 it is enough to show that aµg = µag on K (X). Let K ∈ K (X). By formula (5) and part (II) of Remark 17 we have: µag(K) = RµK (ag) = aRµK (g) = aµg(K). Thus, µag = aµg. If g = 0 then by part (II) of Remark 17µg(K) = RµK (0) = 0 for each K ∈ K (X). Thus, µg = 0. The argument for νg is the same. 18 S. V. BUTLER, UCSB (b2) We need to check that µg+h = µg + µh on K (X). For K ∈ K (X) by part (II) of Remark 17 µg+h(K) = RµK (g + h) = RµK (g) + RµK (h) = µg(K) + µh(K). The proof for νg is the same. (b3) By Remark 11 it is enough to check that µg ≤ µh (or νg ≤ νh) on K (X), which easily follows from part (II) of Remark 17. (b4) Let A ∈ O(X) ∪ C (X), µ(A) < ∞. By formula (9) we have a′µ(A) ≤ a′ µ(A) dt = b′µ(A), a′ R2,µA,f (t) dt ≤ a′µ(A) +R b′ RA f dµ ≤ a′µ(A) +R b′ and the statement follows. (b5) If F ∈ C (X) and g = h on F then R2,µF ,g(t) = R2,µF ,h(t) for every t > 0. If U ∈ O(X) and g = h on U then R1,µU ,g = R1,µU ,h for every t ≥ 0. From formula (8) (applied on [a, b] that contains f (X) and g(X)) we see that µg(A) = µh(A) for A = F or A = U . For νg we apply a similar argument using formula (6). (b6) Let K ∈ K (X). Using [9, Lemma 7.12(z8)] and part (v7) of Theorem 24 we have: µg(K) − µf (K) = RµK (g) − RµK (f ) ≤ µK(X)kg − f k = µ(K)kg − f k. (b7) Follows from formula (11). (b8) Follows from part (b7) and regularity of µ. (b9) It is enough to check that µg(K) = cλg(K) + dδg(K) (respectively, νg(K) = cλg(K) + dδg(K)) for each K ∈ K (X), which follows easily from formula (8) (respectively, from (6)). (b10) Let K ∈ K (X). By part (v4) of Theorem 24 µK ≤ λK . By [9, Theorem 8.7 (II)] µg(K) = RµK (g) ≤ RλK (g) = λg(K). The statement follows. (b11) Follows from formula (11) and [11, Theorem 4.5(I), Section 4]. (b12) See first Remark 31. Now let g > 0. Assume that µg is a Radon measure. By [11, Theorem 4.3, Section 4] we may write µ = m + λ where m is a Radon measure and λ is a proper deficient topological measure. By part (b9) µg = mg + λg. But since µg and mg are both Radon measures, by uniqueness of decomposition in [11, Theorem 4.3, Section 4] we see that λg = 0. Then from part (b4) it follows that λ(K) = 0 for every K ∈ K (X). Then λ = 0 and µ = m is a Radon measure. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES19 (b13) Let P = Cozg. Since R1,µP ,g(t) = R1,µX ,g(t) for all t ≥ 0, by formula (8) we have µg(P ) = RµP (g) = RµX (g) = µg(X). (b14) Follows from (12). (i) It is enough to show that νg+c = νg + cµ on K (X); the other statement then will follow from part (b1). For K ∈ K (X) by [9, Lemma 7.12(i)] and part (v7) of Theorem 24 we have: (13) νg+c(K) = RµK(g + c) = RµK (g) + cµK (X) = νg(K) + cµ(K). (ii) Suppose first 0 ≤ g ≤ c, g = c on supp h. We need to check that νg+h = νg + νh on K (X). Let K ∈ K (X). From [9, Theorem 7.10, Lemma 4.4] it follows that RµK satisfies the c-level condition (see [9, Definition 4.3], so νg+h(K) = RµK (g + h) = RµK (g) + RµK (h) = νg(K) + νh(K). For the general case g ≤ c, g = c on supp h choose constant b ≥ 0 such that 0 ≤ g + b ≤ c + b on supp h. For any K ∈ K (X) we have: νg+b+h(K) = νg+b(K)+νh(K), which by (13) gives νg+h(K)+bµ(K) = νg(K) + bµ(K) + νh(K), i.e νg+h(K) = νg(K) + νh(K). Thus, νg+h = νg + νh. (iii) Using [9, Lemma 7.12(ii), ] and part (v7) of Theorem 24, as in part (b6) we obtain the statement. (cid:3) Remark 34. Example 30 and similar examples show that inequalities in part (b4) of Theorem 33 can be realized as equalities. Theorem 35. Suppose X is locally compact and g ∈ C + 0 (X), or X is compact and g ∈ C(X). Suppose µ is a compact-finite topological measure on X. Suppose A, B, A ⊔ B ∈ O(X) ∪ C (X), and µ(A ⊔ B) < ∞. Then ZA⊔B g dµ ≤ZA g dµ +ZB g dµ + µ(A ⊔ B) max{ω(g, A), ω(g, B)}, where ω(f, A) = supx∈A f (x) − inf x∈A f (x). 20 S. V. BUTLER, UCSB Proof. Let g(X) ⊆ [a, b] where a ≤ 0. Suppose first that D ⊔ V = U , where D ⊆ C (X), V, U ∈ O(X), and µ(U ) < ∞. Let p = inf x∈D g(x), q = supx∈D g(x). If a < t < p then D ⊆ g−1((t, ∞)), and so g−1((t, ∞)) ∩ U = (g−1((t, ∞)) ∩ V ) ⊔ D. Thus, R1,µU ,g(t) = R1,µV ,g(t) + µ(D). If t > q then g−1((t, ∞)) ∩ D = ∅, so g−1((t, ∞)) ∩ U = g−1((t, ∞)) ∩ V , i.e. R1,µU ,g(t) = R1,µV ,g(t). Also, Z q p (R1,µU ,g(t) − R1,µV ,g(t))dt ≤Z q p R1,µU ,g(t)dt ≤ (q − p)µ(U ) = ω(g, D)µ(D ⊔ V ). By superadditivity µ(U ) − µ(V ) − µ(D) ≥ 0. Using also (6) and (11) we obtain: ZU g dµ −ZV g dµ = a(µ(U ) − µ(V )) +Z b a (R1,µU ,g(t) − R1,µV ,g(t))dt = a(µ(U ) − µ(V )) +Z p µ(D)dt +Z q a p (R1,µU ,g(t) − R1,µV ,g(t))dt ≤ a(µ(U ) − µ(V )) + (p − a)µ(D) + ω(g, D)µ(D ⊔ V ) = a(µ(U ) − µ(V ) − µ(D)) + pµ(D) + ω(g, D)µ(D ⊔ V ) ≤ pµ(D) + ω(g, D)µ(D ⊔ V ) ≤ZD g dµ + ω(g, D)µ(D ⊔ V ). The statement now follows. All other possible cases are proved similarly. For example, if D ⊔V = C, where D, C ⊆ C (X), V ∈ O(X) and µ(C) < ∞, we use the fact that R2,µC ,g(t) = R2,µD,g(t) + µ(V ) for t < p = inf x∈V g(x), and R2,µC ,g(t) = R2,µD ,g(t) for t > q = supx∈V g(x). (cid:3) Remark 36. Theorem 35 says that for a deficient topological measure µg or a signed deficient topological measures νg obtained in Theorem 29 we have (14) µg(A ⊔ B) ≤ µg(A) + µg(B) + µ(A ⊔ B) max{ω(g, A), ω(g, B)}, νg(A ⊔ B) ≤ νg(A) + νg(B) + µ(A ⊔ B) max{ω(g, A), ω(g, B)}, whenever A, B, A ⊔ B ∈ O(X) ∪ C (X), and µ(A ⊔ B) < ∞. When X is compact, in [19, Theorem 5.10] it was shown that µg(A ⊔ B) ≤ µg(A) + µg(B) + µ(X) max{ω(g, A), ω(g, B)}, and it inspired Theorem 35. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES21 By Theorem 6, µg is a topological measure iff µg(U ) ≤ µg(K) + µg(U \ K) for any open U and any compact K ⊆ U . Example 32 shows that even when µ is a finitely defined simple topological measure and g is strictly positive, µg may not be a topological measure. This is a contrast to the situation with measures (see part (b12) of Theorem 33). Inequality (14) is what we can say in general. Using Theorem 7, we can also say, that, unless mg is a measure, any inequality that estimates µg(A ⊔ B) above using µg(A) and µg(B) (like (inequality 14)) must have the form µg(A⊔B) ≤ µg(A)+µg(B)+µ(A⊔B)+δ(A, B) with δ(A, B) > 0 for some A, B ∈ O(X) or A, B ∈ K (X). 3. INTEGRALS OVER A SET VIA FUNCTIONALS In this section we present another interpretation of the deficient topological mea- sure µg given by Theorem 29 in the case where µ is finite. Let X be locally compact. Let µ be a finite deficient topological measure on X with the corresponding functional ρ = Rµ according to part (I) of Remark 17. Let g ∈ C + 0 (X). Let ρg be a functional on C0(X) defined by ρg(f ) = ρ(f g). Since ρ is a d-functional, so is ρg. Definition 37. Letcµg be the deficient topological measure corresponding to ρg by part (III) of Remark 17. Remark 38. By part (III) of Remark 17 cµg(K) = inf{ρg(f ) = ρ(f g) : f ≥ 1K , f ∈ Cc(X)} = inf{ρg(f ) = ρ(f g) : 1K ≤ f ≤ 1, f ∈ Cc(X)}, for K ∈ K (X), and for U ∈ O(X) cµg(U ) = sup{ρg(f ) = ρ(f g) : 0 ≤ f ≤ 1, f ∈ Cc(X), supp f ⊆ U }. Remark 39. If µ is a regular Borel measure, it is easy to see that cµg(A) =ZA g dµ for every A ∈ O(X) ∪ C (X). In the next theorem we will show that the equality cµg(A) =RA g dµ also holds when µ is a finite deficient topological measure. 22 S. V. BUTLER, UCSB Theorem 40. Let X be locally compact. Let µ be a finite deficient topological measure on X. Let g ∈ C + 0 (X). Let cµg be as in Definition 37, and let µg be the deficient topological measure given by Theorem 29. Thencµg = µg. Proof. It is enough to show that cµg = µg on K (X). Let K ∈ K (X). Let g(x) ⊆ [0, b], b ≥ 1. By formula (8) (15) µg(K) = RµK (g) =Z b 0 R2,µK ,g(t), where R2,µK ,g(t) is given by formula (7). Take any f ∈ Cc(X) such that 1K ≤ f ≤ 1. Then [0, b] contains the ranges of functions g and gf . By formula (2) ρ(f g) =Z b 0 R2,µ,f g(t) dt. Note that g−1([t, ∞)) ∩ K ⊆ (f g)−1([t, ∞)), so R2,µK ,g(t) ≤ R2,µ,f g(t). Then by (15) µg(K) ≤ ρ(f g) = ρg(f ). From Remark 38 we see that µg(K) ≤ cµg(K). Now we shall show the opposite inequality. By Lemma 3 choose U, V ∈ O(X) and C = V ∈ K (X) such that K ⊆ V ⊆ C ⊆ U . Let K = {C ∈ K (X) : ∃V with C = V , K ⊆ V ⊆ C ⊆ U }. K is a directed set with respect to inverse inclusion and TC∈K C = K. Let C ∈ K. Let f be the Urysohn function such that 1K ≤ f ≤ 1V . For every t > 0 we have (f g)−1([t, ∞)) ⊆ g−1([ f −1((0, ∞)) ⊆ g−1([t, ∞)) ∩ C, so R2,µ,f g(t) ≤ R2,µC ,g(t). Then t f (t) , ∞)) ∩ cµg(K) ≤ ρg(f ) = ρ(f g) =Z b 0 R2,µ,f g(t) dt ≤Z b 0 R2,µC ,g(t). Each function R2,µC ,g is bounded above by µ(U ) < ∞. Also, R2,µC ,g is left- continuous on [0, b] (where left-continuity of R2,µC ,g(t) for t > 0 follows from 0 R2,µC ,g(t) dt → (cid:3) τ -smoothness on compact sets in Remark 8). By Corollary 21R b R b 0 R2,µK ,g, and socµg(K) ≤R b We can say more about the values of the deficient topological measure µg using 0 R2,µK ,g(t) dt = µg(K). the measure nr on X given by [9, Definition 7.15]: Theorem 41. Suppose µ is a finite deficient topological measure on a locally com- pact space X, g ∈ C + 0 (X). Then (i) For E = g−1(A), where A ⊆ R is open or closed, µg(E) ≤RE g dnr. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES23 g−1(A), where A ⊆ (−∞, 0) is open or closed, (ii) µg(Kerg) = RKerg g dnr = 0, and µg(E) = RE g dnr = 0 for E = (iii) If µ is a topological measure then µg(E) = RE g dnr for E = g−1(A), (iv) If X is compact and µ is a topological measure, then νg(E) = RE g dnr (v) If X is compact then µ is a topological measure on X iff µg(E) =RE g dnr for E = g−1(A), where A ⊆ R is open or closed. where A ⊆ R \ {0} is closed or A ⊆ R is open. for any g ∈ C +(X) and E = g−1(C) for any closed C ⊆ R. Proof. (i) First, let D = g−1(C), where C ⊆ R is closed. Let g(X) ⊆ [0, b]. Set the measure p to be the restriction of the measure nr to D. Then 0 F (t)dt, where F is the distribution function for RD g dnr =RX g dp =R b the measure p ◦ g−1 defined by F (t) = p ◦ g−1([t, ∞)) = p(g−1([t, ∞))) = nr(g−1([t, ∞)) ∩ D). For t > 0 using [9, Lemma 7.17] we have: F (t) = nr(g−1([t, ∞)) ∩ D) = nr(g−1([t, ∞)) ∩ g−1(C)) = nr(g−1([t, ∞) ∩ C)) ≥ µ(g−1([t, ∞) ∩ C)) = µ(g−1([t, ∞)) ∩ D) = R2,µD ,g(t), so ZD g dnr =Z b 0 F (t)dt ≥Z b 0 R2,µD ,g(t)dt = µg(D). The case U = g−1(W ), where W ⊆ R is open, is similar. We use intervals (t, ∞) instead of [t, ∞) in the distribution function F . Then as above, F (t) ≥ R1,µU ,g(t) andRU g dnr =R b 0 F (t)dt ≥R b µg(U ). 0 R1,µU ,g(t)dt = (ii) Using part (i), 0 ≤ µg(Kerg) ≤RKerg g dnr = 0. If A ⊆ R is an open or closed subset of (−∞, 0) then E = g−1(A) = ∅. So µg(E) =RE g dnr = 0. (iii) Let A ⊆ R \ {0} be closed or open and E = g−1(A). The argument as in the proof of part (i), where by [9, Lemma 7.17(iii)] inequalities become equalities, shows that µg(E) = RE g dnr. We are only left to consider the 24 S. V. BUTLER, UCSB case µr(U ), where U = g−1(W ) for an open set W ⊆ R such that 0 ∈ W . Write W = W1 ⊔ {0}, U1 = g−1(W1). Then U = U1 ⊔ Kerg and g dnr +ZKerg µg(U ) ≥ µg(U1) + µg(Kerg) =ZU1 Together with part (i) this gives µg(U ) =RU g dnr. g dnr =ZU g dnr. (iv) The argument is essentially the same as the one for part (i), where by [9, Lemma 7.17(iv)] inequalities become equalities. For example, for D we have F (t) = R2,µD ,g(t) for any t ∈ [a, b] = g(X), and then ZD g dnr = ap(X) +Z b a F (t)dt = aµ(D) +Z b a R2,µD ,g(t)dt = νg(D). (Note that when g ≥ 0, one may also use the argument from [18, Theorem 20].) (v) The proof is basically one from [18, Theorem 20] and is given here for completeness. Let Z ⊆ X be a zero set. Say, Z = f −1(0). We may assume that 0 ≤ f ≤ 1. Let g = 1 − ≤ g ≤ 1, and g−1(1) = Z. Let U = X \ Z. Using formula (11) and part (iv) f . Then g ∈ C +(X), 1 − 1 n 1 n we have: (1 − 1 n )µ(X) ≤ µg(X) =ZX g dnr =ZZ g dnr +ZU g dnr = µg(Z) + µg(U ) ≤ µ(Z) + µ(U ). It follows that µ(X) ≤ µ(Z) + µ(U ) = µ(Z) + µ(X \ Z). Every locally compact space is completely regular, and so the family of all zero sets is a base for closed sets. From Remark 8 it follows that µ(X) ≤ µ(C) + µ(X \ C) for any closed set C, and by Theorem 6 µ is a topological measure. (cid:3) 4. INTEGRATION OVER ZERO AND COZERO SETS Integration with respect to a deficient topological measure sometimes gives the same results as integration with respect to a measure would give. For example, we have the following lemma: Lemma 42. Supppose X is locally compact, µ is a deficient topological measure on X, f ∈ Cb(X). If A ∈ O(X) ∪ C (X) is such that µ(A) = 0 thenRA f dµ = 0. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES25 Proof. Let f (X) ⊆ [a, b]. Consider first V ∈ O(X) such that µ(V ) = 0. Since R1,µV ,f (t) = 0 for any t ∈ [a, b], using formula (6) we see thatRV f dµ = 0. For C ∈ C (X), µ(C) = 0 we may apply a similar argument using formula (6) and the fact that R2,µC ,f (t) = 0. (cid:3) Integration with respect to a deficient topological measure sometimes is also very different from integration with respect to a measure. If µ is a measure and µ(Coz(f )) = 0, then RX f dµ = 0. Example 43 below shows that this is gener- ally false if µ is a deficient topological measure. (On the other hand, part (y1) of Theorem 44 below shows that this is true for nonnegative f , assuming that µ(Cozf ) < ∞ (in particular, for finite µ); this theorem also generalizes Lemma 42). Example 43. Let X = R, D = [1, 4] and µ be a deficient topological measure such that µ(A) = 1 if D ⊆ A and µ(A) = 0 otherwise, for any A ∈ O(X) ∪ C (X). (See [7, Section 6]). Let f = (x − 2) ∧ 0, so f (X) = [−2, 0], Coz(f ) = (−2, 2) and µ(Coz(f )) = 0. Note that R2,µ,f (t) = 1 if t ∈ [−2, −1] and R2,µ,f (t) = 0 if t ∈ (−1, 0]. By formula (3) we have: ZX f dµ = R(f ) = −2µ(X) +Z 0 −2 R2,µ,f (t)dt = −1. Now let g = −f , so g(X) = [0, 2]. Since R2,µ,g(t) = 0 for every t > 0, we have Thus, (−f )dµ =Z 2 0 f dµ = −1, ZX ZX R2,µ,g(t)dt = 0. ZX (−f )dµ = 0. Theorem 44. Suppose X is locally compact, µ is a deficient topological measure on X, f ∈ Cb(X). Then (y1) If f ≥ 0, then for any open set V with µ(V ) < ∞ ZV f dµ =ZV ∩Coz(f ) f dµ, and ZV f dµ = 0 ⇐⇒ µ(V ∩ Coz(f )) = 0. 26 S. V. BUTLER, UCSB In particular, for a finite µ we haveRX f dµ =RCoz(f ) f dµ, and ZX f dµ = 0 ⇐⇒ZCoz(f ) f dµ = 0 ⇐⇒ µ(Coz(f )) = 0. (y2) If V ∈ O(X), µ(V ) < ∞, then µV (Z(f )) = µ(V ) =⇒ZV f dµ = 0. In particular, if µ is finite and µ(Z(f )) = µ(X), thenRX f dµ = 0. (y3) If X is compact, V ∈ O(X), and f ≤ 0 then µV (Z(f )) = µ(V ) ⇐⇒ZV f dµ = 0. In particular, ZX f dµ = 0 ⇐⇒ µ(Z(f )) = µ(X). Proof. (y1) Let f ∈ Cb(X), f ≥ 0. Take any V ∈ O(X) with µ(V ) < ∞, and let U = V ∩ Coz(f ). For t > 0 we have U ∩ f −1((t, ∞)) = V ∩ Coz(f ) ∩ f −1((t, ∞)) = V ∩ f −1((t, ∞)), so from (7) and ( 8) we see that R1,µU ,f (t) = R1,µV ,f (t), andRV f dµ = RV ∩Coz(f ) f dµ. Now suppose RV f dµ = R b 0 R1,µV ,f (t) = 0. Since the integrand in the second integral is nonnegative, using the right-continuity of R1,µV ,f (t) (see [9, Lemma 6.3]), we must have µ(V ∩ Coz(f )) = R1,µV ,f (0) = 0. On the other hand, if µ(V ∩Coz(f )) = 0, then R1,µV ,f (t) = 0 for every t > 0, and by formula (6)RV f dµ = 0. (y2) Suppose f (X) ⊆ [a, b], a ≤ 0, b ≥ 0, and µV (Z(f )) = µ(V ) < ∞. Using part (v7) of Theorem 24 and superadditivity (see Remark 8), we have µ(V ) = µV (X) ≥ µV (Z(f )) + µV (Coz(f )), and so we must have µV (Coz(f )) = µ(V ∩ Coz(f )) = 0. With W = (f −1(t, ∞)) ∩ Coz(f ), we have: µV (W ) = 0. Since f −1((t, ∞)) ∩ Z(f ) = ∅ for t ≥ 0 and (f −1(t, ∞)) ∩ Z(f ) = Z(f ) for t < 0, we see that f −1((t, ∞)) = W if t ≥ 0, and f −1((t, ∞)) = Z(f ) ⊔ W if t < 0. Then R1,µV ,f (t) = µV (W ) = 0 for t ≥ 0 and R1,µV ,f (t) = µV (f −1((t, ∞))) ≥ µV (Z(f ))+ INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES27 µV (W ) = µV (Z(f )) = µ(V ) for t < 0. Since also R1,µV ,f (t) ≤ µ(V ), we have R1,µV ,f (t) = µ(V ) for t < 0. Then ZV f dµ = aµ(V ) +Z b a R1,µV ,f (t) = aµ(V ) +Z 0 a µ(V )dt = 0. (y3) Suppose that X is compact. (=⇒) is given by part (y2). (⇐=) Assume thatRV f dµ = 0, i.e. by formula (6) f dµ = aµ(V ) +Z 0 R2,µV ,f (t)dt =Z 0 ZV a a (R2,µV ,f (t) − µ(V ))dt = 0. The integrand function in the last integral is nonpositive. Since X is com- pact, by [9, Lemma 6.3(V)] R2,µV ,f (t) is left-continuous at 0. From left continuity it follows that R2,µV ,f (0) − µ(V ) = 0. But R2,µV ,f (0) = µV (Z(f )). (cid:3) Lemma 45. If µ assumes only finitely many values and is not a topological mea- sures then there exists f ≥ 0 such thatRX f dµ = 0, RX(−f )dµ < 0. Proof. Suppose that µ is a deficient topological measure, but not a topological measure and that µ assumes finitely many values. By Theorem 6 there are U ∈ O(X) and C ∈ K (X), C ⊆ U such that µ(C) + µ(U \ C) < µ(U ). We may assume that U is compact. (Otherwise, choose by regularity compact D ⊆ U with µ(D) = µ(U ), then using regularity and Lemma 3 choose U1 ∈ O(X) such that U1 ∈ K (X), D ⊆ U1 ⊆ U, and µ(U1) = µ(D) = µ(U ). Then µ(C) + µ(U1 \ C) ≤ µ(C) + µ(U \ C) < µ(U ) = µ(U1), and we replace U by U1.) We have (16) µ(C) + µ(U \ C) + µ(X \ U ) < µ(U ) + µ(X \ U ) ≤ X. Pick a compact B ⊆ U \ C with µ(B) = µ(U \ C). Let V = (U \ C) \ B. By superadditivity µ(V ) = 0. Set a closed set G = (X \ U ) ⊔ B ⊔ C. By Remark 8 µ(G) = µ(X \ U ) + µ(B) + µ(C), which is equal to the left-hand side of (16), so µ(G) < µ(X). Now choose W ∈ O(X) such that G ⊆ W, µ(W ) = µ(G). Let δ = µ(X) − µ(W ) > 0. 28 S. V. BUTLER, UCSB Let K = X \W . Then K ⊆ X \G = V ⊆ U , so K is compact. Let f ∈ Cc(X) be a Urysohn function such that f = 1 on K and supp f ⊆ V . Then Coz(f ) ⊆ V , so µ(Coz(f ) = 0, and by part (y1) of Theorem 44RX f dµ = 0. Now let g = −f , so g(X) ⊆ [−1, 0]. For t ∈ (−1, 0) we have g−1((t, ∞)) ⊆ g−1((−1, ∞)) ⊆ X \ K = W , so R1,µ,g(t) ≤ µ(W ). Then by formula (1) ZX g dµ = −µ(X) +Z 0 −1 R1,µ,g(t)dt ≤Z 0 −1 (µ(W ) − µ(X))dt = −δ < 0. (cid:3) Remark 46. Example 43 illustrates Lemma 45. Also, in Example 43 RX f dµ = −1, while µ(Coz(f )) = 0. This shows that part (y1) of Theorem 44 may not hold if the condition f ≥ 0 is relaxed. Remark 47. Parts (v1) - (v6) of Theorem 24 for the compact space are stated with- out proof in [19, Proposition 5.2]. Some statements from Theorem 29 and Lemma 28 are related to [19, Theorem 5.4]. Parts (b1) - (b4), (b9) - (b14) of Theorem 33 are generalizations of results presented in [18, Theorem 22] and [19, Theorem 5.7, Proposition 5.9]; for part (i) of Theorem 33 see also [19, Theorem 5.7]. The- orem 40 is inspired by [18, Theorem 19]. Statements "In particular..." in Theorem 44 are generalizations of [19, Theorem 3.6, 1-3], and Lemma 45 generalizes [19, Theorem 3.6, 4]. 5. CONVERGENCE THEOREMS Theorem 48. Let X be a locally compact space, µ a finite deficient topological measure on X. (I) Suppose fα ∈ Cb(X) and fα ր f in the topology of uniform convergence. Then ZX fα dµ −→ZX f dµ. (II) Suppose fα ∈ C + 0 (X) and fα ց f in the the topology of uniform conver- gence. Then ZX fα dµ −→ZX f dµ. (I) Since f ∈ Cb(X), we may considerR f dµ. For any t, f −1 f −1((t, ∞)), so by Remark 8 µ(f −1 α ((t, ∞))) → µ(f −1((t, ∞))). α ((t, ∞)) ր i.e. Proof. INTEGRATION WITH RESPECT TO DEFICIENT TOPOLOGICAL MEASURES ON LOCALLY COMPACT SPACES29 R1,µ,fα(t) → R1,µ,f (t). Applying formula (1) on [a, b] containing f (X) we see that ZX fα dµ →ZX f dµ. (II) The proof is similar. Note that for any t > 0 the set f −1 α ([t, ∞)) is com- pact, f −1 α ([t, ∞)) ց f −1([t, ∞)), and we may again apply Remark 8. (cid:3) Theorem 49. Let µ be a compact-finite deficient topological measure on a locally compact space X. Suppose fα converges uniformly to f , fα, f ∈ Cc(X), fα ≥ 0. If supp fα, supp f ⊆ K for some compact K, then RX fα dµ −→ RX f dµ. If X is compact, fα ∈ C(X) and fα converges uniformly to f then RX fα dµ −→ RX f dµ. Proof. Follows from parts (b6) and (iii) of Theorem 33. (cid:3) Acknowledgments: The author would like to thank the Department of Mathe- matics at the University of California Santa Barbara for its supportive environment. REFERENCES [1] J. F. Aarnes, Quasi-states and quasi-measures, Adv. Math. 86, no. 1 (1991), 41 -- 67. [2] Aarnes J. Pure quasi-states and extremal quasi-measures. Math. Ann. 295 (1993), 575 -- 588. [3] J. F. Aarnes, Construction of non-subadditive measures and discretization of Borel mea- sures, Fund. Math. 147 (1995), 213 -- 237. [4] J. F. Aarnes, S. V. Butler, Super-measures and finitely defined topological measures Acta Math. Hungar. 99 (1-2) (2003), 33 -- 42. [5] V. I. Bogachev, Measure Theory, vol. 2, Regular and Chaotic Dynamics, Izhevsk 2003, English transl., Springer-Verlag, Berlin, 2007. [6] S. Butler. Ways of obtaining topological measures on locally compact spaces Bull. of Irkutsk State Univ., Series "Mathematics" 25 (2018), 33 -- 45. [7] S. V. Butler, Deficient topological measures on locally compact spaces, arXiv: 1902.02458 [8] S. V. Butler, Quasi-linear functionals on locally compact spaces, arXiv: 1902.03358 [9] S. V. Butler, Non-linear functionals, deficient topological measures, and representation theorems on locally compact spaces, arXiv: 1902.05692 [10] S. Butler, Signed topological measures on locally compact spaces, arXiv: 1902.07412 [11] S. V. Butler, Decompositions of signed deficient topological measures, arXiv: 1902.07868 [12] Dugundji, Topology, Allyn and Bacon, Inc., Boston, 1966. 30 S. V. BUTLER, UCSB [13] M. Entov, L. Polterovich, Quasi-states and symplectic intersections, Comment. Math. Helv. 81 (2006), 75 -- 99. [14] ∅. Johansen, A. Rustad, Construction and Properties of quasi-linear functionals, Trans. Amer. Math. Soc. 358, no. 6 (2006), 2735 -- 2758. [15] L. Polterovich, D. Rosen Function theory on symplectic manifolds CRM Monograph series, vol. 34, American Mathematical Society, Providence, Rhode Island, 2014. [16] A. B. Rustad, Unbounded quasi-integrals, Proc. Amer. Math. Soc. 129, no. 1 (2000), 165 -- 172. [17] M. G. Svistula, A Signed quasi-measure decomposition, Vestnik Samara Gos. Univ. Es- testvennonauchn. 62, no. 3 (2008), 192 -- 207 (Russian). [18] M. G. Svistula, Deficient topological measures and functionals generated by them, Sbornik: Mathematics 204, no. 5 (2013), 726 -- 761. [19] M. G. Svistula, On integration with respect to a DT-measure, Positivity, 20, no. 3, (2016), 579 -- 598. DEPARTMENT OF MATHEMATICS, UNIVERSITY OF CALIFORNIA SANTA BARBARA, 552 UNI- VERSITY RD., ISLA VISTA, CA 93117, USA E-mail address: [email protected]
1109.2030
1
1109
2011-09-09T14:51:38
Smoothness spaces of higher order on lower dimensional subsets of the Euclidean space
[ "math.FA" ]
We study Sobolev type spaces defined in terms of sharp maximal functions on Ahlfors regular subsets of the Euclidean space and the relation between these spaces and traces of classical Sobolev spaces.
math.FA
math
SMOOTHNESS SPACES OF HIGHER ORDER ON LOWER DIMENSIONAL SUBSETS OF THE EUCLIDEAN SPACE LIZAVETA IHNATSYEVA AND RIIKKA KORTE Abstract. We study Sobolev type spaces defined in terms of sharp maximal functions on Ahlfors regular subsets of Rn and the relation between these spaces and traces of classical Sobolev spaces. This extends in a certain way the results of Shvartsman [20] to the case of lower dimensional subsets of the Euclidean space. 1. Introduction A. Calder´on proved in [7] that a function belongs to the Sobolev space on Rn if and only if the function and its sharp maximal function of the corresponding order are both in an Lp-space; see also [8]. This characterization does not use the notion of derivatives and therefore it can be used to at least formally define Sobolev spaces in more general settings. Some recent results show that this approach is reasonable. In particular, P. Shvartsman proved in [20] that the trace of a Sobolev space to an arbtirary Ahlfors n-regular subset of the Euclidean space admits an intrinsic Calder´on type characterization. Note that this kind of subsets may even have an empty interior. See also [14], where a description in terms of sharp maximal functions for Sobolev spaces on extension domains was given. The main purpose of this paper is to extend Shvartsman's results to lower dimensional closed subsets of the Euclidean space. In [10] A. Jonsson characterized the trace spaces of Sobolev spaces to Ahlfors regular sets as certain Besov type spaces. Therefore, our problem can be also formulated as comparison of these Besov spaces and Calder´on type spaces. Since traces of Sobolev spaces to lower dimensional sub- sets are of essentially different character than classical Sobolev spaces, an exact characterization of Calder´on type seems not to be possible on such subsets. However, in our main result, Theorem 4.1, we show that Calder´on type spaces lie between certain Besov spaces. This result in particular implies that the trace space of a Sobolev space is embedded in the Calder´on type space of the corresponding order and that it con- tains any Calder´on space of higher order, see Corollary 4.4. Our results Date: August 11, 2021. 2010 Mathematics Subject Classification. 46E35. 1 have a similar spirit as Theorem 4 in [15], where relations between the trace spaces of first order Sobolev spaces and Haj lasz-Sobolev spaces were explored. In particular, their result is also of embedding type, not a sharp characterization, which is not surprising since the Haj lasz- Sobolev space coincides with a Calder´on type space in their context. In this paper, we only consider Ahlfors regular sets whose codimension is less than one. This lower bound for the dimension was due to the observation that usual properties of sharp maximal functions on Rn, studied e.g. in [9], remain valid on sets preserving Markov's inequalities for polynomials, and by [12] Ahlfors s-regular sets with n − 1 < s ≤ n have this property. This class of sets includes, for example, many interesting Cantor type sets and self-similar sets. There are very few approaches to spaces of higher order smoothness even on such kind of sets, in spite of the fact that the first order smooth- ness spaces have been extensively studied in different situations. One of the goals of the paper is to show the advantage of Calder´on's approach or, more precisely, its local polynomial approximation interpretation in [4], [9] and [20], in defining Sobolev type spaces in more general setting; see the related discussion in Section 5. 2. Preliminaries Let H s denote the s-dimensional Hausdorff measure on Rn and let Q = Q(x, r) be a closed cube in Rn centered at x with side length 2r and sides parallel to the coordinate axes. We say that a subset S ⊂ Rn is an s-set (or Ahlfors s-regular) if there are constants c1, c2 > 0 such that for every cube Q = Q(x, r) with center at S and diam Q ≤ diam S, we have c1rs ≤ H s(Q(x, r) ∩ S) ≤ c2rs. In this paper, we will always assume that S ⊂ Rn is an s-set with n − 1 < s ≤ n. 2.1. Sobolev spaces. Let Lp(Rn) be the Lebesgue space of Lp-integrable functions in Rn. For a non-negative integer k and 1 ≤ p ≤ ∞, the Sobolev space W k,p(Rn) consists of all functions f ∈ Lp(Rn) having distributional derivatives Djf , j ≤ k, in Lp(Rn). There are several approaches to the notion of smoothness spaces of fractional order. One of them is as follows. 2.2. Potential spaces. The Bessel kernel of order α > 0 is the func- tion Gα ∈ L1(Rn) defined by G(ξ) = (1 + 4π2ξ2)−α/2. 2 The potential space Lp α(Rn), α ≥ 0, 1 ≤ p ≤ ∞, is Lp α(Rn) = {f = Gα ∗ g : g ∈ Lp(Rn)}, α > 0, and Lp 0(Rn) = Lp(Rn). It was shown already by Calder´on [6] that if 1 < p < ∞ and α is a non-negative integer, the Sobolev spaces and potential spaces coincide, i.e. Lp k(Rn) = W k,p(Rn), 1 < p < ∞, k ∈ N. 2.3. Besov spaces. Another scale of spaces which is widely used in the study of fractional order smoothness properties of functions is the family of Besov spaces. For α > 0 and 1 ≤ p, q ≤ ∞, the Besov space α (Rn) may be defined in the following way. Let k be the integer Bp,q such that 0 ≤ k < α ≤ k + 1. Then Bp,q α (Rn) consists of functions f ∈ Lp(Rn) such that Xj≤k kDjf kp + Xj=k(cid:18)ZRn kDjf (· + h) − Djf (·)kq p hn+(α−k)q dh(cid:19)1/q < ∞, (2.1) if k < α < k + 1 and 1 ≤ p, q < ∞. If q = ∞, then (2.1) shall be interpreted in the usual limiting way and if α = k + 1, the first difference of Djf in (2.1) shall be replaced by the second difference. For more details, see [3]. There are several equivalent characterizations of Besov spaces Bp,q α (Rn), for a general theory of these spaces see, for example, monographs [3], [21] and the references therein. We are interested in Besov spaces as spaces of traces of functions from Sobolev or, more general, poten- tial spaces to subsets of Rn. See the next paragraph for more precise formulations. Jonsson and Wallin [11] extended the notion of a Besov space to more general setting. They introduced a definition of Besov spaces on general s-sets, 0 < s ≤ n, in Rn. The definition is rather technical, but when n − 1 < s ≤ n, it admits a more simple formulation, which is based on the local polynomial approximation approach, see Theorem 5 on p. 135 of [12]. In this paper, we will use this formulation. Definition 2.2. Let S be an s-set with n−1 < s ≤ n. Let 1 ≤ p, q ≤ ∞ and α > 0. Then a function f is in the Besov space Bp,q α (S) if f ∈ Lp(S) and there is a sequence {cν}∞ ν < ∞, such that for every net π with mesh size 2−ν, ν = 0, 1, . . . , there is a function Pπf ∈ P[α](π) satisfying ≤ 2−ναcν. (2.3) ν=0, Pν cq f − Pπf p dH s(cid:19)1/p 3 (cid:18)ZS Here Pk(π) denotes the set of all functions g such that the trace of g to a cube Q ∈ π is a polynomial of degree at most k, and [α] is the largest integer that is not greater than α. Note that in [12], the definition is stated for s-sets preserving Markov's inequality. However, by Theorem 3 on p. 39 of [12], all s-sets with s > n − 1 satisfy this condition. 2.4. Traces of Sobolev functions. We say that f can be pointwisely defined at x if the limit ¯f (x) = lim r→0ZQ(x,r) f (y) dy = lim r→0 1 Q(x, r)ZQ(x,r) f (y) dy exists. By Lebesgue's theorem f = ¯f a.e. in Rn. At every x ∈ S where ¯f (x) exists, we define the trace of a function f to S by f S(x) := ¯f (x). If f ∈ Lp β(Rn), 1 < p < ∞, then the Hausdorff dimension of the set of points x ∈ Rn where ¯f (x) does not exists is at most n − βp, see for example [1]. Thus the trace of a function f ∈ Lp β(Rn) to an s-set, s > n − βp, is well defined i.e. f S is defined at H s-a.e. point of S. The next statement, which was proved by A. Jonsson in [10], gives a characterization of the trace of the potential space to an s-set. Theorem 2.4. Let S be an s-set, 0 < s < n, 1 < p < ∞ and α = β − n−s p > 0. Then Lp β(Rn)S = Bp,p α (S), where the equality means that the trace operator R : f 7→ f S satisfies the inequality kRf kBp,p α (S) ≤ ckf kLp β(Rn) for some constant c and for all functions f ∈ Lp ists an extension operator E : Bp,p constant c, we have α (S) → Lp β(Rn), and there ex- β(Rn) such that for some kEgkLp β(Rn) ≤ ckgkBp,p α (S) for all functions g ∈ Bp,p α (S). Since for 1 < p < ∞ and for nonnegative integers β, the potential space Lp β(Rn) coincides with the Sobolev space W β,p(Rn), the theorem above, in particular, gives a characterization for traces of Sobolev spaces. 4 3. The sharp maximal functions on s-sets and corresponding smoothness spaces In this section, we introduce Calder´on type smoothness spaces on s- regular subsets of Rn which are defined in terms of fractional sharp maximal functions. We start with the basic notions. 3.1. Local polynomial approximations. Let f ∈ Lu(S, H s), 0 < u ≤ ∞, Q be a cube in Rn and QS = Q ∩ S. Then the normalized local best approximation of f on Q in Lu(S) norm is Ek(f, Q)Lu(S) := inf p∈Pk−1(cid:18)ZQS f − pu dH s(cid:19)1/u , where Pk, k ≥ 0, is a family of all polynomials on Rn of degree at most k. We also set P−1 := {0}. Note that for every pair of cubes such that Q1 ⊂ Q2, we have and by the s-regularity of a set S, this further implies that H s(Q1 ∩ S)(cid:19)1/u Ek(f, Q1)Lu(S) ≤(cid:18) H s(Q2 ∩ S) r1(cid:19)s/u Ek(f, Q1)Lu(S) ≤ c(cid:18)r2 Ek(f, Q2)Lu(S) (3.1) Ek(f, Q2)Lu(S). (3.2) Here Qi = Q(xi, ri). In the setting of the Euclidean space, Ek(f, Q)Lu(Rn) is the main object of the theory of local polynomial approximation and, in particular, it gives a unified framework for the description of various spaces of smooth functions, see for example the survey [4]. 3.2. Maximal functions. Fix α > 0 and set k = −[−α], i.e. the greatest integer strictly less than α+1. For a locally integrable function f on S, we define the fractional sharp maximal function f ♯ α,u,S(x) := sup t>0 1 tα Ek(f, Q(x, t))Lu(S), x ∈ S. α,S instead of f ♯ α,1,S for short. From now on, we will write f ♯ (3.3) Since S is an s-set, it follows from (3.2) that the supremum over cubes centered at x in the definition above can be replaced by the supremum over all cubes with centers in S containing point x. When S = Rn, maximal functions of this type were first introduced by Calder´on [7] (see also the paper of Calder´on and Scott [8]). It follows from the results of [7] that a function belongs to the Sobolev space W k,p(Rn), 1 < p < ∞, if and only if f and f ♯ k,Rn are both in Lp(Rn). 5 Motivated by Calder´on's characterization of Sobolev spaces define the following function spaces on s-sets C p α(S) = {f ∈ Lp(S) : kf kC p α = kf kp + kf ♯ α,Skp < ∞}, p ≥ 1. (3.4) Remark 3.5. If, in the definition (3.3), we make another choice for the degree of projection, namely, we set k = [α] + 1 i.e. the smallest integer that is strictly larger than α, we will get another variant of a fractional maximal function. We will denote it by f ♭ differs from f ♯ α,S. Clearly, f ♭ α,S only if α is an integer. α,S Fractional sharp maximal functions on Rn and the corresponding smooth- ness spaces were studied in detail in the monograph of R. DeVore and R. Sharpley [9]. Note that in this paper, we use the same notation as [20], but it differs from the one in [9]. P. Shvartsman proved in [20] that, when S is an n-regular subset of Rn, the trace space to S of the Sobolev space can be characterized via sharp maximal functions, namely, W k,p(Rn)S = C p k(S), p > 1. We aim to study the relationship between the trace spaces of W k,p(Rn) to an s-set S, n − 1 < s < n, and the spaces of functions defined in terms of sharp maximal functions on S. Since in this case the trace space W k,p(Rn)S coincides with the Besov space Bp,p α (S), α = k − n−s p > 0 (see Theorem 2.4), the problem can be also formulated as the comparison of Bp,p α(S). Note that in this case, α is not an integer and consequently the exact choice of k for integer α does not matter, see Remark 3.5. α (S) with C p 3.3. Projectors. For the study of sharp maximal functions (3.3), it is useful to construct for every cube Q ⊂ Rn a projection operator PQ from L1(Q ∩ S) onto the subspace Pk−1(Rn)Q∩S, k ∈ N, such that Ek(f, Q)Lu(S) ≈ (H s(Q ∩ S))−1/ukf − PQf kLu(Q∩S). This is possible due to the following property of polynomials. Proposition 3.6. Let S be an s-set with n − 1 < s ≤ n and 1 ≤ q, u ≤ ∞. Then for every polynomial p of degree k and every cube Q centered at S, we have (cid:18)ZQ∩S pq dH s(cid:19)1/q ≤ c(cid:18)ZQ∩S pu dH s(cid:19)1/u , (3.7) where the constant c > 1 depends on n, k and S. 6 See Proposition 3 on p. 36 in [12] for the proof. See also [5], where a more general inequality of such kind is proved. Actually, the reverse Holder inequality for polynomials (3.7) guarantees that the maximal functions f ♯ α,S have most of the properties of their counterparts defined on Rn. In particular, we use it to show that, in the definition of the space C p α,S can be replaced with f ♯ α(S), the function f ♯ α,u,S, 1 < u ≤ p, without changing the space. Recall that QS = Q ∩ S. We fix now one more notation, namely, for a cube Q and a function f ∈ Lu(S, H s), 1 ≤ u ≤ ∞, we denote Ek(f, Q)Lu(S) := inf p∈Pk−1(cid:18)ZQS f − pu dH s(cid:19)1/u . Proposition 3.8. Let k ∈ N and Q be a cube centered at S. Then there exists a linear operator PQ : L1(QS) → Pk−1 such that for every 1 ≤ u ≤ ∞ and every f ∈ Lu(S) (cid:18)ZQS f − PQf u dH s(cid:19)1/u ≤ cEk(f, Q)Lu(S), with some constant c independent of Q. Proof. Following the construction of PQ from [20], let {pβ : β ≤ k −1} denote an orthonormal basis in the linear space Pk−1 with respect to the inner product hf, gi =ZQS f g dH s. (3.9) Note that since s > n−1, formula (3.9) defines an inner product indeed. Set PQf := Xβ≤k−1(cid:18)ZQS f pβ dH s(cid:19)pβ. We estimate the operator norm of PQ in Lu norm. For every f ∈ Lu(QS), we have By the Holder inequality, kPQf kLu(QS) ≤ Xβ≤k−1(cid:12)(cid:12)(cid:12)(cid:12) kPQf kLu(QS) ≤(cid:18) Xβ≤k−1 ZQS kpβkLu(QS ). f pβ dH s(cid:12)(cid:12)(cid:12)(cid:12) kpβkLu(QS)kpβkLu′ (QS)(cid:19)kf kLu(QS ), and by Proposition 3.6, kpβkLu(QS)kpβkLu′ (QS) ≤ c(cid:0)(H s(QS)) 1 u − 1 2 kpβkL2(QS)(cid:1)(cid:0)(H s(QS)) 7 1 u′ − 1 2 kpβkL2(QS)(cid:1) = c. Hence kPQf kLu(QS) ≤ ckf kLu(QS). Now, let pQ denote a polynomial of degree k − 1 satisfying (cid:18)ZQS f − pQu dH s(cid:19)1/u = Ek(f, Q)Lu(S). Then we can write f − PQf = (f − pQ) − PQ(f − pQ) and, consequently, we get the estimate (cid:16)ZQS f − PQf u dH s(cid:17)1/u ≤ (1 + kPQf kLu(QS ))Ek(f, Q)Lu(S) ≤ cEk(f, Q)Lu(S). (cid:3) The proposition above together with the definition of the sharp maxi- mal function (3.3) implies that f ♯ α,u,S(x) ≈ sup t>0 1 tα(cid:18)ZQ(x,t)∩S f − PQ(x,t)f u dH s(cid:19)1/u . (3.10) Now we consider some properties of the projectors PQ. Lemma 3.11. Let function f ∈ L1 be centered at x ∈ S, then: loc(Q∩S) and cube Q = Q(x, r) ⊂ Rn (1) PQ(λ) = λ for any λ ∈ R; (2) PQf (y) ≤ cf Q∩S, y ∈ Q ∩ S; (3) If Q′ centered at S is such that Q′ ⊂ Q and H s(Q′ ∩ S) ≥ cH s(Q ∩ S), then PQf (z) − PQ′f (z) ≤ cZQ∩S f − PQ dH s, z ∈ Q′ ∩ S; (4) If Q′ = Q(y, r), y ∈ S, such that Q′ ∩ Q 6= ∅, then PQf (z1) − PQ′f (z1) ≤ cZQ(z2,2r)∩S for every z1, z2 ∈ Q ∩ Q′ ∩ S. 8 f − PQ(z2,2r)f dH s Proof. Properties (1) and (2) directly follow from the construction of projectors PQ. Let us prove (3). By (3.7), we have sup Q′ S PQf − PQ′f dH s S PQf − PQ′f ≤ cZQ′ ≤ c(cid:20)ZQ′ ≤ c(cid:20)ZQS S f − PQ′f dH s(cid:21) f − PQf dH s +ZQ′ f − PQf dH s + Ek(f, Q′)L1(S)(cid:21) ≤ cZQS S f − PQf dH s. Note that if cubes Q = Q(x, r) and Q′ = (y, r′) are such that Q∩Q′ 6= ∅ then Q, Q′ ⊂ Q(z2, 2r) for any z2 ∈ Q ∩ Q′. Then, since PQf (z1) − PQ′f (z1) ≤PQf (z1) − PQ(z2,2r)f (z1) + PQ′f (z1) − PQ(z2,2r)f (z1), the statement (4) easily follows from (3). (cid:3) Remark 3.12. If x ∈ S is a Lebesgue point of a function f ∈ L1 then, by definition, loc(S), lim r→0ZQ(x,r)∩S f − f (x) dH s = 0. (3.13) By statements (1) and (2) of Lemma 3.11, we have PQ(x,r)f (x) − f (x) = PQ(x,r)[f − f (x)](x) ≤ cZQ(x,r)∩S f − f (x)dH s. Since almost every point of S is a Lebesgue point of a function f ∈ L1 loc(S) (see for example [16]), we have lim r→0 PQ(x,r)f (x) = f (x) a.e. on S. (3.14) The following lemma is a special case of Theorem 1 in [17] (see also Theorem 1 in [18]). For the sake of completeness, we will sketch the proof here. Lemma 3.15. Suppose that α > 0, q ≥ 1 and f ∈ L1 any cube Q = Q(x, r), x ∈ S, we have loc(S). Then for (cid:18)ZQ∩S f − PQf q dH s(cid:19)1/q ≤ crα(cid:18)Z2Q∩S (f ♯ α,S)σ dH s(cid:19)1/σ , (3.16) where 1 σ = 1 q + α s . Proof. Let x0 be a Lebesgue point of a function f . We will show that f (x0)−PQ(x0,r)f (x0) ≤ crα(cid:18)f ♯ α,S(x0)(cid:19)1−ασ/s(cid:18)ZQS(x0,r) 9 (f ♯ α,S)σ dH s(cid:19)α/s . (3.17) For every cube Q(x, t) ⊂ Rn, let QS(x, t) denote the set Q(x, t) ∩ S and consider u(x, t) = f − PQ(x,t)f dH s, 1 tα ZQS(x,t) By Proposition 3.8 and (3.2), we have u(x, τ ) ≤ cu(x, t) if τ ≤ t ≤ 2τ. (3.18) If Qk = Q(x0, 2−kr), k ≥ 0, then by (3.14), Lemma 3.11 and (3.18) respectively, we obtain ∞ (PQk+1f (x0) − PQkf (x0))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (x0) − PQ(x0,r)f (x0) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=0 Xk=0 2−kαu(x0, 2−kr) ≤ crα ∞ tαu(x0, t) dt t 0 ≤ crαu(x0, r) + cZ r α,S)σ dH s(cid:19)1/σ (f ♯ I =(cid:18)ZQS(x0,r) Let and consider two cases: (1) If f ♯ α,S(x0) ≤ I then by (3.10), we have . (3.19) (3.20) tα−1u(x0, t) dt ≤ crαf ♯ α,S(x0) ≤ crα(f ♯ α,S(x0))1− ασ s I ασ s . Z r 0 (2) If f ♯ α,S(x0) > I, then define τ = rI σ/s(f ♯ α,S)−σ/s < r and write Z r 0 tα−1u(x0, t) dt =(cid:18)Z τ 0 Then +Z r τ (cid:19)tα−1u(x0, t) dt ≡ I1 + I2. I1 ≤ cτ αf ♯ α,S(x0) = crα(f ♯ α,S(x0))1− ασ s I ασ s . To estimate I2, note that for every t ≤ r, we have u(x0, t) ≤ c(cid:18)ZQS(x0,t) and therefore (f ♯ α,S)σ dH s(cid:19)1/σ t(cid:19)s/σ I, (3.21) I2 ≤ cIrs/σZ r τ tα−1−s/σ dt ≤ cI(cid:18) r τ α ≤ crαI, Consequently, we have the same estimate as in case (1) for the integral in (3.19). 10 ≤ c(cid:18)r τ(cid:19)s/σ To estimate u(x0, r), we use (3.21) with t = r. Thus u(x0, r) = [u(x0, r)]1− ασ s [u(x0, r)] which finishes the proof of (3.17). ασ s ≤ [f ♯ α,S(x0)]1− ασ s I ασ s , Now consider (cid:18)ZQS f − PQf q dH s(cid:19)1/q ≤(cid:18)ZQS +(cid:18)ZQS f (y) − PQ(y,r)f (y)q dH s(y)(cid:19)1/q PQ(y,r)f (y) − PQf (y)q dH s(y)(cid:19)1/q ≡I1 + I2. By (3.17), we have I1 ≤ crα(cid:18)ZQS (f ♯ α,S(y))q(1−ασ/s)(cid:16)ZQS(y,r) (f ♯ α,S)σ dH s(cid:17)qα/s dH s(y)(cid:19)1/q , and since for every y ∈ Q(x, r), the cube Q(y, r) ⊂ Q(x, 2r), we have I1 ≤ crα(cid:18)ZQS(x,2r) ≤ crαZQS(x,2r) (f ♯ (f ♯ α,S)σ dH s(cid:19)α/s(cid:18)ZQS α,S)σ dH s(cid:19)1/σ . (f ♯ α,S)σ dH s(y)(cid:19)1/q If y ∈ Q(x, r)∩S and z ∈ Q(x, 2r)∩S, then by statement (4) of Lemma 3.11, we have PQ(y,r)f (y) − PQf (y) ≤ cZQS(x,2r) f − PQ(x,2r)f dH s ≤ crαf ♯ α,S(z). Since the last inequality holds for any z ∈ QS(x, 2r), we have PQ(y,r)f (y) − PQf (y) ≤ crα(cid:18)ZQS(x,2r) (f ♯ α,S)σ dH s(cid:19)1/σ for every y ∈ QS. This completes the proof. (3.22) (cid:3) Applying the Holder inequality and Lemma 3.15 respectively, we get the following statement. Lemma 3.23. Let α > 0, u > 1 and f ∈ L1 loc(S), then f ♯ α,S ≤ f ♯ α,u,S(x) ≤ cMσ(f ♯ α,S)(x), (3.24) where 1/σ = 1/u + α/s, M is the Hardy-Littlewood maximal operator and Mσ(g) = [M(gσ)]1/σ. 11 Remark 3.25. Recall that the function space C p α(S) is defined as the set of functions f ∈ Lp(S) such that f ♯ α,1,S ∈ Lp(S). By Lemma 3.23 and the Lq-boundedness of the maximal operator for q > 1, the set of functions such that f ♯ α,u,S ∈ Lp(S) is independent of u as long as 1 ≤ u ≤ p. Thus, we can use any value of u in the definition of the space C p α(S), it is enough to consider local best approximations on cubes with side length less than any fixed positive number. α(S). Furthermore, the next lemma shows that to define C p Lemma 3.26. Let p > 1, 1 ≤ u ≤ p and γ > 0. Then C p with the space α(S) coincides {f ∈ Lp(S) : sup 0<t<γ t−αEk(f, Q(·, t))Lu(S) ∈ Lp(S)}. Proof. First let 1 ≤ u < p. For every x ∈ S, we have 1 tα Ek(f, Q(x, t))Lu(S) ≤ γ−α sup sup t≥γ ≤ c(M(f u)(x))1/u t≥γ (cid:18)ZQ(x,t)∩S f u dH s(cid:19)1/u (3.27) and the claim follows from the Lq-boundedness of the maximal operator for q = p/u > 1. If u = p, we take some 1 ≤ q < p. By (3.24), we have kf ♯ α,p,SkLp(S) ≤ ckf ♯ α,q,SkLp(S) ≤ k sup 0<t<γ 1 tα Ek(f, Q(x, t))Lq(S)kLp(S) + k sup 1 tα Ek(f, Q(x, t))Lq(S)kLp(S), where the first summand is bounded by the assumption and the second by (3.27). (cid:3) t≥γ 4. Comparison with Besov spaces The following theorem is the main result of the present paper. Theorem 4.1. Let S be an s-set with n − 1 < s ≤ n, 1 < p ≤ ∞ and α be a non-integer positive number. Then Bp,p α (S) ⊂ C p α(S) ⊂ Bp,∞ α (S). (4.2) Remark 4.3. If α > 0 is an integer then in the embeddings (4.2) the space C p α(S) shall be replaced with the space {f ∈ Lp(S) : f ♭ α,S ∈ Lp(S)}, see Remark 3.5 for the difference between functions f ♯ α,S and f ♭ α,S. 12 Theorem 4.1 is an analogue of Theorem 7.1 in [9] for S = Rn. Examples similar to the ones constructed in [9] show that the embeddings (4.2) are the best possible within the scale of Besov spaces. The case of s-set with s strictly less than n is of our current interest due to the characterization for traces of potential spaces to s-sets given by A. Jonsson, see Theorem 2.4. Corollary 4.4. Let S be an s-set, n − 1 < s < n, 1 < p < ∞, k ∈ N and α = k − (n − s)/p > 0. Then for any 0 < ε < (n − s)/p C p α+ε(S) ⊂ W p k (Rn)S ⊂ C p α(S). (4.5) To prove Theorem 4.1, we need the following representation for the norm of Besov spaces. Theorem 4.6. Let S be an s-set, n − 1 < s ≤ n, α > 0, 1 ≤ p, q ≤ ∞ and k = [α] + 1. Then, when q < ∞, we have kf kBp,q α (S) ≈ kf kLp(S) +(cid:18)Z 1 0 (cid:18)kEk(f, Q(·, t))Lp(S)kLp(S) tα t (cid:19)1/q (cid:19)q dt and kf kBp,∞ α (S) ≈ kf kLp(S) + sup 0<t≤1 t−αkEk(f, Q(·, t))Lp(S)kLp(S). Remark 4.7. Such characterization of Besov spaces is fairly standard, see for example [4], [21] for the case when S = Rn and [12], [20] for the case of n-sets. Proof. First, suppose that the right-hand-side is finite. We note that by (3.1), we can replace the integral by the sum ∞ Xν=0 2ναq(cid:18)ZS E p k (f, Q(x, 2−ν))Lp(S) dH s(x)(cid:19)q/p . (4.8) Take a net π = {Qi, i = 1, 2, . . . } with mesh size 2−ν and let Pπf be a function from Pk(π) which will be chosen later. Clearly, ZS f −Pπf p dH s =XQ∈πZQ∩S f −Pπf p dH s = XQ∈π′ZQ∩S f −Pπf p dH s, where π′ = {Q ∈ π : Q ∩ S 6= ∅}. Set t = 2−ν−1. For any cube Q = Q(x, t) from π′, choose a point y ∈ Q ∩ S and set K = Q(y, 2t). Then Q ⊂ K and χK ≤ c, XQ∈π′ where constant c depends only on n. 13 The center of every cube K is in S. Hence, by Proposition 3.8, there is a projector PK : L1(K ∩ S) → P[α] such that ZK∩S f − PKf p dH s ≤ cH s(K ∩ S)E p k (f, K)Lp(S), with constant c independent of f and K. Define Pπf (x) = PKf (x), Q. For any point z ∈ K ∩ S we have K ⊂ Q(z, 4t) and x ∈ Q, and Pπf (x) = 0 if x /∈ SQ∈π′ f − Pπf p dH s =ZQ∩S ≤ZK∩S ZQ∩S f − PKf p dH s f − PKf p dH s ≤ cH s(K ∩ S)E p ≤ cH s(K ∩ S)E p k (f, K)Lp(S) k (f, Q(z, 4t))Lp(S), where the last inequality holds by (3.1). Then we integrate the inequal- ity over the set K ∩ S to obtain ZQ∩S f − Pπf p dH s ≤ cZK∩S E p k (f, Q(z, 4t))Lp(S) dH s(z). Remember that we set t = 2−ν−1. Thus we have (cid:18)ZS f − Pπf p dH s(cid:19)1/p ≤ c XQ∈π′ZK∩S ≤ c(cid:18)ZS E p E p k (f, Q(z, 2−ν+1))Lp(S) dH s(z)!1/p k (f, Q(·, 2−ν+1))Lp(S) dH s(cid:19)1/p . Let now cν be equal to the last integral multiplied by 2να. Then ∞ Xν=1 cq ν = c ∞ Xν=1 2ναq(cid:18)ZS E p k (f, Q(x, 2−ν))Lp(S) dH s(x)(cid:19)q/p < ∞, so that, by (4.8), f ∈ Bp,q holds. α (S) and the wanted estimate for its norm Suppose now that f ∈ Bp,q α (S) and π is a net with mesh size 2t, t > 0. Denote by π′ a family of all cubes Q from π such that Q ∩ S 6= ∅. If Q ∈ π′ and x ∈ Q∩S, then Q(x, t) ⊂ 2Q, H s(Q(x, t)∩S) ≈ H s(2Q∩S) and by (3.1) E p k (f, Q(x, t))Lu(S) ≤ c E p k (f, 2Q)Lu(S). 14 Hence, ZS E p k (f, Q(x, t))Lu(S) dH s(x) =XQ∈π′ZQ∩S ≤ cXQ∈π′ E p k (f, Q(x, t))Lu(S) dH s(x) H s(2Q ∩ S)E p k (f, 2Q)Lu(S). It is easy to see that family of cubes π = {2Q : Q ∈ π′} can be represented as π = ∪m i=1πi, where m = 2n and every πi is a subfamily of a net with mesh size 4t. Set k = [α] + 1 and t = 2−ν, ν = 2, . . . . Since f ∈ Bp,q functions Pπif ∈ Pk−1, i = 1, . . . , m, such that α (S), there are ZS E p k (f, Q(x, 2−ν))Lu(S) dH s(x) m f − Pπif u dH s ≤ c ≤ c sup = c sup k (f, Q)Lp(S) H s(Q ∩ S)E p f − Pπif p dH s, (H s(Q ∩ S))1−p/pZQ∩S Xi=1 XQ∈πi πi XQ∈πi πi XQ∈πiZQ∩S πi ZS k (f, Q(x, 2−ν))Lu(S) dH s(x)(cid:19)q/p E p ≤ c sup f − Pπif p dH s ≤ c2(−ν+2)αpcp ν−2 and, consequently ∞ Xν=2 2ναq(cid:18)ZS ≤ c ∞ Xν=0 cq ν < ∞. (cid:3) Proof of Theorem 4.1. We start with the first embedding and use here the characterization of the spaces C p α(S) given by Lemma 3.26. By property (3.1) of local best approximation, we have 1 tαp E p k (f, Q(x, t))Lp(S) ≤ c ∞ Xν=1 ≤ cZ 1 0 2−ναpE p k (f, Q(x, 2−ν))Lp(S) E p k (f, Q(x, t))Lp(S) tαp dt t . sup 0<t≤ 1 2 Thus, kf ♯ α,Skp Lp(S) ≤ cZSZ 1 = cZ 1 0 E p k (f, Q(x, t))Lp(S) tαp dt t dx 0 (cid:18)kEk(f, Q(·, t))Lp(S)kLp(S) tα 15 (cid:19)p dt t . For non-integer α > 0 the number k = −[−α] is strictly greater than α, hence, by Theorem 4.6 the last term can be estimated by kf kBp,p α (S). To prove the second embedding, we notice that for every α > 0, k ∈ N and t > 0, we have kEk(f, Q(·, t))Lp(S)kLp(S) tα ≤ k sup t>0 Ek(f, Q(·, t))Lp(S) tα kLp(S). (4.9) Setting k − [−α] and taking the supremum over the interval (0, 1] in (4.9) we get kf kBp,∞ α (S) ≤ kf ♯ α,SkLp(S). (cid:3) Since the statements of Lemmata 3.15, 3.23, 3.26 hold true for the sharp maximal functions f ♭ α,S as well, the case of integer α can be treated with the slight modification of the last proof; see Remark 4.3. 5. Sobolev spaces on s-sets As mentioned above, the definition (3.4) of the function space C p k (S) yields the Sobolev space W k,p(Rn) if S = Rn [7] or the space W k,p(S) if S is a W k,p-extension domain [14]. Motivated by these facts, one could ask a natural question: Can the spaces C p k(S) be relevant analogs of classical Sobolev spaces in some more general settings? If S is an n-set, then C p k(S) is the trace space of W k,p(Rn) to S [20] and therefore functions from C p k (S) possess certain distinctive properties In the case of s-sets with n − 1 < s < n, we of Sobolev spaces. can not derive the corresponding properties from the trace reasoning. Nevertheless, some results which are known for Sobolev spaces of higher order can be obtained. In particular, we have a version of Sobolev- Poncar´e inequality given by Lemma 3.15, and an analogue of Sobolev embedding theorem which is proved below. Proposition 5.1. Let S be an s-set, n − 1 < s < n, p ≥ 1, kp < s and q = sp/(s − kp). Then kf kLq(S) ≤ c(kf ♯ α,SkLp(S) + (diam S)−αkf kLp(S)) (5.2) Proof. By Lemma 3.15 and by statement (2) from Lemma 3.11 (cid:18)ZQS f q dH s(cid:19)1/q ≤(cid:18)ZQS ≤ c(cid:20)rα(cid:18)Z2QS f − PQf q dH s(cid:19)1/q α,S)p dH s(cid:19)1/p (f ♯ +(cid:18)ZQS +ZQS PQf q dH s(cid:19)1/q f dH s(cid:21). Choosing Q = Q(x, diam S), where x is any point in S, and using the s-regularity of S, we get (5.2). (cid:3) 16 For the first order Sobolev space, the definition in terms of Lp-properties of sharp maximal functions makes sense also in general situation of a metric measure space [13]. So far, very little is known about the higher order case. The problem is that in the definition of the sharp maximal function, we need a family of polynomials with special properties. In case of s-sets in Rn, n − 1 < s ≤ n, such families do exist, see Sec- tion 3. In more general situation, one could use the related technique assuming that some polynomial type functions exist. This kind of an approach is used for example in [19], where a version of polynomials on metric spaces equipped with a doubling measure have been proposed. See also [2], where a characterization of higher order Sobolev spaces via a quadratic multiscale expression is used to propose an other definition for Sobolev functions on any metric space. References [1] D. R. Adams and L. I. Hedberg, Function spaces and potential theory, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Vol. 314 (Springer-Verlag, Berlin, 1996). [2] R. Alabern, J. Mateu, and J. Verdera, A new characterization of Sobolev spaces on Rn, Preprint, arXiv:1011.0667v2 (2010). [3] O. V. Besov, V. P. Il'in, and S. M. Nikol'ski, Integral representations of func- tions and imbedding theorems. Vol. I -- II (V. H. Winston & Sons, Washington, D.C., 1978 -- 1979). [4] Y. Brudnyi, Sobolev spaces and their relatives: local polynomial approxima- tion approach, in: Sobolev spaces in mathematics. II, , Int. Math. Ser. (N. Y.), Vol. 9 (Springer, New York, 2009), pp. 31 -- 68. [5] A. Brudnyi and Y. Brudnyi, Remez type inequalities and Morrey-Campanato spaces on Ahlfors regular sets, Interpolation theory and applications, Con- temp. Math., Vol. 445 (Amer. Math. Soc., Providence, RI, 2007), pp. 19 -- 44. [6] A. P. Calder´on, Lebesgue spaces of differentiable functions and distributions, Proc. Sympos. Pure Math., Vol. IV (Amer. Math. Soc., Providence, R.I., 1961), pp. 33 -- 49. [7] A. P. Calder´on, Estimates for singular integral operators in terms of maximal functions, Studia Math. 44, 563 -- 582 (1972). [8] A. P. Calder´on and R. Scott, Sobolev type inequalities for p > 0, Studia Math. 62, 75 -- 92 (1978). [9] R. A. DeVore and R. C. Sharpley, Maximal functions measuring smoothness, Mem. Amer. Math. Soc. 47, 1 -- 115 (1984). [10] A. Jonsson, The trace of potentials on general sets, Ark. Mat. 17, 1 -- 18 (1979). [11] A. Jonsson and H. Wallin, A Whitney extension theorem in Lp and Besov spaces, Ann. Inst. Fourier (Grenoble) 28, 139 -- 192 (1978). [12] A. Jonsson and H. Wallin, Function spaces on subsets of Rn, Math. Rep. 2, 1 -- 221 (1984). [13] P. Haj lasz and J. Kinnunen, Holder quasicontinuity of Sobolev functions on metric spaces, Rev. Mat. Iberoamericana 14, 601 -- 622 (1998). [14] P. Haj lasz, P. Koskela, and H. Tuominen, Sobolev embeddings, extensions and measure density condition, J. Funct. Anal. 254, 1217 -- 1234 (2008). [15] P. Haj lasz and O. Martio, Traces of Sobolev functions on fractal type sets and characterization of extension domains, J. Funct. Anal. 143, 221 -- 246 (1997). 17 [16] J. Heinonen, Lectures on analysis on metric spaces, Universitext (Springer- Verlag, New York, 2001). [17] E. V. Ignat'eva, An inequality of Sobolev-Poincar´e type on metric spaces in terms of sharp-maximal functions, Mat. Zametki 81, 140 -- 144 (2007). [18] I. A. Ivanishko and V. G. Krotov, Generalized Poincar´e-Sobolev inequality on metric spaces, Trudy Instituta Matematiki NAS Belarusi 14 (2006), in Russian. [19] Y. Liu, G. Lu, and R. L. Wheeden, Some equivalent definitions of high or- der Sobolev spaces on stratified groups and generalizations to metric spaces, Math. Ann. 323, 157 -- 174 (2002). [20] P. Shvartsman, Local approximations and intrinsic characterization of spaces of smooth functions on regular subsets of Rn, Math. Nachr. 279, 1212 -- 1241 (2006). [21] H. Triebel, Theory of function spaces. II, Monographs in Mathematics, Vol. 84 (Birkhauser Verlag, Basel, 1992). Addresses: L.I.: Department of Mathematics, P.O. Box 11100, FI-00076 Aalto University, Finland. E-mail: [email protected] R.K.: Department of Mathematics and Statistics, P.O. Box 68, FI- 00014 University of Helsinki, Finland. E-mail: [email protected] 18
1511.09045
1
1511
2015-11-29T17:11:49
Stein Domains in Banach Algebraic Geometry
[ "math.FA", "math.AG", "math.NT" ]
In this article we give a homological characterization of the topology of Stein spaces over any valued base field. In particular, when working over the field of complex numbers, we obtain a characterization of the usual Euclidean (transcendental) topology of complex analytic spaces. For non-Archimedean base fields the topology we characterize coincides with the topology of the Berkovich analytic space associated to a non-Archimedean Stein algebra. Because the characterization we used is borrowed from a definition in derived geometry, this work should be read as a contribution towards the foundations of derived analytic geometry.
math.FA
math
STEIN DOMAINS IN BANACH ALGEBRAIC GEOMETRY FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Abstract. In this article we give a homological characterization of the topology of Stein spaces over any valued base field. In particular, when working over the field of complex numbers, we obtain a characterization of the usual Euclidean (transcendental) topology of complex analytic spaces. For non-Archimedean base fields the topology we characterize coincides with the topology of the Berkovich analytic space associated to a non-Archimedean Stein algebra. Because the characterization we used is borrowed from a definition in derived geometry, this work should be read as a contribution towards the foundations of derived analytic geometry. Contents Introduction 1. 1.1. Notation 2. Bornological Algebraic Geometry 2.1. Quasi-abelian categories, bornological spaces and dagger analytic geometry 2.2. Bornological vector spaces 2.3. Tensor product of unbounded complexes 3. Some results in functional analysis 3.1. Proper bornological vector spaces 3.2. Relations between bornological and topological vector spaces 3.3. Nuclear bornological spaces and flatness 3.4. A relative flatness lemma 3.5. Strict exact sequences in bornological and topological settings 3.6. Derived functors of the inverse limit functor 4. Stein domains 4.1. Bornological Fr´echet algebras 4.2. Stein algebras and Stein spaces 2 3 4 4 8 9 11 11 14 16 24 28 30 34 34 36 The first author acknowledges the support of the University of Padova by MIUR PRIN2010-11 “Arith- metic Algebraic Geometry and Number Theory”, and the University of Regensburg with the support of the DFG funded CRC 1085 “Higher Invariants. Interactions between Arithmetic Geometry and Global Analy- sis” that permitted him to work on this project. The second author acknowledges the University of Oxford and the support of the European Commission under the Marie Curie Programme for the IEF grant which enabled this research to take place. The contents of this article reflect the views of the authors and not the views of the European Commission. We would like to thank Konstantin Ardakov, Francesco Baldassarri, Elmar Grosse-Klonne, Fr´ed´eric Paugam, and J´erome Poineau for interesting conversations. 1 2 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER 5. Homological characterization of the Stein topology References 43 51 1. Introduction This paper can be thought as a continuation of [5] and [4], on which we build our main results. The problem addressed in the present work is to extend the approach proposed in [5] and [4] to Stein domains of analytic spaces. More precisely, the main theorems of this paper prove that the characterization of open embeddings with the homotopy monomorphism property, given in [5] for affinoid domains and in [4] for dagger affinoid domains, extends to Stein domains. Since Stein spaces are defined as spaces which have a suitable exhaustion by (dagger) affinoid subdomains, the strategy of our proofs is to rely on the results of [5] and [4], and deduce the theorems for Stein spaces using their exhaustions. In particular, this strategy naturally leads to consider projective limits of bornological spaces and the issue about the commutation of the bornological projective tensor product with projective limits. We devote Section3 to provide basic results about these issues for which it seems that there is no literature available. Our main difficulty is the fact that neither the projective tensor product nor the injective tensor product of bornological vector spaces commutes with projective limits in general. This is also the main obstacle to the generalization of our results to more general notion of domains, such as dagger quasi-Stein domains. The paper is organized as follows: in Section2 we recall the main results from [4] that will be used, we fix the notation and we introduce some key notions that will be used in our main proofs. Section3 contains our technical results in functional analysis that are required to prove Theorems 5.5 and 5.7. In Section3.1 we describe the class of proper bornological spaces and we prove its main property in Proposition 3.15, which shows that the closure of a subspace of a proper bornological space is equal to the set of its (bornological) limit points. Section 3.3 begins with the definition of nuclear bornological spaces and the study of their main properties. We remark that, although proper and nuclear bornological vector spaces have been previously considered in literature, see for example [19] and [21], this is the first time, at our knowledge, that a detailed study of their property is done also for non- Archimedean base fields. The main result of Section3.3 is Theorem 3.50 which shows that nuclear bornological vector spaces are flat in CBornk with respect to its natural monoidal structure, i.e. counterpart of the well-known result of the theory of locally convex spaces. The rest of Section3 deals with projective limits of bornological vector spaces: In particular, Section 3.4 addresses the issue of commutation of projective limits with the bornological complete projective tensor product; Section 3.5 deals with a bornological version of the Mittag-Leffler Lemma for Fr´echet spaces and Section 3.6 contains some lemmas about the computation of the derived functor of the projective limit functor for quasi-abelian categories. Section 4 starts by providing some results on Fr´echet bornological algebras and then it continues by giving the definition of Stein spaces and Stein algebras suitable for our context. The main result of this section, Theorem 4.25, is a generalization of Forster’s Theorem about the anti-equivalence between the category of complex Stein algebras and complex Stein spaces, the endofunctor p´qpbkF is exact if F is nuclear. This is a bornological 3 to arbitrary base fields. Section 5 contains our main results. We characterize the open embeddings of Stein spaces by the maps having the homotopy monomorphism property (see Theorem 5.5 and Theorem 5.7). Giving a disjoint union of Stein spaces mapping to a fixed Stein space, we characterize in Theorem 5.13 the surjectivity of such a morphism by a conservativity property for transversal modules. See Definition 2.4 for the notion of transversal module, called a RR-quasicoherent module, after the work of Ramis and Ruget [35]. The homological treatment of holomorphic functional calculus, as developed by Taylor, leads naturally to derived methods (see for example [37]). Let pC,b, eCq be a closed sym- metric monoidal elementary quasi-abelian category with enough flat projectives. In [6] a Grothendieck topology of (homotopy) Zariski open immersions in CommpsCqop is defined where sC is the closed symmetric monoidal model category of simplicial objects in C. The homotopy monomorphism condition that we use in this article (in the case that C “ CBornk or IndpBankq) is a restriction of Definition 1.2.6.1 (3) of [41] from the homotopy category of CommpsCqop to the opposite category of dagger Stein algebras (thought of as constant sim- plicial objects). The model structure to be explained in [6] is compatible in a natural way with the quasi-abelian structure on C. This allows us to relate our work to the work of Toen and Vezzosi from [40, 41], as shown in [6]: CommpsCqop satisfies their axioms on a monoidal model category so according to their approach one can do derived geometry relative to it, and apply their results. In the case that our base field is the complex numbers, some of our results are already present in the work of Pirkovskii, for instance [28]. 1.1. Notation. The notation used here will be totally consistent with the notation of [4], which is the following: ‚ If C is a category we will use the notation X P C to indicate that X is an object of C. ‚ If C is a category then IndpCq will denote the category of Ind-objects of C. ‚ k will denote a field complete with respect to a fixed non-trivial valuation, Archimedean or non-Archimedean. ‚ Vectk is the closed symmetric monoidal category of vector spaces (with no extra structure) over k. ‚ SNrmk the category of semi-normed modules over k, remarking that, if not otherwise stated, by a semi-normed space over non-Archimedean base field we mean a k-vector space equipped with a non-Archimedean semi-norm. ‚ Nrmk the category of normed modules over k. ‚ Bank the category of Banach modules over k. ‚ For V P SNrmk, V s “ V {p0q is the separation and pV P Bank is the separated comple- tion. ‚ Bornk the category of bornological vector spaces of convex type over k and CBornk the category of complete bornological vector spaces of convex type over k. ‚ For E P Bornk and B a bounded absolutely convex subset, (i.e. a bounded disk) of E, EB is the linear subspace of E spanned by elements of B equipped with the gauge 4 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER functor in a category. functor in a category. k denotes the category of dagger affinoid algebras over k. E denotes the category of bounded absolutely convex subsets B of semi-norm (also called the Minkowski functional) defined by B (see Remark 3.40 of [4] for a review of the notion of gauge semi-norm). ‚ For E P Bornk, DE denotes the category of bounded absolutely convex subsets of E. ‚ Afnd: ‚ For E P Bornk, Dc E for which EB P Bank. ‚ The notation limÑ refers to a colimit (also known as inductive or direct limit) of some ‚ The notation limÐ refers to a limit (also known as projective or inverse limit) of some ‚ For polyradii ρ “ pρiq P Rn`, the notation ρ ă ρ1 means that ρ and ρ1 have the same number of components and every component of ρ is strictly smaller than the corresponding component of ρ1. ‚ With the notation CD we will denote the category of covariant functors D Ñ C. In particular, if I is a filtered set we will denote CI the category of functors I Ñ C, when I is thought as a category. ‚ A cofiltered projective system tEiuiPI of objects of a category is said to be epimorphic if for any i ă j the system map Ej Ñ Ei is an epimorphism. Similarly, a filtered direct system tEiuiPI is said to be monomorphic if for any i ă j the system map Ei Ñ Ej is a monomorphism. 2. Bornological Algebraic Geometry 2.1. Quasi-abelian categories, bornological spaces and dagger analytic geometry. We suppose that the reader is familiar with the theory of quasi-abelian categories as devel- oped in [36]. In this section pC,b, eCq will be a closed symmetric monoidal quasi-abelian category and Hom will denote the internal hom functor. To any closed symmetric monoidal category is associated a category of commutative monoids, denoted CommpCq and a category of affine schemes AffpCq “ CommpCqop. The duality functor CommpCq Ñ AffpCq is denoted by spec. To any A P CommpCq we can associate the category of A-modules ModpAq, which is quasi-abelian closed symmetric monoidal with respect to a bifunctor bA, naturally induced by b. Moreover, since ModpAq is quasi-abelian we can always associate to A the derived cat- egories of ModpAq, denoted DpAq, and using the left t-structure of DpAq we define Dď0pAq, Dě0pAq and DbpAq. Notice also that in Proposition 2.1.18 (c) of [36] it is shown that if C is elementary quasi-abelian ModpAq is elementary quasi-abelian. Definition 2.1. A morphism specpBq Ñ specpAq is said to be a homotopy monomorphism if the canonical functor Dď0pBq ÝÑ Dď0pAq is fully faithful. In a dual way, we say that the correspondent morphism of monoids A Ñ B is a homotopy epimorphism. The following characterization of homotopy monomorphisms is the useful one for pratical purposes. Lemma 2.2. Assume that p : specpBq Ñ specpAq is a morphism in AffpCq and that the functor ModpAq Ñ ModpBq given by tensoring with B over A is left derivable to a functor Dď0pAq Ñ Dď0pBq. Then, p is a homotopy monomorphism if and only if BbL AB – B. Proof. See Lemma 2.24 of [4]. l Lemma 2.3. Let f : specpAq Ñ specpBq, g : specpBq Ñ specpCq be two morphisms of affine schemes such that g f and g are homotopy monomorphisms, then also f is a homotopy monomorphism. Proof. The hypothesis mean that we have a diagram of functors 5 Dď0pAq f pg fq g Dď0pCq Dď0pBq such that g is fully faithful and g f is fully faithful. Hence for any V, W P Dď0pAq HomDď0pAqpV, Wq – HomDď0pCqppg fqpV q,pg fqpWqq – HomDď0pBqpfpV q, fpWqq which precisely means that f is fully faithful. l We recall the following notion from [4] and [5]. Definition 2.4. Consider an object A P CommpCBornkq. We define a sub-category ModRRpAq AB Ñ of ModpAq whose modules M satisfy the property that the natural morphism MpbL MpbAB is an isomorphism in Dď0pBq, for all homotopy epimorphisms A Ñ B. We call these modules RR-quasicoherent modules. Homotopy epimorphisms are the morphisms that we use to endow CommpCq with a Grothendieck topology. The following definition is based on definitions in [39] and [41]. Definition 2.5. Consider a full sub-category A Ă AffpCq such that the base change of a homotopy monomorphisms in A is a homotopy monomorphism. On A we can define the homotopy Zariski topology which has as its covers collections tspecpBiq Ñ specpAquiPI where there exists a finite subset J Ă I such that ‚ for each i P J, the morphism A Ñ Bi is of finite presentation and the resulting morphisms Dď0pBiq Ñ Dď0pAq is fully faithful; ‚ a morphism in ModRRpAq is an isomorphism if and only if it becomes an isomorphism in each ModRRpBjq for j P J after applying the functor M ÞÑ MbL ABj. Such a family is called conservative. One can drop the requirement on the maps of the covering A Ñ Bi to be finitely presented, obtaining another topology called the formal homotopy Zariski topology. Later on, we will relax the condition on coverings allowing the subset J Ă I to be countable. We will discuss this issue at the end of Section5. We will also make an extensive use of flat objects in C, in the following sense. Definition 2.6. Let pC,b, eCq be a closed, symmetric monoidal, quasi-abelian category. We call an object F of C flat if for any strictly exact sequence 0 Ñ E1 Ñ E Ñ E2 Ñ 0 6 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER the resulting sequence is strictly exact, i.e. if the endofunctor E ÞÑ EbF is an exact functor in the terminology of [36]. 0 Ñ E1bF Ñ EbF Ñ E2bF Ñ 0 We conclude this section by defining free resolutions in closed symmetric monoidal quasi- abelian categories and by showing some properties. Definition 2.7. Let pC,b, eCq be a closed symmetric monoidal quasi-abelian category and let A P CommpCq. An object E P ModpAq is called free if for some V P C. Definition 2.8. Let pC,b, eCq be a closed symmetric monoidal quasi-abelian category and let A P CommpCq. A free resolution of E P ModpAq is the data of a strict complex Ñ L2pEq Ñ L1pEq Ñ L0pEq Ñ 0 L‚pEq – E and a strict quasi-isomorphism where each LipEq is free in ModpAq and E is thought as a complex concentrated in degree 0. Lemma 2.9. Let pC,b, eCq be a closed symmetric monoidal quasi-abelian category with enough projectives. Let A P CommpCq and E P ModpAq. Then, E admits a free resolution. If in addition, both E and A are flat as objects in C then each term of the free resolution can be chosen to be a flat object in ModpAq. E – AbV Proof. Consider looooomooooon ApEq “ AbpAbbA L n bEq n times dn “ n´1ÿ ApEq is by definition a free A-module. Defining the differentials dn : L n where we think the first A factor as an A-module and the other factors as objects of C. In ApEq Ñ this way L n A pEq in the following way: Let mA : AbA Ñ A denotes the multiplication map of A L n´1 and ρE : AbE Ñ E the action of A on E, then i“0 p´1qiidAbbmAbbidAbidE ` p´1qnidAbbidAbρE, (2.1) where mA is at the ith place. Standard computations show that the complex L ‚ resolution of E. This complex is a resolution because L ‚ the maps A pEq ApEq is a free ApEq has a splitting over C given by ApEq ÐÝ L n´1 L n 1AbidL n´1 A pEq, (2.2) where 1A is the constant morphism to the identity of A. Therefore, we can deduce that the ApEq Ñ E is a strictly exact complex in Dď0pCq. By Proposition 1.5.1 cone of the map L ‚ of [36] a morphism in ModpAq is strict if and only if is strict as a morphism in C, hence the ApEq Ñ E is also a strictly exact complex in Dď0pAq. cone L ‚ ApEq. Since bA is right exact we only ApEq. Consider a strictly exact need to show only the left exactness of the functor p´qbAL n sequence of morphisms of ModpAq It remains to show the claim about the flatness of L n 7 0 Ñ F Ñ G Ñ H. Applying p´qbAL n (2.3) which can be rewritten as 0 Ñ FbpAbbA ApEq we obtain the sequence 0 Ñ FbAL n looooomooooon ApEq Ñ GbAL n looooomooooon bEq Ñ GbpAbbA ApEq Ñ HbAL n ApEq looooomooooon bEq Ñ HbpAbbA bEq. n times n times n times Therefore, the hypothesis that A and E are flat objects of C directly implies that the sequence l (2.3) is strictly exact. Remark 2.10. The splitting maps of equation (2.2) are not A-linear in general. Definition 2.11. Let pC,b, eCq be a closed symmetric monoidal quasi-abelian category. Let A P CommpCq and E P ModpAq. We define the Bar resolution of E to be the free resolution introduced in Lemma 2.9. Using the Bar resolution we can prove the following important lemma. Lemma 2.12. Let C be elementary quasi-abelian. Let tAiuiPI be a filtered inductive system in CommpCq such that all system morphisms are homotopy epimorphisms. Then, for any j P I the canonical maps Aj Ñ limÑ iPI Ai are homotopy epimorphisms. Ai is a homotopy epimorphism we need plimÑ iPI to check that Proof. Let’s fix a j P I. To show that Aj Ñ limÑ iPI Aiq – limÑ iPI AiqbL in Dď0pCq. Consider the Bar resolution L ‚ AiqbAj plimÑ iPI Aiq. More explicitly, for each n P N AjplimÑ loooooomoooooon iPI bplimÑ AiqbAj AjbpAjbbAj Aiq “ plimÑ iPI iPI is a representative of plimÑ iPI AiqbAj AjplimÑ iPI AjplimÑ iPI AjplimÑ L ‚ iPI AiqbL AjplimÑ iPI Aiq. Then, the complex plimÑ iPI L n Ai Aiq n times Aiqq 8 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER which simplifies to plimÑ iPI loooooomoooooon AiqbpAjbbAj n times bplimÑ iPI Aiqq – limÑ iPI loooooomoooooon pAib AjbbAj bAiq. n times is an exact functor when I is filtered we Now, because by Proposition 2.1.16 (c) of [36] limÑ iPI have limÑ iPI AiqbL plimÑ iPI – LlimÑ iPI AjplimÑ iPI AjplimÑ L ‚ iPI . Therefore, Aiq – plimÑ iPI Aiq – limÑ iPI l We also introduce the following complex needed for computing Cech cohomology in the AjpAiqq – limÑ L ‚ iPI Aiq – limÑ iPI AiqbAj pAibAj pAibL Aj Ai. theories we will develop. Definition 2.13. Given a collection of morphisms U “ tA Ñ BiuiPI in CommpCq and M P ModpAq we have the Cech-Amitsur complex C ‚ and its augmented version which we call the Tate complex T ‚ i i,j ApM, Uq ź ź pbABjq Ñ pMpbABi pMpbABiq Ñ ź ź pMpbABiq Ñ pMpbABi ź pMpbABi1 pbABjq Ñ pbA pbABidq ApM, Uq “ ApM, Uq i i,j 0 Ñ M Ñ where we use the degree convention ApM, Uq “ T d C d i1,...,id for d ě 1. 2.2. Bornological vector spaces. In this section we recall very briefly the theory of bornological vector spaces. We refer the reader to Section3.3 of [4] for more details. Definition 2.14. Let X be a set. A bornology on X is a collection B of subsets of X such that (1) B is a covering of X, i.e. @x P X,DB P B such that x P B; (2) B is stable under inclusions, i.e. A Ă B P B ñ A P B; (3) B is stable under finite unions, i.e. for each n P N and B1, ..., Bn P B, The pair pX,Bq is called a bornological set, and the elements of B are called bounded subsets of X (with respect to B, if it is needed to specify). A family of subsets A Ă B is called a basis for B if for any B P B there exist A1, . . . , An P A such that B Ă A1 YY An. A morphism of bornological sets ϕ : pX,BXq Ñ pY,BY q is defined to be a bounded map ϕ : X Ñ Y , i.e. a map of sets such that ϕpBq P BY for all B P BX. Definition 2.15. A bornological vector space over k is a k-vector space E along with a bornology on the underlying set of E for which the maps pλ, xq ÞÑ λx and px, yq ÞÑ x` y are bounded. Bi P B. nŤ i“1 Definition 2.16. A bornological vector space is said to be of convex type if it has a basis made of absolutely convex subsets. We will denote by Bornk the category whose objects are the bornological vector spaces of convex type and whose morphisms are bounded linear maps between them. 9 Remark 2.17. For every bornological vector space of convex type E there is an isomorphism E – limÑ BPDE EB where B varies over the family of bounded absolutely convex subsets of E and EB is the vector subspace of E spanned by elements of B equipped with the gauge semi-norm (also called Minkowski functional) defined by B and each EB is equipped with the bornology induced by this semi-norm. Reasoning as last remark, one can show that there is a functor diss : Bornk Ñ IndpSNrmkq which is fully faithful, which commutes with all projective limits and direct sums and whose essential image is the sub-category of essential monomorphic objects of IndpSNrmkq, i.e. by ind-objects isomorphic to systems which have monomorphisms as system maps. The functor diss does not commute with cokernels in general. Definition 2.18. A bornological vector space over k is said to be separated if its only bounded linear subspace is the trivial subspace t0u. Definition 2.19. A bornological vector space E over k is said to be complete if there is a small filtered category I, a functor I Ñ Bank and an isomorphism E – limÑ iPI Ei for a filtered colimit of Banach spaces over k for which the system morphisms are all injective and the colimit is calculated in the category Bornk. Let CBornk be the full subcategory of Bornk consisting of complete bornological spaces. Collecting together several facts from the theory of bornological vector spaces. See Section3.3 of [4] for a more detailed account. Lemma 2.20. CBornk has all limits and colimits and the inclusion functor CBornk Ñ Bornk commutes with projective limits. Both Bornk and CBornk are closed symmetric monoidal elementary quasi-abelian categories with enough projectives. product and the one of CBornk ispbπ,k (often written just aspbk) and it is called the completed The monoidal structure of Bornk is denoted by bπ,k and is called the projective tensor projective tensor product. We conclude by noticing that the fact that CBornk has enough projectives implies that we can always derive right exact functors from CBornk to a quasi- abelian category, as happens in the theory of abelian categories. 2.3. Tensor product of unbounded complexes. In this section we extend to the quasi- abelian settings some results of [9] on the derived functor of the tensor product for abelian categories. We see how the derived functor p´qbLp´q : Dď0pCq Dď0pCq Ñ Dď0pCq (dis- cussed extensively in [4]) extends to a functor p´qbLp´q : DpCq DpCq Ñ DpCq. Since 10 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER we are not looking for the utmost generality of the result we suppose in this section that C is an elementary quasi-abelian closed symmetric monoidal category, although it would be sufficient to suppose only that C has exact direct sums and enough projectives instead of supposing it to be elementary. Recall that the hypothesis of C being elementary implies that C has enough projectives, cf. Proposition 2.1.15 (c) of [36]. Lemma 2.21. Let X P DpCq, then the morphism τďnpXq Ñ ho limÑ nPN limÑ nPN τďnpXq is an isomorphism. Proof. We can apply the dual statement of Remark 2.3 of [9] because by Proposition 2.1.15 (a) of [36], LHpCq satisfies the axiom AB4. l Lemma 2.22. Every object in DpCq is strictly quasi-isomorphic to a complex of projective objects. Proof. Let X P DpCq. Consider the truncated complex τďnpXq (recall that we are using the left t-structure of DpCq). There is a canonical map τďnpXq Ñ X. Since C has enough projectives we can always find strict quasi-isomorphisms P ďn Ñ τďnpXq, where P ďn is a complex of projectives which is zero in degree ą n. The diagram P ďn P ďn`1 τďnpXq τďn`1pXq defines a morphism P ďn Ñ P ďn`1 in DpCq, because P ďn`1 Ñ τďn`1pX‚q is a strict quasi- isomorphism and this morphism can be realized as a morphism of complexes because P ďn and P ďn`1 are made of projectives. So, there is a sequence of morphisms limÑ nPN τďnpXq Ñ ho limÑ nPN τďnpXq Ñ ho limÑ nPN τďnpP ďnq τďnpXq – X and the last map is an isomorphism because is a homotopy colimit where limÑ nPN of strict quasi-isomorphisms. The first map is an isomorphism by Lemma 2.21. Therefore, X – ho limÑ l nPN τďnpP ďnq which by construction is a complex of projectives. Definition 2.23. Let T be a triangulated category with direct sums. We say that subcate- gory of T is localizing if it is closed under direct sums. We use the following notation: KpCq the homotopy category of C and KpPq the smallest localizing subcategory of KpCq containing the complexes made of projectives. Remark 2.24. Notice that if C is elementary, then both KpCq and DpCq have all the direct sums. Indeed, the claim on KpCq is trivial and for the one on DpCq, using Proposition 2.1.12 one deduce that C is derived equivalent to an elementary abelian category and therefore DpCq has all direct sums as consequence of Corollary 1.7 of [9]. Lemma 2.25. The composition functor KpPqãÑKpCq Ñ DpCq 11 is an equivalence. Proof. Since KpPq is thick subcategory of KpCq, we can use the dual statements of Lemmata 2.8-11 of [9] to deduce that KpPq is a localizing subcategory of KpCq. Using Lemma 2.22 we see that all objects of DpCq are isomorphic to objects of KpPq. l Theorem 2.26. The tensor product can be derived to a functor p´qbLp´q : DpCq DpCq Ñ DpCq. The restriction to Dď0pCqDď0pCq agrees with the derived tensor product based on projective resolutions. Proof. The tensor product p´qbp´q extends to a functor KpCq KpCq Ñ KpCq. By Lemma 2.25, under our hypothesis, DpCq is equivalent to a subcategory of KpCq. Therefore, we can define p´qbLp´q as the restriction to KpPq of the extension of p´qbp´q to KpCq. l 3. Some results in functional analysis This section is devoted to prove results in the theory of bornological vector spaces needed in the rest of the paper. 3.1. Proper bornological vector spaces. Definition 3.1. Let E be a bornological k-vector space and txnu a sequence of elements of E. We say that txnunPN converges (bornologically) to 0 if there exists a bounded subset B Ă E such that for every λ P k there exists an n “ npλq for which xm P λB,@m ą n. We say that txnunPN converges (bornologically) to a P E if txn ´ aunPN converges (bornolog- ically) to zero. The idea of this definition is due to Mackey and it is sometimes called Mackey convergence. Definition 3.2. Let E be a bornological vector space over k. ‚ a sequence txnunPN Ă E is called Cauchy-Mackey if the double sequence txn ´ xmun,mPN converges to zero; ‚ a subset U Ă E is called (bornologically) closed if every sequence of elements of U which is bornologically convergent in X converges bornologically to an element of U . Definition 3.3. A bornological vector space is called semi-complete if every Cauchy-Mackey sequence is convergent. 12 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER The notion of semi-completeness is not as useful as the notion of completeness in the theory of topological vector spaces. We remark that any complete bornological vector space is semi-complete, but the converse is false. Remark 3.4. The notion of bornological convergence on a bornological vector space of convex type E “ limÑ EB, where DE denotes the family of bounded disks of E, can be BPDE restated in the following way: txnunPN is convergent to zero in the sense of Mackey if and only if there exists a B P DE and N P N such that for all n ą N , xn P EB and xn Ñ 0 in EB (equipped with the semi-norm induced by B). Remark 3.5. The notion of bornologically closed subset induces a topology on E, but this topology is neither a linear topology nor group topology in general. However, an arbitrary intersection of bornological closed subsets of a bornological vector space is bornologically closed. From Remark 3.5 follows that the following definition is well posed. Definition 3.6. Let U Ă E be a subset of a bornological vector space. The (bornological) closure of U is the smallest bornologically closed subset of E in which U is contained. We denote the closure of U by U . Definition 3.7. Let E be a bornological vector space over k. We will say that a subset U Ă E is bornologically dense if the bornological closure of U is equal to E. Proposition 3.8. Let f : E Ñ F be a morphism in CBornk, then ‚ f is a monomorphism if and only if it is injective ‚ f is an epimorphism if and only if fpEq is bornologically dense in F ‚ f is a strict epimorphism if and only if it is surjective and F is endowed with the quotient bornology; ‚ f is a strict monomorphism if and only if it is injective, the bornology on E agrees with the induced bornology from F and fpEq is a bornologically closed subspace of F . l Proof. See Proposition 5.6 (a) of [30]. Definition 3.9. A bornological vector space is called proper if its bornology has a basis of bornologically closed subsets. Remark 3.10. (1) All bornological vector spaces considered in [2] and [4] are proper. (2) Let E be a separated bornological vector space (i.e. t0u is a closed subset of E), then the morphism E Ñ pE is injective (cf. Proposition 17 page 113 of [18]). Indeed, in this case, in E – limÑ BPDE EB then pE “ {limÑ BPDE EB – limÑ BPDE xEB. The main drawback, for our scope, of proper bornological vector spaces is that although they form a closed symmetric monoidal category, this category is not quasi-abelian. We will also need the following property of proper bornological vector spaces. Proposition 3.11. Any projective limit of proper objects in Bornk is proper. Proof. [18], Proposition 12, page 113. 13 l Lemma 3.12. Any subspace of proper bornological vector space, endowed with the induced bornology, is proper. Proof. Let E be a proper bornological vector space and let tBiuiPI be a basis for its bornology made of bornologically closed bounded subsets. Given a subspace F Ă E, the family tBi X FuiPI is a basis for the bornology that E induces on F . Consider a sequence of elements txnunPN Ă Bi X F converging bornologically to x P F then, since the inclusion map F Ñ E is bounded, it preserves bornological limits. Therefore, txnunPN converges to x also as a sequence of elements of E and since Bi is closed x P Bi. It follows that x P F X Bi and l therefore the claim. We have to warn the reader that the closure of a subset X Ă E of a bornological vector space is rarely equal to the limit points of convergent sequences of elements of X. So, we introduce the following notation. Definition 3.13. Let E be a bornological vector space and X Ă E, then Xp1q denotes the elements of E that are bornological limits for sequences of elements of X. Thus, we have the (in general strict) inclusion Xp1q Ă X. We now show that this inclusion is an equality for an important special case: linear subspaces of complete, proper bornological spaces. We need a preliminary lemma. Lemma 3.14. Let F Ă E be a linear subspace of a complete bornological vector space over k. Then, F is (bornologically) closed if and only if F is complete. Proof. This is a classical result in the theory of bornological vector spaces. For the convenience of the reader we reproduce here the proof that can be found in [19], chapter 3, Proposition 3.2.1. Moreover our proof deals also with the non-Archimedean case of the lemma not covered in [19]. Suppose that F Ă E is a complete subspace. Let txnunPN be a sequence of elements of F that converges bornologically to x P E. By Remark 3.4 there exists a bounded disk B Ă E such that xn Ñ x in EB. Since B X F is a bounded disk in F and F is complete can find a disk B1 Ă F such that B Ă B1 and FB1 is Banach. This shows that txnunPN is a Cauchy sequence in FB1, because the map FB Ñ FB1 is by construction bounded. Hence, the limit of txnunPN is an elements of F , which is hence closed. For the converse, suppose that F Ă E is a closed subspace. Let Dc E denotes the family of completant bounded disks of E, i.e. bounded disks B for which EB is Banach. It is enough to show that for any B P Dc E the bounded disk B X F of F is completant. So, let txnunPN be a Cauchy sequence in FBXF . This implies that txnunPN is a Cauchy sequence in EB and therefore it must converge to a point of EB. But FBXF is a closed subspace of EB, since it is the intersection of EB Ă E with a closed subspace of E, hence the limit of txnunPN is in l FBXF . This proves that FBXF is a k-Banach space and concludes the proof. Proposition 3.15. Let F Ă E be a subspace of a complete, proper bornological vector space over k, then F p1q “ F , where F p1q is as in Definition 3.13. 14 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Proof. Let E “ limÑ Ei be a presentation of E as a direct limit of Banach spaces over k as iPI in Definition 2.19. The bornology on F induced by the inclusion F Ă E can be described by where each F X Ei is endowed by the norm induced by Ei. By Lemma 3.12 F endowed with this bornology is a proper bornological vector space which is easily seen to be separated. By Remark 3.10 (2), the inductive system t{F X EiuiPI obtained by applying the comple- {F X Ei in which F embeds. Since pF is the completion of bornological vector space pF “ limÑ tion functor of Banach spaces, is monomorphic and hence it defines a complete and proper F and the inclusion F Ñ E must factor through a map iPI F – limÑ iPI pF X Eiq, φ : pF Ñ E. F X Ei Ñ {F X Ei Ñ Ei Moroever, φ is a strict monomorphism, because for each i P I the maps are strict, since all spaces are endowed with the restriction of the same norm, by construction. pF is a closed linear subspace of E because it is complete and hence we can apply Lemma 3.14. This implies that F Ă impφq – pF . Finally, let x P impφq then x “ φpyq for some y P {F X Ei and for some i P I. So, there exists a sequence of elements txnunPN Ă F X Ei which converges in norm to x, which shows that pF – impφq Ă F p1q Ă F . l We remark that in Proposition 3.15 the hypothesis that F Ă E is a (linear) subspace is crucial. It is possible to construct subsets S Ă E for which Sp1q Ă S strictly even when E is a proper object of CBornk (see [20] Theorem 1, SectionII.7). Remark 3.16. Proposition 3.15 in the case k “ C was proven in [25], cf. Proposition 4.14. The arguments of [25] are similar to ours, but the terminology and the context are very different. 3.2. Relations between bornological and topological vector spaces. We recall the definitions of two functors from [18], p´qt : Bornk Ñ Tck and p´qb : Tck Ñ Bornk, where Tck is the category of locally convex topological vector spaces over k. To a bornological vector space E we associate the topological vector space Et in the following way: we equip the underlying vector space of E with a topology for which a basis of 0-neighborhoods is given by bornivorous subsets, i.e. subsets that absorb all bounded subsets of E. If E is a locally convex space, Eb is defined to be the bornological vector space obtained by equipping the underlying vector space of E with the von Neumann (also called canonical) bornology, whose bounded subsets are the subsets of E absorbed by all 0-neighborhoods. In chapter 1 of [18] one can find details of these constructions and their properties of which the main one is that (3.1) is an adjoint pair of functors. p´qt : Bornk Ô Tck : p´qb Lemma 3.17. It is shown in [18] (cf. Chapter 2, §2 n 6) that there are full sub-categories of Bornk and Tck for which these functors give an equivalence of categories. 15 Definition 3.18. We call the objects of the categories mentioned in Lemma 3.17 normal bornological vector spaces or normal locally convex spaces (depending on the ambient cat- egory we are thinking them in). We call the corresponding full sub-categories of normal objects NBornk Ă Bornk and NTck Ă Tck. For two elements E, F P NTck (or E, F P NBornk) the notion of boundedness and conti- nuity of a linear map f : E Ñ F is equivalent. Remark 3.19. Any normal bornological vector space is proper (cf. Definition 3.9). One can show this fact by first noticing that in any topological vector space the closure of bounded subset is bounded. Hence, the von Neumann bornology of a locally convex topological vector space always has a basis made of closed (for the topology) subsets. Moreover, every subset which is closed for the topology is also bornologically closed for the von Neumann bornology. For more details on this topic see [18], Chapter 2, §3, n 6 and 7. Observation 3.20. The functor p´qb : Tck Ñ Bornk commutes with projective limits and p´qt : Bornk Ñ Tck commutes with inductive limits. Observation 3.21. Let E be a bornological vector space over k. It follows that a subset U Ă E is bornologically dense then U Ă Et is topologically dense. Indeed, all bornologically closed subsets for the von Neumann bornology of Et must be bornologically closed for the bornology of E and, as said in Remark 3.19, all topologically closed subsets of Et are bornologically closed for the von Neumann bornology. Therefore, the closure of U in Et contains the closure of U in E. Example 3.22. ‚ All semi-normed spaces over k are in an obvious way normal objects in Bornk and Tck. ‚ All metrizable locally convex vector spaces and metrizable bornological vector spaces of convex type are normal. See the beginning of page 109 of [18] or Proposition 3 at page 50 of [19]. ‚ The underlying object in CBornk of a k-dagger affinoid algebra is a normal bornolog- ical vector space (see [2], Chapter 3). We introduce some classes of normal spaces that will be used throughout the paper. Definition 3.23. A Fr´echet space over k is a complete, metric, locally convex topological vector space over k. A bornological Fr´echet space over k is a bornological vector space E P CBornk such that E – F b for a Fr´echet space F P Tck. Definition 3.24. A LB space (respectively LF space) over k is a locally convex topological vector space over E such that (3.2) E – limÑ nPN En where En is a Banach space (respectively a Fr´echet space) and the system maps are injective. 16 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER All LB spaces are normal locally convex spaces but the functor p´qb may not commute with the colimit (3.2). Definition 3.25. An LB space or an LF space over k is said to be regular if Eb – limÑ nPN Eb n. Definition 3.26. We say that E P Bornk is a bornological LB space if E – F b where F P Tck is a regular LB space. The following criterion will be very important in following sections. Proposition 3.27. Consider a projective system tEiuiPI of objects of NBornk. The following conditions are equivalent Eiqt. (1) limÐ iPI (2) limÐ iPI Proof. If limÐ iPI pEt iq is normal in Tck; iq – plimÐ pEt iPI iq is normal, then by Observation 3.20 pEt Eiqt. i qqt – plimÐ pEtb iPI pEt iq – plimÐ On the other hand, the same calculations also shows that if limÐ iPI iPI pEt Eiqt – limÐ iq. iPI iq – pplimÐ pEt iPI iqbqt – plimÐ iPI pEtb i qqt – plimÐ iPI Et iqbqt – plimÐ iPI limÐ iPI Eiqt pplimÐ Et iPI iq is normal. pEt Therefore, limÐ iPI l Finally, in next section we will use the following class of bornological vector spaces. Definition 3.28. A bornological vector space is said to be regular if it has a basis for its bornology made of subsets which are closed subsets with respect to the topology of Et. Remark 3.29. The condition of Definition 3.28 is strictly stronger than the condition of properness of Definition 3.9, where the requirement is that E has a basis of bornologically closed subsets. 3.3. Nuclear bornological spaces and flatness. The functor CBornk Ñ CBornk, given by E ÞÑ Epbπ,kF where F is a complete bornological vector space, is right exact (even strongly right exact) but not left exact, in general. In order to have a sufficient condition to ensure the exactness of such functors, we use the notion of nuclear operator. This notions was introduced by Grothendieck in the context of locally convex spaces over Archimedean base fields. We will also need to introduce the injective tensor product over maximally complete fields (non-Archimedean or not). Definition 3.30. Conside a morphism f : E Ñ F in Bank. The morphism f is called nuclear if there exist two sequences tαnu Ă HompE, kq and tfnu Ă F where ‚ 8ř n“0 ‚ }αn}}fn} Ñ 0, for n Ñ 8 if k is non-Archimedean; }αn}}fn} ă 8 if k is Archimedean; 8ÿ such that fpxq “ αnpxqfn n“0 17 for all x P E. Remark 3.31. A morphism f is nuclear in the sense of Definition 3.30 if and only if f is in the image of the canonical morphism Hompk, E_pbkFq Ñ Hompk, HompE, Fqq – HompE, Fq, where E_ “ HompE, kq. This definition makes sense in any closed, symmetric monoidal category. In this abstract context, the pre-composition of a nuclear morphism with any morphism is a nuclear morphism. Similarly, the post-composition of a nuclear morphism with any morphism is nuclear (see Proposition 2.1 of [22] for a proof of these facts). Definition 3.32. Let V be a locally convex k-vector space. A subset B Ă V is called compactoid if for any neighborhood of the origin U there is a finite subset F Ă V such that B Ă U ` ΓpFq, where Γ denotes the absolute convex hull of F . Recall that a subset B Ă V of a locally convex k-vector space is called pre-compact if for any neighborhood of the origin U for a finite set F Ă V . Proposition 3.33. Let V be a locally convex k-vector space. If k is locally compact then a subset B Ă V is compactoid if and only if it is pre-compact. B Ă U ` F The map kn Ñ E given by pλ1, . . . , λnq ÞÑ nř Proof. Let B Ă V be a compactoid subset and U Ă V be an absolutely convex neigh- borhood of zero. Then, there is a finite set F “ tx1, . . . , xnu Ă V such that B Ă U ` ΓpFq. λixi is continuous and it maps the compact set pkqn onto ΓpFq. Therefore, ΓpFq is compact. This means that there is a finite set F 1 Ă V such that ΓpFq Ă U ` F 1, so i“1 B Ă U ` U ` F 1. If k is non-Archimedean we can conclude that B Ă U ` F 1 proving that B is pre-compact. If k is Archimedean, we can notice that the same reasoning can be done with U 2 in place of U , so that ` U 2 B Ă U 2 ` F 1 Ă U ` F 1 proving that B is pre-compact also in this case. 18 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER l The converse statement, that every pre-compact subset is compactoid, is obvious. Thanks to Proposition 3.33, a compactoid subset in the case when k is Archimedean is simply pre-compact subset. The notion of compactoid is necessary when the base field is not locally compact, because in this case the definition of pre-compact subset is not useful. Definition 3.34. The family of compactoid subsets of a locally convex topological vector space V over k forms a bornology, which is denoted by CptpV q. Compactoid subsets are always bounded subsets in the sense of von Neumann, but the converse is often false. For example, one can show that if on a Banach space every bounded subset is compactoid, then the Banach space must be finite dimensional. Definition 3.35. Consider a morphism f : E Ñ F in Bank. The morphism f is called compactoid if the image of the unit ball of E is a compactoid (in the sense of Definition 3.32) subset of F . Remark 3.36. A morphism f : E Ñ F in Bank is compactoid if and only it is a bounded morphism when considered as a map of bornological spaces f : Eb Ñ CptpFq. Notice that what we called a compactoid map in literature is just called a compact map, in the case in which k is Archimedean. If k is Archimedean a nuclear morphism is compact but there exist compact morphisms which are not nuclear. However, the non-Archimedean case is simpler. Proposition 3.37. Let k be non-Archimedean and f : E Ñ F a bounded morphism in Bank. Then f is nuclear if and only if it is compactoid. Proof. In classical terminology what we called nuclear maps over a non-Archimedean field are called completely continuous maps. With this terminology, the proposition is proved as l Proposition 2 in page 92 of [15]. Definition 3.38. An object of CBornk is said to be nuclear if there exists an isomorphism E – limÑ iPI Ei as in Definition 2.19 where for any i ă j in I the map Ei Ñ Ej is a nuclear monomorphism of k-Banach spaces. Therefore, by definition, nuclear bornological spaces are always complete bornological spaces. We now limit the discussion to the case where k is maximally complete. This restriction will be removed later. Definition 3.39. Let k be a maximally complete valuation field. Let E, F be objects of Bank and let with E_ “ HomBankpE, kq and F _ “ HomBankpF, kq. Then, the (completed) injective tensor product is defined to be the completion of the algebraic tensor product Ebk F equipped with the semi-norm ÿ ÿ where pE_q and pF _q denote the unit balls. We denote the injective tensor product by } ei b fi},k “ E b,k F ad the completed one by Epb,kF . sup αPpE_q,βPpF _q αpeiqβpfiq The following is Theorem 1 of page 212 of [18]. Lemma 3.40. Let k be a non-Archimedean, maximally complete valuation field. Then, for any E, F P Bank, the natural morphism in Bank E bπ,k F Ñ E b,k F is an isomorphism. 19 Definition 3.41. Let k be a maximally complete valuation field and let E, F be two objects of CBornk. We define the injective tensor product of E and F as the colimit in CBornk of an the monomorphic inductive system: E b,k F “ EBE b,k FBF , (3.3) where DE and DF are the family of bounded disks of E and F respectively. The completed injective tensor product of E and F is defined limÑ BEPDE ,BFPDF Epb,kF “ {E b,k F . Remark 3.42. The inductive system (3.3) is monomorphic because p´qb,kp´q is a strongly left exact functor, cf. Corollary 2 page 210 of [18]. Definition 3.41 differs from the one given in [18] chapter 4, §2, but is equivalent to it for thanks to the following lemma. Lemma 3.43. Let k be a maximally complete valuation field. Let tEiuiPI and tFjujPJ be two filtered inductive systems of complete bornological vector spaces such that all Ei and Fi are separated and regular (in the sense of Definition 3.28) and all system morphisms are injective. Then, in CBornk we have that E b,k F – limÑ pi,jqPIJ pEi b,k Fiq – plimÑ iPI Eiq b,k plimÑ jPJ Fjq, where on the right hand side the injective tensor product is the one considered in [18]. Proof. See [18] Proposition 4, page 210. l Remark 3.44. The injective tensor product does not commute with filtered inductive limits in general. Since all Banach spaces are obviously regular and complete bornological vector spaces are essentially monomorphic objects in IndpBankq, Lemma 3.43 readily implies that our definition of injective tensor product agrees with the definition of [18] page 208, for proper complete bornological vector spaces. Proposition 3.45. Let k be a maximally complete valuation field. Let E and F be objects of CBornk. In the Archimedean case we assume that F is nuclear. Then, the natural morphism Epbπ,kF ÝÑ Epb,kF in CBornk is an isomorphism. 20 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Proof. We start with the case in which k is non-Archimedean. Write E “ limÑ iPI F “ limÑ jPJ Ei and Fj as in Definition 2.19. The isomorphisms Ei bπ,k Fj Ñ Ei b,k Fj from Lemma 3.40 gives an isomorphism of systems limÑ pi,jqPIJ pEi bπ,k Fjq Ñ limÑ pi,jqPIJ pEi b,k Fjq. The projective tensor product commutes with all colimits and by Lemma 3.43 also the injec- tive tensor product commutes with colimit we are calculating. Hence, we get an isomorphism plimÑ iPI Eiq bπ,k plimÑ jPJ Fjq Ñ plimÑ iPI Eiq b,k plimÑ jPJ Fjq and the functoriality of the completion yields the claimed isomorphism. We now consider the case in which k is Archimedean. First recall the following fact: let f : E1 Ñ E2 be a nuclear morphism of Banach spaces and V another Banach space, then the linear map of vector spaces f bk idV : E1 b,k V Ñ E2 bπ,k V is bounded. This is precisely the content of Proposition 2.4, page I-15 of [13], where the complex case is discussed, but the same arguments work also over R. Then, consider F “ limÑ Ei presentations of F and E as in Definition 2.19 and as in Definition jPJ 3.38 respectively. For any pi1, j1q ă pi2, j2q the morphisms Fj and E “ limÑ iPI αi1,j1 : Ei1 b,k Fj1 Ñ Ei2 b,k Fj1 Ñ Ei2 bπ,k Fj2 are bounded morphisms of Banach spaces because of the mentioned Proposition 2.4 of [13]. By Lemma 3.43 we have that the map E b,k F – plimÑ iPI limÑ pi,jqPIJ Eiq b,k plimÑ jPJ pEi bπ,k Fjq – plimÑ iPI Fjq – limÑ pi,jqPIJ Eiq bπ,k plimÑ jPJ pEi b,k Fjq αÑ Fjq – E bπ,k F obtained by the morphisms αi1,j1 is a bounded map, and is easy to check that the composition of α with the canonical map E bπ,k F Ñ E b,k F is the identity on both sides. We conlcude the proof using again the functorialiy of the l completion of bornological vector spaces. Lemma 3.46. Let k be a maximally complete field and F P CBornk. The functor p´qb,k F is strongly left exact. l Proof. See [18] Corollary 2, page 210. Lemma 3.47. The completion functor yp´q : Nrmk Ñ Bank is exact. Proof. Proposition 4.1.13 of [31] shows that the completion functor is exact for locally 21 convex spaces. Since the functor as locally convex spaces, we can use that proposition to deduce our lemma. Notice that [31] deals only with Archimedean base fields but the same reasoning works also for non- Archimedean base fields since only the uniform structures of the spaces are considered. l yp´q : Nrmk Ñ Bank precisely agrees with the completion yp´q : Nrmk Ñ Bank is not strongly left exact and does not preserve monomor- Remark 3.48. phisms. Lemma 3.49. If k is non-Archimedean, then any object in the category of non-Archimedean Banach spaces over k is flat. Proof. If k is maximally complete, then the claim follows from the combination of Lemma 3.46, Lemma 3.40 and Lemma 3.47 . If k is not maximally complete we can choose a maximal completion K{k. For every E P Bank the canonical map ιE : E Ñ E bk K is a strict monomorphism (even an isometry onto its image, cf. Lemma 3.1 of [29] applied with A “ k, B “ K and C “ E). Consider a strict exact sequence (3.4) 0 Ñ ker f Ñ E fÑ F in Bank. We can always suppose that ker f is equipped with the restriction of the norm of E, in which case ker f Ñ E is an isometric embedding. Applying again Lemma 3.1 of [29], we obtain a diagram 0 0 ker f E f F ker f bk K E bk K F bk K where all vertical maps are isometric embeddings. The bottom row is exact algebraically. It is also strictly exact, because we can apply Lemma 3.1 of [29] to the isometric embedding ker f Ñ E (precisely using A “ ker f, B “ E and C “ K this time) to deduce that ker f bk K Ñ E bk K is an isometric embedding. Then, consider and a G P Bank. Tensoring (3.4) with G we obtain a diagram 0 0 ker f bk G E bk G f F bk G pker f bk Gq bk K pE bk Gq bk K pF bk Gq bk K 22 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER where the top row is algebraically exact. We need only to check that ker f bk G Ñ E bk G is a strict monomorphism. Using the isomorphisms pE bk Gq bk K – pE bk Kq bK pG bk Kq, pF bk Gq bk K – pF bk Kq bK pG bk Kq pker f bk Gq bk K – pker f bk Kq bK pG bk Kq last diagram becomes 0 0 ker f bk G E bk G f F bk G pker f bk Kq bK pG bk Kq pE bk Kq bK pG bk Kq pF bk Kq bK pG bk Kq where the bottom row is strictly exact because K is maximally complete. Hence, in the diagram ker f bk G E bk G pker f bk Kq bK pG bk Kq pE bk Kq bK pG bk Kq all maps are known to be strict monomorphism but the top horizontal, which is therefore a strict monomorphism too. This shows that the functor p´q bπ,k p´q is strongly exact, and l applying Lemma 3.47 we deduce that p´qpbπ,kp´q is exact. then F is flat with respect the category pCBornk,pbk, kq. If k is Archimedean then the same Theorem 3.50. Let F be complete bornological vector spaces over k. If k is non-Archimedean We now look at flatness in the closed symmetric monoidal category CBornk. conclusion holds provided that F is nuclear. Proof. When k is Archimedean, it is of course maximally complete. So, the theorem follows immediately from combining Lemma 3.46 and Proposition 3.45. Suppose now that k is a non-Archimedean complete valuation field and fix a representation Fj as a monomorphism filtered inductive system of Banach spaces. Unravelling the F “ limÑ definitions, the functor p´qpbπ,kF can be written as jPJ p´qpbπ,kF “ {p´q bπ,k F “ { {p´q bπ,k Fj p´q bπ,k Fj “ limÑ jPJ limÑ jPJ functor p´qpbπ,kF can be written as a composition of exact functors. where last colimit is calculated in CBornk, meaning that it is the separated colimit of bornological vector spaces. Since CBornk is an elementary quasi-abelian category, the fil- tered colimits are exact. Therefore, applying Lemma 3.49 and Lemma 3.47 we see that the l We conclude this section with some results about nuclearity needed in subsequent sections. E – limÑ iPI pE X Fiq ś ś and E X Fi are k-Banach spaces. It is enough to check that the morphisms E X Fi Ñ E X Fj are nuclear for each i ă j. Thanks to Proposition 3.37 this is equivalent to check that the maps E X Fi Ñ E X Fj are compactoid. This follows from [27], Theorem 8.1.3 (ii) and (iii). Now, we check that countable products of nuclear bornological spaces are nuclear. Thus, suppose that tEnunPN is a family of nuclear bornological spaces and put E “ En. Fix for any n a completant bounded disk Bn Ă En such that there exists a completant bounded disk An Ă En for which Bn Ă An and the inclusion EBn Ñ EAn is a nuclear map (or equivalently compactoid, cf. Proposition 3.37). Let B “ It is clear that by ś varying Bn over a final family of bounded completant disks of En, for each n, the family of disks B Ă E and A Ă E obtained in this way form a final family of bounded disks for the bornology of E. By construction the map EB Ñ EA is bounded, where EA is the subspace EAn consisting of bounded sequences equipped with the supremum norm, and EB is of EBn. We need to check that EB Ñ EA sends the unit ball of the analogous subspace of EB to a compactoid subset of EA. Fix a α P k such that α ă 1. By rescaling the norms of EBn and EAn we can assume that Bn is contained in the ball of radius αn of EAn. Since Bn is compactoid in EAn, by Theorem 3.8.25 of [27] there exists a zero sequence tx i uiPN such pnq that Bn Ă Γptx i uiPNq, where Γ denotes the absolutely convex hull. Then, we can consider pnq the sequence Bn and A “ ś ś nPN An. nPN nPN nPN nPN Proposition 3.51. Any countable projective limit in CBornk of nuclear objects is nuclear (and therefore by Theorem 3.50 flat). 23 Proof. The proposition is proved in [21] Theorem 4.1.1, page 200, for the case in which k is an Archimedean base field. So, we check the claim only for non-Archimedean base fields. First we show that a closed subspace of a nuclear bornological space is nuclear. Let Fi as in Definition 3.38. As E Ă F be a closed subspace of a nuclear space and let F – limÑ iPI bornological vector spaces, there is an isomorphism xpnq “ tpx pnq i Since by hypothesis, for each i we have that quiPN Ă EA. then the sequence xpnq is a zero sequence of EA whose absolutely convex hull is B. So l applying Theorem 3.8.24 of [27] we can deduce that B is compactoid in EA. nÑ8x lim pnq i EAn “ 0 Proposition 3.52. Any small inductive limit in CBornk of nuclear objects is nuclear. Proof. It is an easy consequence of Definition 3.38 that any monomorphic filtered in- ductive limit of bornological nuclear spaces is a nuclear bornological space. Then, by the description of coproducts in CBornk of Lemma 2.7 of [4] (which agree with coproducts in IndpBankq) and applying Proposition 3.51 to finite coproducts, we obtain that coproducts of 24 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER bornological nuclear spaces are nuclear bornological spaces. Therefore, it remains only to check that quotients of nuclear spaces are nuclear. The Archimedean case of the proposition is proved in Theorem 4.1.1, page 200 of [21]. So, let k be non-Archimedean, F a nuclear bornological space over k and E Ă F a closed subspace. Fixing an isomorphism F – limÑ iPI – limÑ iPI Fi as in Definition 3.38, we can write FiXE Ñ Thanks to Proposition 3.37 it is enough to prove that the system morphisms φi,j : FjXE for any i ď j are compactoid maps. By hypothesis Fi Ñ Fj is a compatoid map and since the image of compactoid subsets by bounded maps are compactoid subsets, then also Fi Ñ Fj FjXE is a compactoid map. By Theorem 8.1.3 (xi) of [27] this is equivalent to say that l φi,j is a compactoid map. We recall that a dagger affinoid algebra is a bornological algebra which is isomorphic to a kpρq of over-convergent analytic functions on the polycylinder of quotient of the algebra W n polyradius ρ. The isomorphism F E Fj Fi Fi X E . Fi kpρq – limÑ W n rąρ k prq T n k prq are the k-Banach algebras of strictly convergent power-series on the polycylinder where T n kpρq with a bornology and dagger affinoid algebras are endowed of polyradius ρ, endow W n with the quotient bornology. We refer to Section4 of [4] of chapter 3 of [2] for more details. Proposition 3.53. The underlying bornological vector spaces of dagger affinoid algebras are nuclear. If k is Archimedean one can write W n Proof. Since by Proposition 3.52 quotients of bornological nuclear spaces are nuclear, kpρq is nuclear. We first look at the case in which k is non- it is enough to check that W n k pρ1q, for ρ1 ă ρ k pρq Ñ T n Archimedean. It is easy to check the canonical restriction maps T n are compactoid maps, which proves the claim. For details the reader can see Theorem 11.4.2 kpρq. and Remark 11.4.3 of [27] where there A:pρq is what in our notation is W n kpρq as the direct limit of the filtered system of Fr´echet spaces of holomorphic functions on open polydisks of polyradius bigger than ρ, see Section3.3 of [2] for a detailed proof of this fact. Thanks to a celebrated theorem of Montel these Fr´echet spaces are nuclear Fr´echet spaces (see for example Proposition 4 on page 26 of [13] for a proof of this fact) and by Lemma 3.60 this is equivalent to say that these spaces are nuclear as bornological spaces. So, since by Lemma 3.52 direct limits of nuclear bornological spaces are nuclear we deduce that W n l 3.4. A relative flatness lemma. In last section we proved that if F is a nuclear bornolog- ical vector space then the functor p´qpbkF is exact. Now we want to find conditions on p´qpbkF to preserve a different kind of projective limit: The cofiltered ones. The notion of kpρq is nuclear. projective tensor product was introduced by Grothendieck, in the category of locally convex spaces, in order to find a topological tensor product which commutes with projective limits. This is the origin of the name projective tensor product. But the bornological projective 25 tensor product does not commute with projective limits, in general. So, we need to find some sufficient conditions to ensure this commutation for the projective limits we will study in Section5. This study involves comparing the bornological and topological projective tensor products. Thus, we start by recalling the notion of topological projective tensor product. Definition 3.54. Given E, F P Tck one defines E bπ,k F P Tck as the algebraic tensor product equipped with the locally convex topology whose base of neighborhoods of zero is given by the family of absolutely convex hulls of subsets of E bk F of the form U b V “ tx b y x P U, y P V u space obtained in this way is denoted Epbπ,kF . where U and V vary over the family of neighborhoods of E and F respectively. The complete projective tensor product is defined as the separated completion of the space E bπ,k F . The Definition 3.55. E P Tck is said to be nuclear if E – limÐ Ei for a cofiltered epimorphic pEj Ñ pEi of Banach spaces induced on the separated completions are nuclear maps. iPI projective system of semi-normed spaces tEiuiPI such that, for each i ă j in I, the maps Definition 3.56. Given E, F P Tck one defines the injective tensor product E b,k F P Tck as the algebraic tensor product equipped with the semi-norms t} }pibqjui,jPIJ , where for each i, j P I J the semi-norm } }pibqj is defined as in Definition 3.39, where tpiuiPI is a family of semi-norms that defines the topology of E and tqjujPJ is a family of semi-norms that defines the topology of F . The separated completion of E b,k F is by definition the complete injective tensor product and is denoted Epb,kF . Lemma 3.57. If E P Tck is nuclear then the functors p´qpb,kE and p´qpbπ,kE are naturally isomorphic. Proof. For the Archimedean case one can refer to Theorem 1 of page 25 of [13]. For the non-Archimedean case the result is a direct consequence of Theorem 8.5.1 and Theorem l 10.2.7 of [27]. Definition 3.58. Given E P NBornk, we say that E is binuclear is E is a nuclear bornological space and Et is a nuclear topological vector space. Lemma 3.59. Let E P Tck. Suppose there exists a countable monomorphic compactoid inductive system of Banach spaces tEnu be such that E – limÑ En (in particular E is a nPN normal space, cf. Definition 3.18). Then, E is nuclear as locally convex space and Eb is nuclear as bornological vector space. Proof. If k is Archimedean, then this lemma is proved in [21] Theorem 7, page 160, since compactoid inductive limits are complete. Let k be non-Archimedean. Then, by Theorem En 11.3.5 (v) of [27] E is a regular LB-space (in the sense of Definition 3.25), hence Eb – limÑ nPN as bornological vector space, so Eb is nuclear. And by Theorem 11.3.5 (ix) of [27] E is l nuclear as locally convex space. For Fr´echet spaces the situation is a bit more complicated. 26 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Lemma 3.60. If k is Archimedean, then a Fr´echet space E P Tck is nuclear as locally convex vector space if and only if Eb is nuclear as bornological vector space. For non-Archimedean base fields, if a Fr´echet space E P Tck is nuclear as locally convex vector space then Eb is nuclear as bornological space. Proof. If k is Archimedean then this lemma is proved in [21] Theorem 7 (i) and (ii), page 160. If k is non-Archimedean then the nuclear Fr´echet space E is Montel by Corollary 8.5.3 of [27]. Therefore, by Theorem 8.4.5 (δ) of [27], each bounded subset of E is compactoid, i.e. there is an isomorphism of bornological vector spaces CptpEq – Eb. Let’s write as in Remark 2.17, where DE is the family of bounded disks of E, and Eb – limÑ BPDE EB CptpEq – limÑ BPCptE EB limÑ BPCptE EB – limÑ BPDE EB EB Ñ EB1 where CptE is the family of compactoid disks of E. The family CptE is characterized as the family of bounded disks B of E for which there exists another bounded disk B1 such that B Ă B1 and the canonical map EB Ñ EB1 is compactoid. This characterization, in combination with the isomorphism imply that for every bounded disk B of E there exists another bounded disk B1 such that B Ă B1 and the canonical map is compactoid. Since E is complete we can always suppose that B and B1 are Banach disks (i.e. that EB and EB1 are Banach). This proves that Eb satisfies the conditions of Definition l 3.38. Remark 3.61. If k is non-Archimedean there exists a Fr´echet space E for which Eb is nuclear whereas E is not nuclear. This is due to the fact that, for non-Archimedean base fields, a morphism of Banach spaces is nuclear if and only if it is compactoid. But the Archimedean analogous of the notion of compactoid map is what is called a compact map (or operator), in classical functional analysis over R and C . Therefore, a nuclear Fr´echet space over a non-Archimedean base field, following this terminology, is analogous to what in Archimedean functional analysis is a called Schwartz-Fr´echet space, (see [38] and the first chapter of [21] for a detailed account of the properties of Schwartz-Fr´echet spaces). In the Archimedean case it is known that a Fr´echet space is bornologically Schwartz if and only if it is Montel (cf. Theorem 8 of page 20 of [21]). Also in the non-Archimedean case a Fr´echet space is bornologically nuclear if and only if it is Montel (cf. Corollary 8.5.3 of [27]). One can prove (in both settings) that there exist Fr´echet-Montel spaces which are not Schwartz. For explicit examples of such spaces, see Counterexamples 9.8.2 (vi) of [27], for k non-Archimedean, and §4 of [38] for k Archimedean. We will be interested in finding conditions for which the functors p´qt and p´qb intertwine the complete bornological and the complete topological projective tensor products. In general p´qt and p´qb preserve neither the complete nor the incomplete projective tensor products. Proposition 3.62. Let E, F P Bornk be bornological Fr´echet spaces one of which is binuclear, then 27 If E, F P Tck are Fr´echet spaces one of which is nuclear, then pEpbπ,kFqt – Etpbπ,kF t. pEpbπ,kFqb – Ebpbπ,kF b. Proof. In page 215 of [18] it is proved that for metrizable topological or bornological spaces and pE b,k Fqt – Et b,k F t pE b,k Fqb – Eb b,k F b. Since for metric topological or bornological spaces the bornological and the topological notion of convergence agree (see the last lines of page 108 of [18] or Proposition 1.17 of [3] for a detailed proof of this fact) it follows that for metric spaces the notion of bornological and topological completeness agree (see also Corollary 1.18 of [3]). Therefore, we deduce that and pEpb,kFqt – Etpb,kF t pEpb,kFqb – Ebpb,kF b. Finally, since one of E or F is binuclear we deduce that the complete injective and projective tensor products coincide by Lemma 3.57 and Proposition 3.45 (in both categories Tck and l Bornk), obtaining the required isomorphisms. We also underline that in Theorem 2.3 of [25] the Archimedean version of Proposition 3.62 is discussed. Lemma 3.63. Let tEiuiPI be a cofiltered projective system in Tck whose projective limit we call E and F P Tck. Then Epbπ,kF “ plimÐ Eiqpbπ,kF – limÐ pbπ,kFq pEi iPI iPI Proof. Cf. Proposition 9 at page 192 of [18]. l Corollary 3.64. Let tEiuiPN be a cofiltered projective system of Banach vector spaces such that limÐ Ei is a binuclear Fr´echet bornological vector space and F a Fr´echet bornological vector iPN space, over k. Then, the canonical map Epbπ,kF “ plimÐ Eiqpbπ,kF Ñ limÐ pbπ,kFq pEi iPN iPN is an isomorphism of bornological vector spaces. 28 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Proof. The claim follows form the chian of isomorphisms Epbπ,kF 3.22– pEpbπ,kFqtb 3.62– pEtpbπ,kF tqb 3.63– plimÐ pbπ,kFq 3.20– limÐ iPN pbπ,kF tqb – limÐ pEt pEi iPN iPN i pbπ,kF tqqb – pEt i where last isomorphism is immediate from the definition of projective tensor product (com- l pare Definition 3.54 with Definition 3.57 of [4]). Corollary 3.65. Let tEiuiPN be a cofiltered projective system of binuclear Fr´echet bornological vector spaces and F bornological Fr´echet vector space, over k. Then, the canonical map Epbπ,kF “ plimÐ Eiqpbπ,kF Ñ limÐ pbπ,kFq pEi iPN iPN is an isomorphism of bornological vector spaces. Proof. Notice that Lemma 3.51, together with the well-known fact that a projective limit of nuclear spaces in Tck is a nuclear space and Proposition 3.27 imply that E is binuclear. So, also for this corollary we have the chian of isomorphisms Epbπ,kF 3.22– pEpbπ,kFqtb 3.62– pEtpbπ,kF tqb 3.63– plimÐ pbπ,kFq. 3.20– limÐ iPN pbπ,kF tqb 3.62– limÐ pEt pEi iPN iPN i pEt i pbπ,kF tqqb – l 3.5. Strict exact sequences in bornological and topological settings. This section contains some results about how the notions of strictly short exact sequence in Bornk and Tck are related, which are needed later. Lemma 3.66. Let f : E Ñ F be a surjective continuous map between Fr´echet spaces. For any compactoid subset B Ă F , there exists a compactoid subset B1 Ă E such that In particular, the functor Cpt : Tck Ñ CBornk preserves strict short exact fpB1q “ B. sequences of Fr´echet spaces. Proof. When k “ R, C, see [24] Theorem 1.62, and for the non-Archimedean case see l Theorem 3.8.33 [27]. Lemma 3.67. Let F be a nuclear Fr´echet space, then the von Neumann and the compactoid bornology conicides, i.e. the identity map gives an isomorphism F b – CptpFq in CBornk. Proof. In both cases one can show that nuclear Fr´echet spaces are Montel spaces, and for Montel spaces the von Neumann bornology and the compactoid bornology agree (by definition). Possibile references for these results are sections 8.4 and 8.5 of [27] for non- Archimedean base fields and Theorem 8 of page 20 of [21] for Archimedean base fields. l Lemma 3.68. Let E be a nuclear Fr´echet space and txnu a sequence of elements of E. Then, the following are equivalent 29 (1) txnu converges topologically to 0; (2) txnu converges bornologically to 0 for the bornology of Eb; (3) txnu converges bornologically to 0 for the bornology of CptpEq. Proof. This result is proven for Archimedean base fields in [25] Corollary 3.8. For the general case: For the equivalence of (1) and (2) we can refer to [3], Proposition 1.17. Finally, l conditions (2) and (3) are equivalent by Lemma 3.67. Remark 3.69. We conjecture that Lemma 3.68 holds for all Fr´echet spaces, without the nuclearity hypothesis. Indeed the proof given in [25] for Archimedean base fields, does not use the nuclearity hypothesis but it easily adapts to the non-Archimedean case only when the base field is locally compact. Notice that the only missing implication for a general non-Archimedean base field is (3) implies (2) or (3) implies (1). Corollary 3.70. Let f : E Ñ F be a morphism of nuclear Fr´echet spaces. Then, the following are equivalent: (1) f is an epimorphism in the category of complete locally convex vector spaces; (2) f : Eb Ñ F b is an epimorphism in CBornk (3) f : CptpEq Ñ CptpFq is an epimorphism in CBornk; (4) f : Eb Ñ F b has bornologically dense image. Proof. The characterization of epimorphisms given in Proposition 3.8 in combination with Lemma 3.67 yields that conditions (2), (3), and (4) are equivalent. Then, Lemma 3.68 implies that (1) is equivalent to (4) because Fr´echet spaces are metric spaces. Therefore the topological closure of the image of E in F is equal to its limit points (for the topology of F ). And since F b is a proper bornological vector space (because it is normal) we can apply Theorem 3.15 to deduce that the bornological closure of E in F agrees with its bornological l limit point. Lemma 3.71. If a short sequence of bornological Fr´echet spaces is strictly exact in CBornk then is strictly exact in Tck. 0 Ñ E Ñ F Ñ G Ñ 0 0 Ñ Et Ñ F t Ñ Gt Ñ 0 Proof. The p´qt functor is a left adjoint, so it preserve all colimits. Therefore, we have to check that it preserves kernels of morphisms of Fr´echet spaces. Let’s consider the kernel of the strict morphism f t : F t Ñ Gt. kerpf tq is a Fr´echet space, and therefore it is a normal locally convex space. Hence, we can apply Proposition 3.27 to deduce that kerpf tq – kerpfqt – Et, and the lemma is proved. l One problem that complicates our work is that the converse statement of Lemma 3.71 does not hold in general. However, we have the following result. 30 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Lemma 3.72. A short sequence of nuclear Fr´echet spaces 0 Ñ E Ñ F Ñ G Ñ 0 is strictly exact in Tck if and only if 0 Ñ Eb Ñ F b Ñ Gb Ñ 0 is strictly exact in CBornk. Proof. By Lemma 3.71 the strict exactness of the second sequence implies the strict exactness of the first one. Applying the functor p´qb to the first sequence we obtain a strict exact sequence because p´qb is a right adjoint functor. Since the sequence 0 Ñ Eb Ñ F b Ñ Gb Ñ 0 0 Ñ Eb Ñ F b Ñ Gb is manifestly algebraically exact it remains to check that F b Ñ Gb is a strict epimorphism. By the nuclearity hypothesis on E, F and G we can apply Lemma 3.67 to deduce that the von Neumann bornology of these spaces coincides with the compactoid one. Finally, we can l apply Lemma 3.66 to conclude the proof. Remark 3.73. The functor p´qb does not preserve strict exactness of short exact sequences in general. It is known that there exist Montel-Fr´echet spaces V whose quotient V {W by a closed subspace W is not Montel. Since for such a V one has V b – CptpV q whereas pV {Wqb fl CptpV {Wq, Lemma 3.66 implies that CptpV {Wq – V b W b . So, p´qb does not preserve cokernels in this case. 3.6. Derived functors of the inverse limit functor. We assume the reader is familiar with the notion of a family of injective objects with respect to a functor between quasi- abelian categories, in the sense of Schneiders (Definition 1.3.2 of [36]). In this section we recall how to derive the inverse limit functors in quasi-abelian categories, and then focus on the case of CBornk. Conditions for the existence of the derived functor of the inverse limit functors are given the Proposition 3.74, which is proven in Prosmans [32]. We will discuss then a bornological version of the classical Mittag-Leffler Lemma for Fr´echet spaces, which is an important example of the use of homological algebra towards functional analysis. There is an extensive literature on the study of the derived functor in the category of locally convex spaces: Palamadov [26], Vogt [42], and Retakh [34] are only some of the main contributions. In this section we bound ourself to obtain bornological versions of the basic results for Fr´echet spaces, without looking for the utmost generality. Proposition 3.74. Let I be a small category and C a quasi-abelian category with exact products. Then, the family of objects in CI op -acyclic which are Roos-acyclic form a limÐ iPI family. In particular, the functor : CI op Ñ C limÐ iPI 8ź i“0 ∆ÝÑ Vi 8ź Vi i“0 (3.6) defined by is right derivable to a functor and for any object V P CI op (3.5) D`pCI opq Ñ D`pCq , we have a canonical isomorphism Vi – RoospV q R limÐ iPI where the right hand side is the Roos complex of V . Proof. See Section3.3 of [32]. 31 l -acyclic objects form a family of injective objects relative Corollary 3.75. The family of limÐ iPI to the functor limÐ iPI : CI Ñ C. In the case of a tower (i.e. when I “ N with its natural order), the situation is easier to deal with. Consider a functor V : Nop Ñ C where C is a quasi-abelian category with exact products. We can consider the complex in degree 0 and 1 ∆p. . . , a2, a1, a0q “ p. . . , a2 ´ a3, a1 ´ a2, a0 ´ a1q where ai denotes the image of ai P Vi inside Vi´1. In this particular case, the Roos complex reduces to the complex of equation (3.6). Lemma 3.76. Let C be a quasi-abelian category with exact products. Consider a functor Vi is isomorphic to the complex in degree 0 and 1 given in equation V : Nop Ñ C. Then RlimÐ iPN (3.6). Proof. Since the cardinality of the natural numbers is less than the second infinite Viq “ 0 for all n ě 2. The fact that cardinal, Theorem 3.10 of [33] implies that LH npRlimÐ iPI RlimÐ iPI Vi is represented by (3.6) follows from the general definition of the Roos complex and l the isomorphism in (3.5). Lemma 3.77. Let V be a projective system of Fr´echet spaces in Tck indexed by N where all system morphisms are dense. Then, the complex (3.7) is strictly exact. 0 Ñ limÐ iPN Vi Ñ ź iPN ź iPN ∆Ñ Vi Vi Ñ 0 32 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Proof. We can first show exactness in the category Vectk by applying the standard Mittag-Leffler theorem for (topological) Fr´echet spaces (see for example [13], Lemma 1 of page 45) to 0 V1 V1 0 0 V2 0 V3 V1 V2 V1 V2 V3 V1 V1 V2 0 0 0 The lemma in [13] discusses only in the case k “ C, but it is easy to check the proof works over any base field since the only hypothesis used is that the spaces involved are endowed with a metric for which they are complete. It is an immediate consequence of the open mapping theorem for Fr´echet spaces (see [18] Section4.4.7, page 61) that the morphism ∆ is l a strict epimorphism and therefore the sequence is strictly exact. We can now state our bornological version of the Mittag-Leffler Lemma. Lemma 3.78. (Mittag-Leffler) Let E, F, G P CBorn by i P N. Let (3.8) Nop k 0 Ñ tEiuiPN ηÑ tFiuiPN ψÑ tGiuiPN Ñ 0 , be projective systems of bornological Fr´echet spaces over k indexed be a short exact sequence of systems where each ηi and ψi is strict in CBornk. Suppose also, that (1) tEiuiPN is an epimorphic system; (2) limÐ iPN Fi and limÐ iPN Then, the resulting sequence Ei, limÐ iPN Gi are nuclear bornological spaces. 0 Ñ limÐ iPN Ei Ñ limÐ iPN Fi Ñ limÐ iPN Gi Ñ 0 is strictly exact in CBornk, where the limits are calculated in CBornk. Proof. The datum of the short exact sequence (3.8) is equivalent to a sequence of strictly exact sequences i ÞÑ p0 Ñ Ei ψiÝÑ Gi Ñ 0q whose morphisms are compatible with the system morphisms of the projective systems. Then, the application of the functor p´qt to these sequences i Ñ 0q i ÞÑ p0 Ñ Et ψiÝÑ Gt ηiÑ Fi ηiÑ F t i i   o o   o o   o o     o o   o o   o o     o o     o o o o       o o   o o   o o o o o o o o 33 j Ñ Et yields strictly exact sequences in Tck, thanks to Lemma 3.71. The system maps Et i have topologically dense set theoretic image by Observation 3.21. Therefore, we can apply iuiPN to get the Mittag-Leffler lemma for Fr´echet spaces (cf. Lemma 3.77) to the systems tEt a strictly exact sequence in Tck i q Ñ limÐ pF t iPN and p´qt commute in (3.9). Applying of Fr´echet spaces. By Proposition 3.27, the functors limÐ iPN Lemma 3.72 we deduce that the strict exactness of the sequence iq Ñ limÐ pEt iPN 0 Ñ limÐ iPN iq Ñ 0 pGt (3.9) 0 Ñ plimÐ iPN Eiqt Ñ plimÐ iPN Fiqt Ñ plimÐ iPN Giqt Ñ 0 implies the strict exactness of the sequence 0 Ñ limÐ iPN Ei Ñ limÐ iPN Fi Ñ limÐ iPN Gi Ñ 0, l concluding the proof. Corollary 3.79. Let V be a projective system of bornological Fr´echet spaces indexed by N where all system morphisms are dense and limÐ Vi is nuclear. Then, the complex (3.10) ź ź Vi Ñ 0 Vi Ñ ∆Ñ Vi 0 Ñ limÐ iPN iPN iPN is strictly exact. Proof. Using the bornological version of the Mittag-Leffler Lemma, i.e. Lemma 3.78, we l can use the same proof of Lemma 3.77 to deduce the corollary. Corollary 3.80. Let V be a projective system of bornological Fr´echet spaces indexed by N where all system morphisms are dense and limÐ Vi is nuclear. Then (3.11) Vi. RlimÐ iPN Vi – limÐ iPN Proof. Corollary 3.79 is equivalent to say that the Roos complex of V has cohomology l only in degree 0. Therefore, Proposition 3.74 implies (3.11). Lemma 3.81. In any quasi-abelian category C, a complex Ñ V n`1 dn`1Ñ V n dnÑ V n´1 Ñ is strictly exact if and only if the sequences (3.12) 0 Ñ kerpdnq Ñ V n Ñ kerpdn´1q Ñ 0 are strictly exact for each n. Given a projective system of strictly exact complexes (3.13) Ñ V n`1 i dn`1 iÑ V n i dn iÑ V n´1 i Ñ 34 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER where both tkerpdn i quiPN and tV n -acyclic systems, the projective limit i uiPI are limÐ iPN limÐ dn`1 iÑ limÐ iPN V n`1 i V n i limÐ dn iÑ limÐ iPN i Ñ V n´1 Ñ limÐ iPN is strictly exact. Proof. By Corollary 1.2.20 of [36], the complex (3.13) is strictly exact if and only if LH npV q “ 0 for all n. This is equivalent to the two term complexes 0 Ñ coimpdn`1q Ñ kerpdnq Ñ 0 vanishing in the left heart of C (which is a full subcategory of the derived category, cf. Corollary 1.2.20 of ibid.). By this explicit description of the left heart, the condition LH npV q “ 0 is equivalent to that the canonical morphism coimpdn`1q Ñ kerpdnq is an isomorphism. Therefore, the strictly exactness of sequences 0 Ñ kerpdnq Ñ V n Ñ coimpdnq Ñ 0 implies the strictly exactness of the sequences (3.12). For the second statement, observe that the sequences 0 Ñ ker dn Ñ limÐ iPN i Ñ ker dn´1 Ñ 0 V n are strictly exact, being the application of RlimÐ iPN i Ñ V n i Ñ ker dn´1 i Ñ 0 -acyclic objects, thought of as an exact triangle in D`pC 0 Ñ ker dn Nq. of limÐ iPN to the strict short exact sequences l 4. Stein domains 4.1. Bornological Fr´echet algebras. The notion of multiplicatively convex bornological algebra (in short m-algebra) introduced in [4] (cf. Definition 4.1 of ibid.) is not general enough for all purposes of analytic geometry. For example, the bornological Fr´echet algebras of analytic functions on open subsets of An k are not multiplicatively convex in this sense. So, we introduce here a generalization of multiplicatively convex bornological algebras which encompass also bornological Fr´echet algebras. We start by recalling what the spectrum of a general bornological algebra is. Definition 4.1. Let A be a bornological algebra, i.e. an object of CommpBornkq, we define the spectrum of A as the set of equivalence classes of bounded algebra morphisms of A to valued extensions of k. The spectrum is denoted by MpAq and it is equipped with the weak topology, i.e. the weakest topology for which all maps of the form χ ÞÑ χpfq P Rě0, for all f P A, are continuous. This definition extends the one given in Definition 4.2 of [4] for m-algebras, but the spectrum of a general bornological algebra is not as well behaved as the spectrum of a bornological m-algebra. For example, there exist bornological algebras whose underlying bornological vector space is complete and whose spectrum is empty. So, it is important to single out a suitable sub-category of the category of bornological algebras for which the notion of spectrum is not pathological. Definition 4.2. Let A be an object of CommpCBornkq. We say that A is pro-multiplicatively convex (or a pro m-algebra) if there is an isomorphism 35 A – limÐ iPI Ai in CommpBornkq, where I is a cofiltered small category and Ai are complete bornological m-algebras. Definition 4.3. Let A be a pro-multiplicatively convex object of CommpCBornkq. A is called densely defined if there exists an isomorphism A – limÐ Ai of bornological algebras as iPI in Definition 4.2 such that for any i P I the canonical map πi : A Ñ Ai has dense set-theoretic image. We remark that the condition in Definition 4.3 can be stated purely categorically by requiring that the morphisms πi : A Ñ Ai are epimorphisms of the underlying complete bornological vector spaces (cf. Proposition 3.8). Proposition 4.4. Let A be a densely defined pro-multiplicatively convex bornological algebra, such that with An Banach algebras. Then, A is a bornological Fr´echet algebra whose spectrum coincides with the spectrum of MpAtq, i.e. An A – limÐ nPN ď nPN MpAq “ MpAnq Proof. A is a bornological Fr´echet algebra, i.e. the underlying bornological vector space nq is a Fr´echet space as consequence pAt Anqt “ At. Thus, topologically. of A is that of a bornological Fr´echet space. Indeed, limÐ nPN pAt nq – plimÐ of Proposition 3.27, which states that limÐ nPN nPN nqbq – A. ppAt pAtqb – limÐ nPN ď because πi : A Ñ Ai is an epimorphism if and only if πt the Ai are Banach. The identification MpAnq MpAq – MpAtq “ MpAtq “ nPN ď nPN This implies that a character A Ñ K is bounded if and only if it is continuous for At. The densely defined hypothesis implies that MpAnq i : At Ñ At i is an epimorphism, since 36 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER is a well-known result of the classical theory of Frech´et algebras (e.g. , see Section2.5 of [2]). l We end this section recalling a result from [3] that will be used in the next subsection. Definition 4.5. Let E be a separated bornological vector space of convex type over k. A pair pV, bq consisting of mappings V : is called a bornological web if all of the conditions below hold: Nj Ñ PpEq and b : NN Ñ pkqN jPN (1) The image of V consists of disks. (2) Vp∅q “ E. (3) Given a finite sequence pn0, . . . , njq, then Vpn0, . . . , njq is absorbed by ď Ť ř Vpn0, . . . , nj, nq. nPN nPN (4) For every s : N Ñ N the series λpsqnxn, with λpsqn P k, converges bornologically in E, whenever we choose xn P Vpsp0q, . . . , spnqq and λpsqn “ bpsqn. Lemma 4.6. Let A, B be objects of CommpBornkq for which the underlying bornological vector space of A is complete and the one of B is a webbed bornological vector space. Let φ : A Ñ B be a morphism of the underlying objects in CommpVectkq. Suppose that in B there is a family of ideals (cid:61) such that (1) each I P (cid:61) is (bornologically) closed in B and each φ´1pIq is closed in A; (2) for each I P (cid:61) one has dimk B{I ă 8; (3) Ş I “ p0q. Then, φ is bounded. IP(cid:61) l Proof. The proof can be found in [3], Proposition 4.23. And finally the following lemma permits us to apply Lemma 4.6 to the morphisms of bornological algebras we will study next sub-section. Lemma 4.7. The underlying bornological vector space of every k-dagger affinoid algebra is a webbed bornological vector space. Proof. The assertion about dagger affinoid algebras is a direct consequence of Example 2.3 (2) of [3]. l Remark 4.8. The notion of pro-multiplicative algebra is more general than what strictly needed in this paper. However, this notion is very comfortable when both Fr´echet algebras and LB algebras are considered in the same discussion. Moreover, the material in this section will be a reference for future works in which we will analyse dagger quasi-Stein algebras for which the use of the notion of pro-multiplicative bornological algebra is unavoidable. 4.2. Stein algebras and Stein spaces. Definition 4.9. A Weierstrass localization of a k-dagger affinoid algebra A is a morphism of the form A Ñ Axr´1 1 X1, . . . , r´1 n Xny: pX1 ´ f1, . . . , Xn ´ fnq for some f1, . . . , fn P A, priq P Rn`. Definition 4.10. Let A, B, C be a k-dagger affinoid algebras such that B and C are A- algebras. Let f : B Ñ C be a bounded morphism of A-algebras. f is called inner with respect to A if there exists a strict epimorphism π : Axr´1 n Tny: Ñ B such that 1 T1, . . . , r´1 37 ρCpfpπpTiqqq ă ri for all 1 ď i ď n, where ρC is the spectral semi-norm of C. Definition 4.11. Let φ : MpAq “ X Ñ MpBq “ Y be a morphism of k-dagger affinoid spaces. The relative interior of φ is the set IntpX{Y q “ tx P XA Ñ Hpxq is inner with respect to Bu. The complement of IntpX{Y q is called the relative boundary of φ and denoted by BpX{Y q. If B “ k, the sets IntpX{Y q and BpX{Y q are denoted by IntpXq and BpY q and called the interior and the boundary of X. Definition 4.12. A dagger Stein algebra over k is a complete bornological algebra A over k which is isomorphic to an inverse limit of k-dagger affinoid algebras ÝÑ A4 ÝÑ A3 ÝÑ A2 ÝÑ A1 ÝÑ A0 (4.1) in the category CBornk where each morphism is a Weierstrass localization and MpAiq is contained in the interior of MpAi`1q, for each i. The category of dagger Stein algebras is the full sub-category of CommpCBornkq identified by dagger Stein algebras over k and it is denoted by Stnk. ď Definition 4.13. A k-dagger analytic space X is called dagger Stein space if it admits a dagger affinoid covering U1 Ă U2 Ă such that Ui X “ and the restriction morphisms OXpUi`1q Ñ OXpUiq are Weierstrass localizations and Ui is contained in the interior of Ui`1, for each i P N. The category of dagger Stein spaces is the full sub-category of the category of k-analytic spaces of this form. iPN Example 4.14. (1) The open polycylinders are the most basic example of Stein spaces, whose exhaustion by Weierstrass subdomains is given by the closed polycylinders of smaller radius. (2) Analytifications of algebraic varieties are Stein spaces, both for Archimedean and non-Archimedean base fields. (3) Closed subspaces of Stein spaces are Stein spaces (see [23]). Lemma 4.15. Let A be a dagger Stein algebra over k. Then, A is a densely defined, pro- multiplicatively convex, bornological Fr´echet algebra over k. Ai be a presentation of A as in definition of dagger Stein algebra. Let Ai Proof. Let limÐ iPN be the completion of Ai with respect to the norm }f} “ inf BPDAi fpAiqB , 38 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER where pAiqB denotes the gauge semi-norm, as explained in Remark 2.17. Notice that }} is a norm because by hypothesis Ui has non-empty interior, therefore }f} “ 0 ñ f “ 0 (more precisely, since Ui has non-empty interior one can find an injective bounded morphism of Ai to a Banach algebra, which implies that Ai is separated). The requirement of MpAiq to lie in the interior of MpAi`1q can be restated by saying that there are (bounded) diagonal morphisms such that the following diagram is commutative A3 A3 A2 A2 A1 A1 (4.2) where the vertical maps are the completions. Hence, following the diagonal maps we obtain a projective system (4.3) Ñ A3 Ñ A3 Ñ A2 Ñ A2 Ñ A1 Ñ A1, which shows that the sub-system Ñ A3 Ñ A2 Ñ A1 is cofinal in (4.3), and hence (4.4) Then, by Proposition 3.27 we have limÐ iPN of Banach spaces in Tck is a Fr´echet spaces and hence normal (see Example 3.22). So, from the relation A – limÐ Ai. iPN p Aiqqt, because a countable projective limit iq – plimÐ p At iPN A “ limÐ iPN p Aiq – plimÐ iPN p Aiqqtb – plimÐ iPN p At iqqb we can deduce that A is a Fr´echet space in Bornk, in the sense of Definition 3.23. Moreover, it is easy to see all the maps of the system (4.4) have dense images because in the diagram (4.2) all maps but the bottom horizontal ones are clearly epimorphisms, which implies that l the bottom horizontal ones are epimorphism too. Remark 4.16. For non-Archimedean base fields, we could have defined Stein algebras us- ing classical affinoid algebras in place of dagger affinoid algebras (which is what we get in Lemma 4.15 with the algebras Ai). The cofinality argument of Lemma 4.15 shows that the bornological structure we defined on a dagger Stein algebra agrees with the Fr´echet struc- tures defined classically, for example in [7], [11], [12] for the non-Archmedean theory and in [16] for the Archimedean theory. Lemma 4.17. Let A be a dagger Stein algebra over k. Then, the underlying bornological vector space of A is binuclear. Proof. By Proposition 3.53 the underlying bornological vector spaces of dagger affinoid algebras are nuclear. We can apply Proposition 3.51 to any system of the form (4.1) to 39 i pAt the locally convex space At is nuclear. Notice that if we chose a presentation A – limÐ iPN we have that At – limÐ iPN deduce that the underlying bornological vector space of A is nuclear. It remains to see that Ai, iq, as a consequence of Proposition 3.27, and by Lemma 3.59 At are nuclear locally convex spaces. This implies that At is nuclear because projective limits of nuclear locally convex spaces are always nuclear (for a proof of this fact we can refer to [13] Proposition 3.2 at page 23 in the case in which k is Archimedean, and for the non- l Archimedean case to Theorem 8.5.7 of [27]). Lemma 4.18. Let A be a dagger Stein algebra over k. Then, MpAq is a dagger Stein space over k. Proof. The spectrum MpAq is a hemi-compact topological space because, by Lemma MpAiq (topologically), for a 4.15, we can apply Proposition 4.4 to deduce that MpAq “ Ai (because, in the notation of Lemma 4.15, MpAiq – Mp Aiq). The presentation A – limÐ iPN condition of MpAiq being contained in the Berkovich interior of MpAi`1q readily implies that tMpAiquiPN is a cofinal family of compact subsets of MpAq (cf. Exercise 3.8.C (b) of [14]). This family is therefore a Berkovich net on MpAq which induces a structure of k- dagger analytic space. It is straightforward to check that in this way MpAq is endowed with a structure of dagger Stein space because by the definition of dagger Stein algebra the mor- phisms Ai`1 Ñ Ai are Weierstrass localizations and MpAiq are mapped into IntpMpAi`1qq l by these localizations. Lemma 4.19. Let A be a dagger Stein algebra over k. Then, we can always find a presen- tation Ť iPN A – limÐ iPN Ai as in Definition 4.12 such that Ai are strictly dagger affinoid algebras. Proof. Let A – limÐ Ai be a presentation of A as in the definition of dagger Stein algebra. It iPN is enough to show that each localization Ai`1 Ñ Ai factors through Weierstrass localizations Ai`1 Ñ Bi Ñ Ai where Bi are strictly dagger affinoid. It is easy to check that Lemma 2.5.11 of [7] holds also for dagger affinoid algebras. Therefore, for any 0 ă  ă 1 we can find a ρ “ pρiq P Rn` and a strict epimorphism π : kxρ´1 n Xny: Ñ Ai`1 1 X1, . . . , ρ´1 such that the image of MpAiq is contained in the image of MpAxpρ1q´1πpX1q, . . . ,pρnq´1πpXnqy:q in MpAi`1q. Since the diagram n Xny: 1 X1, . . . , ρ´1 kxρ´1 π Ai`1 π1 kxpρ1q´1X1, . . . ,pρnq´1Xny: Ai`1xpρ1q´1πpX1q, . . . ,pρnq´1πpXnqy: 40 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER a a is a push-out square, we see that the morphism π1 is a strict epimorphism, because tensoring preserves strict epimorphisms. Therefore, choosing pρiq´1 P k for all i, which is always a k is dense in R, we see that Ai`1xpρ1q´1πpX1q, . . . ,pρnq´1πpXnqy: possible choice because can always be found to be a strictly dagger affinoid algebra and this is our choice for Bi. Moreover, the morphism Ai`1 Ñ Bi is by construction a Weierstrass subdomain embedding and Ai`1 Ñ Ai is by hypothesis a Weierstrass subdomain embedding which immediately implies that Bi Ñ Ai is a Weierstrass domain embedding, concluding the proof. l Remark 4.20. In Lemma 4.19 we only discussed how to find a presentation of a dagger Stein algebra as an inverse limit of strictly dagger affinoid algebras, but the same argument clearly works in the classical affinoid case. Notice that this lemma crucially depends on the hypothesis that k is non-trivially valued. Lemma 4.21. Let A be a dagger Stein algebra over k and m Ă A a finitely generated maximal ideal. Then, m is bornologically closed in A. Proof. Let A – limÐ Ai be a presentation of A as given by Lemma 4.19. Consider the iPN extensions of m Ă A to Ai, denoted mAi, for every i P N. There are two possible cases to consider: either mAi is a proper ideal of Ai, for some i, or mAi “ Ai for all i. In the i pmAiq must be a proper ideal containing m, and first case, since m is maximal in A then π´1 therefore must be equal to m. Since mAi is bornologically closed in Ai (cf. Theorem 4.9 (2) of [4]), πi is a bounded map and the pre-image of a bornologically closed set by bounded maps is a bornologically closed set, then m is bornologically closed in A. On the other hand, suppose that for every i, mAi “ Ai and let f1, . . . , fn denote a set of generators of m. Since mAi “ Ai, πipfiq have no common zeros in MpAiq. Because this is MpAiq, we can deduce that f1, . . . , fn have no common zeros true for any i and MpAq “ in MpAq. Using Theorem A for dagger Stein spaces (cf. Theorem 3.2 of [17] for the case in which k is non-Archimedean, instead for k Archimedean the classical theorem of Cartan applies) and reasoning in the same way as Theorem V.5.4 and Theorem V.5.5 at page 161 of [16], we can deduce that condition that f1, . . . , fn have no common zeros implies that there figi and hence m “ A. But this is impossible because by hypothesis m is a proper ideal of A. This contradiction shows that there must exist an i such that mAi ‰ Ai. l exist g1, . . . , gn P A such that 1 “ nř i“1 Ť iPN Indeed, dagger Stein algebras can have non-finitely generated maximal ideals. Remark 4.22. The second part of the proof of Lemma 4.21 works only for fintely generated ideals. It follows that such ideals are necessarily non-closed and bornologically dense subsets. Remark 4.23. Notice that the morphism of Grothendieck locally ringed spaces MpAiq Ñ MpAq, induced from the projections πi : A Ñ Ai discussed so far, is an open immersion (in the sense of Definition 4.18 of [2]), when one consider the Berkovich G-topology of analytic domains (cf. chapter 6 of [2] or chapter 1 of [8] for details about this G-topology). This is because we endow MpAq with the structure of k-dagger analytic space given by the iPN MpAiq. Therefore, each MpAiq is an analytic domain in MpAq and MpAiq Berkovich net is identified with its image in MpAq. Lemma 4.24. Any morphism of k-dagger Stein algebras comes from a morphism in PropAfnd: kq by applying the functor limÐ : PropAfnd: kq Ñ CommpBornkq. 41 Ť Ai, Proof. Let f : A Ñ B be a morphism of dagger Stein algebras and let A – limÐ iPN B – limÐ Bj be two fixed presentations as in the definition of dagger Stein algebra. Let jPN Mpfq : MpBq Ñ MpAq denote the morphism of dagger Stein spaces induced by f . Since the family tMpAiquiPN is cofinal in the family of compact subsets of MpAq, every morphisms MpBiq Ñ MpAq must factor through a MpAiq, because the image of MpBiq in MpAq is compact. Therefore, we get morphisms of k-dagger analytic spaces Mpfiq : MpBiq Ñ MpAiq (possibly by re-indexing the systems in a suitable way). These morphisms commute with the morphisms Mpfq : MpBq Ñ MpAq and limÑ Mpfiq “ Mpfq. By Remark 4.16 of [4], there iPN exist maps fi : Ai Ñ Bi of k-dagger affinoid algebras induced by the morphisms of k-dagger affinoid spaces Mpfiq and for these maps we get limÐ l iPN fi “ f , as required. Theorem 4.25. The category of dagger Stein algebras over k is anti-equivalent to the cat- egory of dagger Stein spaces over k. Moreover, the forgetful functor from the category of dagger Stein algebras over k to CommpVectkq is fully faithful. Proof. It is clear that to any dagger Stein algebra one can associate a dagger Stein space and vice versa functorially, as explained so far (cf. Lemma 4.18). So, we check the claim about the boundedness of every algebra morphism between dagger Stein algebras. Fix two dagger Stein algebras with representations A – limÐ Bi, and let φ : A Ñ B be a iPN morphism of their underlying objects in CommpVectkq. We also suppose that all Ai and Bi are strictly dagger affinoid algebras, which is always a possible choice (cf. Lemma 4.19). The data of the morphism φ is equivalent to a system of morphisms φi : A Ñ Bi in CommpVectkq with obvious commuting relations with the system morphism of the representation of B we fixed. To show that φ is bounded, it is enough to show that each φi is bounded. Let m Ă Bi be a maximal ideal. Then, m is finitely generated and Bi m is a finite valued extension of k, because Bi is strictly dagger affinoid. Then, the composition of maps Ai, B – limÐ iPN A Ñ Bi Ñ Bi m identifies A φ´1 pmq with a sub-ring of Bi i m so k Ă . i pmq is a maximal, finitely generated This implies of course that ideal of A, which by Lemma 4.21 must be closed. It follows that the quotient bornology of i pmq Ă Bi A φ´1 pmq is a field. Therefore φ´1 A φ´1 m i 42 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER i pmq is separated, and hence complete and since A pmq as vector space over k is finite φ´1 dimensional, its bornology is necessarily isomorphic to the product bornology of finitely many copies of k. Lemma 4.18 yields A φ´1 i which readily implies MpAq “ MaxpAq “ MpAiq MaxpAiq, ď ď iPN iPN where on the left-hand side only finitely generated maximal ideals are considered. So, there i pmqAj is a maximal ideal of Aj. It follows that there exists a j such that for any j ě i, φ´1 is a canonical isomorphism i pmq – A φ´1 Now consider mn, for any n P N. By elementary commutative algebra one can see that φipmqn Ă φipmnq which implies that there exist canonical quotient maps Aj i pmqAj φ´1 . i pmqn Ñ A A Ñ A i pmnq . φ´1 φ´1 Since MpAjqãÑMpAq is an immersion (i.e. induces isomorphisms on stalks), using the same argument of proposition 7.2.2/1 of [10] we can deduce that i pmqn – A φ´1 Aj i pmqnAj φ´1 , @n P N. Aj pmqnAj are a finite dimensional k-Banach algebra. Therefore, we wrote Notice also that φ´1 i i Ť Ť A is a quotient of a finite dimensional k-Banach algebra, which is necessarily a k- φ´1 pmnq i pmnq is closed in A, for any maximal ideal of Bi and any n P N. Banach. This shows that φ´1 We showed that the family of all powers of maximal ideals of Bi satisfies the hypothesis of Lemma 4.6 (that can be applied to φi thanks to Lemma 4.7), proving that φi and hence φ are bounded maps. To conclude the proof it remains to show that every morphism of dagger Stein spaces is induced by a morphism of dagger Stein algebras. A morphism X Ñ Y of dagger Stein spaces is by definition a morphism of locally ringed Grothendieck topological spaces between X “ MpAiq, when they are equipped with their maximal Berkovich nets generated by the nets tMpAiquiPN and tMpBiquiPN (cf. chapter 6 of [2]). All the elements of the maximal nets which gives the structure of k-analytic spaces to X and Y must be compact subsets of X and Y respectively, and hence by the fact that X and Y are hemi-compact, all the elements of the maximal nets must be contained in some MpAiq or MpBiq, respectively, as explained in the proof of Lemma 4.18. Therefore, every morphism of k-dagger analytic spaces must come from a morphism of systems i ÞÑ pAi Ñ Biq, of the form we described in Lemma 4.24. This shows that to give a morphism between dagger Stein l algebras is the same as to give a morphism of dagger Stein spaces and vice versa. MpBiq and Y “ iPN iPN 43 The following corollary is our version of the classical Forster’s theorem. A similar state- ment holds for the category of affinoid algebras or in the dagger affinoid setting as discussed in Remark 4.18 of [4]. Corollary 4.26. The functor defined by M from the opposite category of k–dagger Stein algebras to the category of locally ringed Grothendieck topological spaces is fully faithful. Therefore, we can identify the category opposite to k-dagger Stein algebras, with a full sub- category of the category of locally ringed Grothendieck topological spaces. Thanks to Theorem 4.25 we can give the following definition. Definition 4.27. A morphism A Ñ B between dagger Stein algebras is called a localization if it can be written as a projective limit of dagger localizations. Definition 4.28. A morphism f : X Ñ Y between dagger Stein spaces is called an open immersion if it is a homeomorphism of X with fpXq and for each x P X the induced morphism of stalks OY,fpxq Ñ OX,x is an isomorphism. Remark 4.29. Using Theorem 4.25 and Proposition 4.22 of [4], one can see that A Ñ B is a localization of Stein algebras if and only if the associated morphism of dagger Stein spaces MpBq Ñ MpAq is an open immersion. 5. Homological characterization of the Stein topology We now need to prove last lemmata for deducing the main results of this work. ApEq be the Bar resolution (see Lemma 5.1. Let A be an object of CommpCBornkq and L ‚ Definition 2.11) of an object E P ModpAq. If both A and E are nuclear, as object of CBornk, then L ‚ ApEq is isomorphic to E in Dď0pAq. Theorem 3.50, nuclear objects of CBornk are flat for pbk. ApFq “ ApbkpA L n Proof. It is enough to check that we can apply Lemma 2.9. This is possible because, by l Let A P CommpCBornkq and E, F P ModpAq. In Definition 2.11 we introduced the Bar whose differentials are given in equation (2.1). Applying to this complex the functor EpbAp´q Ñ EpbAL 2 which is a representative of EpbL EpbAL n kpbkFq “ ApbkpApbk pbkA loooooomoooooon ApFq Ñ L 0 resolution of F as the complex Ñ L 2 pbn ApFq Ñ EpbAF Ñ 0 ApFq Ñ EpbAL 0 ApFq Ñ EpbAL 1 kpbkFq – EpbkA ApFq “ EpbAApbkpA kpbkF. pbn pbn ApE, Fq “ EpbkA kpbkF. pbn Remark 5.2. The terms of this complex can be written in a simplified form as follows ApFq Ñ F Ñ 0 pbkFq Therefore, we will introduce the notation (5.1) L n ApFq Ñ L 1 n times one obtains the complex AF in Dď0pAq. 44 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Lemma 5.3. Let A, B be dagger Stein algebras and let X “ MpBq, Y “ MpAq be the corresponding dagger Stein spaces. Let f : A Ñ B be a morphism and Mpfq : X “ MpBq Ñ Y “ MpAq the corresponding morphism of dagger Stein spaces. Let also C be another dagger Stein algebra and g : A Ñ C an arbitrary morphism in CommpCBornkq Then, if Mpfq is an open immersion there exist projective systems of Banach algebras tAiuiPN, tBiuiPN and tCiuiPN and morphisms of systems tfi : Ai Ñ BiuiPN, tgi : Ai Ñ CiuiPN such that Ci – C and the projection maps A Ñ Ai, B Ñ Bi, Ai – A, limÐ iPN Bi – B, limÐ (1) limÐ iPN iPN C Ñ Ci have dense images; (2) limÐ fi “ f , limÐ gi “ g; (3) for every i P N an isomorphism AipCi, Biq Ñ Ci L ‚ (5.2) pbAiBi of objects of Dď0pBiq. ď Proof. By Lemma 4.24 every morphism of Stein algebras can be represented as a system morphism of dagger affinoid algebras. So, let X “ Ui, Y “ Vi, Z “ MpCq “ iPN ď iPN ď Wi iPN iq and Wi “ MpC1 iq, Vi “ MpA1 i and C1 be representations of X, Y and Z, with Ui “ MpB1 iq where A1 i, B1 i are dagger affinoid algebras and the systems are chosen in a way that both morphisms f and A Ñ C have the same index set (for finite diagram without loops this re-indexing can always be performed; for more details see Proposition 3.3 of the Appendix of i, B1 [1]). Moreover, using Lemma 4.19 we can suppose that all A1 i are strictly dagger i Ñ B1 i and gi : A1 affinoid algebras. Consider the morphisms f1 i, induced by f iq are open immersions of dagger and g : A Ñ C respectively. It is easy to check that Mpf1 affinoid spaces. Applying Proposition 4.22 of [4] we can deduce that f1 i are dagger affinoid localizations. Applying Theorem 5.7 of [4], we obtain a strict isomorphism of complexes of B1 i-modules (5.3) The object C1 i – C1 B1 iq. pC1 i, B1 i can be represented by the complex L ‚ B1 A1 pC1 iq Ñ L 0 iq i, B1 A1 Ñ L 2 A1 iq Ñ L 1 A1 i and C1 pbA1 i Ñ C1 pbL pC1 i, B1 pC1 i, B1 pbL i : A1 (5.4) B1 i. i C1 i i A1 i A1 i i i i i i using notation of equation (5.1). It is clear that the explicit description of the differentials of the Bar resolution given in equation (2.1) implies that the Bar resolution of B1 i admits a uniform indexing as a diagram of ind-Banach modules, once one fixes a representation of B1 i as an ind-Banch object. Moreover, by Theorem 4.9 (4) of [4] morphisms of dagger affinoid algebras can always be written as morphisms of inductive systems of Banach algebras. pC2 ρqi Therefore, we can find representations of A1 ρqi and B1 ρqi, C1 pB2 pA2 i – limÑ ρą1 i – limÑ ρą1 i – limÑ ρą1 such that the isomorphism (5.3) lifts to an isomorphism of complexes Banach modules for any ρ ą 1 small enough. Indeed to choose such a ρ we can proceed as follows. Let ρ1 ą 1 be small enough such that f1 i is representable as a morphism of inductive systems i Ñ B1 i : A1 45 i Ñ C1 ρ1qi Ñ pB1 ρ2qi Ñ pC1 ρ1qi and let ρ2 ą 1 be such that gi : A1 pA2 i can be represented as a morphism of ρ2qi. Then, any ρ such that 1 ă ρ ă mintρ1, ρ2u is a suitable inductive systems pA1 choice. In fact, all the maps of the complex of equation (5.4) can be written as a morphism of inductive systems for such small ρ, because these maps are obtained from the differentials of the Bar resolution by tensoring pB2 ρqi-module) over with pC2 ρqi ρ2qi). Therefore, we have a (which becomes a pA2 quasi-isomorphism of complexes of Banach modules (5.5) ρqi (which becomes an pA2 ρ2qi Ñ pC2 ρqi-modules through the map pA2 r Ñ L 2pA2 ρqiqs Ñ pC2 ρqi for ρ small enough. Moreover, by the fact that the dagger subdomain embeddings ρqiq Ñ L 0pA2 ρqiq Ñ L 1pA2 ρqi,pB2 ρqi,pB2 pbpA2 ρqi,pB2 ρqippC2 ρqippC2 ρqippC2 ρqipB2 ρqi are inner for all i P N, we can find ρU ą 1, ρV ą 1, ρW ą 1 small enough such that Ui Ă Ui`1, Vi Ă Vi`1, Wi Ă Wi`1 MppB2 ρUqiq Ă Ui`1, MppA2 ρV qiq Ă Vi`1,MppC2 ρWqiq Ă Wi`1. Therefore, fixing any ρ ą 1 such that mintρU , ρV , ρW , ρ1, ρ2u ą ρ ą 1, we set for each i Ai “ pA2 ρqi, Bi “ pB2 ρqi, Ci “ pC2 ρqi. The properties (1), (2) and (3) satisfied by the system so obtained as direct consequences of l the choices we made. pbL pbAiBi Bi Ñ Ci Remark 5.4. Notice that the isomorphism (5.2) does not imply that Ci is an isomorphism in general. Because for Archimedean base fields Banach spaces might not be flat in CBornk and therefore the Bar resolution is not a flat resolution. Theorem 5.5. Let i : U Ñ V be an open immersion of dagger Stein spaces corresponding to a morphism AV Ñ BU of dagger Stein algebras over k. For any dagger Stein AV -algebra CW the morphism Ai pbL AV CW BU Ñ CW pbAV BU is an isomorphism in Dď0pBUq. Also (in the case that CW “ BU ) AV Ñ BU is a homotopy epimorphism. Proof. Lemma 5.3, applied to AV , BU , CW and the morphism i : U Ñ V yields three systems of Banach algebras such that Ai – AV , limÐ iPN Bi – BU , Ci – CW limÐ iPN limÐ iPN such that the opposite morphisms to i and the morphism AV Ñ CW can be written as morphisms of these projective systems. Moreover, for each i P N, Lemma 5.3 also yields to a strictly exact complex Ñ L 2 pbAiBi Ñ 0 AipCi, Biq Ñ L 0 AipCi, Biq Ñ L 1 AipCi, Biq Ñ Ci 46 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER The following commutative diagram pbkCj pbAj Cj Bj Bj pbkCi pbAiCi Bi Bi show that the bottom horizontal map is an epimorphism because the rest of the arrows are. Continuing to tensor we get a epimorphism L n to check that these morphisms commutes with differentials, giving a projective system of morphism of complexes AipCi, Biq for any n. It is easy AjpCj, Bjq Ñ L n d2 j d2 i L 1 AjpCj, Bjq L 1 AipCi, Biq d1 j d1 i L 0 AjpCj, Bjq L 0 AipCi, Biq d0 j d0 i Cj Ci pbkAi pbkBiq ÝÑ limÐ pCi pbkBiq ÝÑ limÐ pCi pbAj Bj pbAiBi pbAiBiq ÝÑ 0. 0 0 (5.6) of whose limit we denote by pCi (5.7) ÝÑ limÐ iPN Using the isomorphisms coimpdn`1 squares i iPN iPN q – kerpdn i q and coimpdn`1 q – kerpdn j jq the commutative (5.8) L n AjpCj, Bjq kerpdn´1 j q L n AipCi, Biq / kerpdn´1 i q are seen to have all arrows epimorphisms except the one on between the kernels. Therefore, it also is an epimorphism. Now both the systems tkerpdn epimorphic. Notice that AipCi, BiquiPN are i quiPN and tL n i q. dn i q – kerplimÐ iPN kerpdn limÐ iPN AipCi, Biq is nuclear. Therefore, limÐ L n iPN i q is nuclear Proposition 3.51 implies that limÐ iPN AV pCW , BUq. Therefore, Corollary 3.80 as it can be identified with a closed subspace of L n implies that tkerpdn i quiPN and tL n AipCi, BiquiPN are limÐ iPN kerpdn acyclic. Applying Lemma 3.81 the to the system of complexes (5.6) we deduce that complex of equation (5.7) is strictly acyclic. Moreover, applying Corollary 3.64 to each term in degree strictly less than zero of the complex (5.6), we see that (5.7) is strictly isomorphic to the / /     / complex (5.9) ÝÑ plimÐ iPN CiqpbkplimÐ iPN Thus we showed that the complex ÝÑ CW is strictly exact. Since cokerpL 1 the isomorphism in Dď0pBUq CW AV iPN iPN pCi pCi pbAiBiq ÝÑ 0. Biq ÝÑ plimÐ iPN pbkBU ÝÑ CW AiqpbkplimÐ pbkAV pbL pbAiBi Ñ Bi is an isomorphism by Theorem 5.11 of [4]. We can infer pbL Biq ÝÑ limÐ iPN pbAiBiq ÝÑ 0 pbAV BU we have shown pbAV BU . CiqpbkplimÐ pbkBU ÝÑ limÐ pbAiBiq – CW pbAiBiq – limÐ AV pCW , BUq Ñ L 0 BU Ñ limÐ iPN pCi AV pCW , BUqq – CW Bi – BU BU AV iPN BU Ñ limÐ iPN pBi iPN pbL In the special case that the system tBiuiPN and the system tCiuiPN are equal we have BU “ CW and that Bi that Bi – Bi Ai is an isomorphism, concluding the proof of the theorem. 47 l Lemma 5.6. Let A be a dagger Stein algebra presented by a system of (strictly) dagger affinoid algebras Ai. Then the projection maps A Ñ Ai are homotopy epimorphisms. Proof. We can write Ai as a direct limit Ai – limÑ ρą1 pAiqρ where pAiqρ are Stein algebras of Stein spaces that admit closed embeddings in polydisks and where each morphism pAiqρ Ñ pAiqρ1, for ρ1 ă ρ corresponds geometrically to an open embedding. Such a system of Stein spaces can be found via a presentation of Ai – W n I and writing W n k as the direct limit of the Stein algebras of open polydisks of radius bigger that one, which form a base of neighborhoods of the closed unital polydisk. Applying Theorem 5.5 we can infer that each pAiqρ Ñ pAiqρ1 is a homotopy epimorphism and applying Lemma 2.12 we can deduce that the canonical morphisms pAiqρ Ñ Ai are homotopy epimorphisms. Since A is a Stein algebra and A Ñ Ai corresponds geometrically to an open embedding there exists a ρ ą 1 small enough such that A Ñ Ai factors through A Ñ Aρ and that A Ñ Aρ corresponds geometrically to an open embedding. Applying Theorem 5.5 we can deduce that A Ñ Aρ is a homotopy epimorphism and therefore A Ñ Ai is a homotopy epimorphism because it can be written as a composition of two homotopy epimorphisms. l k And finally, the last result of characterization of open Stein immersions. Theorem 5.7. Let f : AV Ñ BU be a morphism of dagger Stein algebras over k. If f is a homotopy epimorphism as morphism of CommpCBornkq, then it is a localization. Proof. The condition of f being a homotopy epimorphism means that pbL AV BU pbAV BU – BU . BU – BU 48 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Ai – AV , limÐ iPN Bi – BU be representations of AV and BU such that f can be written Let limÐ iPN as a morphism of projective systems fi : Ai Ñ Bi of dagger affinoid algebras. By Lemma 5.6 the projections AV Ñ Ai and BU Ñ Bi are homotopy epimorphisms. Lemma 2.3 applied to the bottom horizontal map of the commutative diagram A Ai B Bi implies that Ai Ñ Bi is a homotopy epimorphism for every i P N. This is equivalent to say that f can be written as a projective system of homotopy epimorphisms of dagger affinoid algebras. Applying Theorem 5.11 of [4] we obtain that the morphisms Ai Ñ Bi are open immersions of dagger affinoid spaces. Therefore, Mpfq : MpBUq Ñ MpAV q can be written as filtered inductive limit of open embeddings, and it is easy to check that this implies that Mpfq is an open immersion. l To conclude our homological and categorical characterization of the topology of Stein spaces it remains to characterize coverings. Consider a Stein space X and an arbitrary covering ď of X made of Stein spaces Yi. By definition the topology of X is hemi-compact and X is also paracompact. Therefore, the family tYiuiPI admits a countable sub-family tYjujPJ , where J Ă I, such that X “ Yi X “ Yj. iPI ď jPJ š iPJ Lemma 5.8. Let A be a dagger Stein algebra and let tfi : A Ñ AViuiPI be a family of localizations such that for some countable subset J Ă I the corresponding family of functors for is conservative. Then, the morphism φ : tFi : ModRRpAq Ñ ModRRpAViquiPJ š MaxpAViq Ñ MaxpAq is surjective. Proof. Assume that the family of functors tFiuiPJ is conservative and that φ : MaxpAViq Ñ MaxpAq is not surjective. We will deduce a contradiction. Let mx P MaxpAq be a point which is not in the image of φ. Consider the quotient A{mx. This is a Stein algebra: see Example 4.14. By 5.5 A{mx P ModRRpAq and it is a non-trivial module. We also have iPJ pbApA{mxq “ 0 AVi for all i, because the extension of mx to AVi is equal to the improper ideal for all i P J. This l proves that the family Fi is not conservative. F pAq the full sub-category of ModRRpAq for which Definition 5.9. We denote with ModRR M P ModRRpAq is a bornological Fr´echet space. Lemma 5.10. Let tEiuiPI be a countable set of binuclear bornological Fr´echet modules over A and F a bornological Fr´echet modules over A, where A is a nuclear Fr´echet algebra. Then, the canonical map ź EpbAF “ p ź FpbAEi “ iPI ź EiqpbAF Ñ cokerpFpbkApbkEi iPI pEi pbAFq d1Ñ FpbkEiq iPI ź iPI is an isomorphism of bornological modules. Proof. where d1 is induced by the differential in degree 1 of the Bar resolution. Since direct products of bornological spaces preserve cokernels (see Proposition 1.9 of [30] for a proof of this fact for bornological spaces over C and Proposition 1.2.12 of [2] for the same proof worked out in a more general settings) we see that 49 l l to which we can apply Corollary 3.65 (cofiltering the infinite direct product by its finite products) to deduce that cokerp ź ź cokerpFpbkApbkEi ź pFpbkApbkEiq d1Ñ iPI iPI ź ź pFpbkApbkEiq d1Ñ pFpbkEiqq ź ź Eiqq – FpbAE. Eiq d1Ñ Fpbkp iPI iPI iPI iPI d1Ñ FpbkEiq – cokerp pFpbkEiqq – cokerpFpbkApbkp ź pEi ź AF “ p EiqpbL AF Ñ iPI EpbL pbL AFq Corollary 5.11. Under the same hypothesis of Lemma 5.10 we have that for any countable set I, iPI iPI is an isomorphism of bornological modules. with the complex L ‚ we have that ApE, Fq “ EpbkA AF representing it ApE, Fq, using the notation introduced so far. Therefore, for each n P N L n Proof. The same reasoning of Lemma 5.10 can be extended to EpbL ź pbnpbkF “ p pbnpbkF. EiqpbkA ś pbkA pbnpbkFq. By writing a coundable iPI pEi pbnpbkFq pbkA ś When I is finite, using the fact that the completed projective tensor product commutes with finite products, we can deduce that L n product as a cofiltered projective system of finite products, we can use Corollary 3.65 to deduce that ApE, Fq – ź ś ApE, Fq – L n for general countable collections. Now Dď0pModpAqq Ñ Dď0pModpAqq because direct products are exact and so we get a strict quasi-isomorphism of the complexes defines a functor pEi iPI iPI iPI iPI representing EpbL AF and ś pEi iPI pbL AFq. 50 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER Lemma 5.12. Let A be a Stein algebra and tViuiPN a countable collection of Stein do- F pAq Ñ mains that covers X “ MaxpAq. Then, the corresponding family of functors ModRR ModRR F pAq be any morphism such that fi : MpbAAVi Ñ NpbAAVi ź F pAViq is conservative. Proof. Let f : M Ñ N in ModRR are isomorphisms for all i. The Cech-Amitsur complex ź (5.10) 0 Ñ A Ñ AVi1 Ñ i1PN AVi1 i1,i2PN pbAAVi2 Ñ is strictly exact as a consequence of the Theorem B for Stein spaces (cf. Fundamental Theorem in page 124 of [16] for the Archimedean version of Theorem B and Satz 2.4 [23] Ap´q to the Ap´q to a functor on the derived category of Ap´q we are left with Ap´q commutes with the relavant countable products (by Corollary 5.11) because M is Fr´echet and products respect cokernels. Furthermore, as M is RR-quasi-coherent, we have for the non-Archimedean one). Theorem 2.26 permits to apply the functor MpbL complex (5.10) because it permits to extend MpbL unbounded complexes. Therefore, by applying the derived functor MpbL an object strictly quasi-isomorphic to zero. Notice that in this case that MpbL pbA pbAAVinq pbAAVi2q Ñ pbA pbAAVinq – MpbApAVi1 MpbL ApAVi1 ź pMpbAAVi1 pMpbAAVi1q Ñ for each ii, . . . in. Therefore, we can apply Corollary 3.65 and by a small computation, obtain a a strictly exact complex 0 Ñ M Ñ ź i1PN i1,i2PN We can do the same thing for N and this yields that the morphisms fi extend uniquely to l morphisms of the complexes resolving M and N . Therefore f is an isomorphism. Theorem 5.13. A collection of homotopy epimorphisms of Stein algebras tAV Ñ AViuiPI is a covering of MaxpAV q if and only if the family of functors tModRR F pAViquiPJ is conservative, with J Ă I a countable subset. F pAq Ñ ModRR 5.12. Proof. The assertion of the theorem is simply the combination of Lemma 5.8 and Lemma l Corollary 5.14. Consider pCBornk,pbk, kq as a closed symmetric monoidal elementary quasi- We summarize the main results of this section in the following corollary. abelian category. The natural inclusion of categories StnkãÑCommpCBornkq permits to define a countable version of the formal homotopy Zariski topology on Stnk as in Definition 2.5. The coverings of Stein spaces by Stein spaces (in the usual sense for Archimedean base fields, and in the rigid sense for non-Archimedean base fields) corresponds precisely with the families of morphisms tAV Ñ AViuiPI in the category CommpCBornkq for which the family of functors F pAViquiPJ is conservative, and AV Ñ AVi is a homotopy epimorphism tModRR for each i P J with J Ă I some countable subset. F pAq Ñ ModRR Proof. Theorems 5.7 and 5.5 precisely mean that in Stnk a morphism is a homotopy epimorphism if and only if it is an open immersion. Whereas, the claim on the coverings is l obtained in Theorem 5.13. 51 Remark 5.15. If instead, we want to consider finite covers of Stein spaces by Stein spaces (so that J is finite), then the analogue of Corollary 5.14gives a description of the formal F with ModRR, using all homotopy Zariski topology and in this case one can replace ModRR RR-quasicoherent modules instead of just the Frechet ones. Notice also that within the proof of Lemma 5.12 we have shown that given a countable open Stein cover of a Stein space and any M P ModRR F pAq that the complex 0 Ñ M Ñ pMpbAAVi1q Ñ pMpbAAVi1 pbAAVi2q Ñ ź i1PN ź i1,i2PN is strictly exact. The same holds for finite open Stein covers of a Stein space and any M P ModRRpAq. References [1] Artin, M., Mazur B., Etale homotopy, Springer, 1969. [2] Bambozzi, F., On a generalization of affinoid varieties, Ph.D. thesis, University of Padova, 2013, available at http://arxiv.org/pdf/1401.5702.pdf. [3] Bambozzi, F., Closed graph theorems for bornological spaces, 2015, available at http://arxiv.org/pdf/1508.01563.pdf. [4] Bambozzi, F., Ben-Bassat, O., Dagger Geometry as Banach Algebraic Geometry, preprint http://arxiv.org/pdf/1502.01401.pdf. [5] Ben-Bassat, O., Kremnizer, K., Non-Archimedean analytic geometry as relative algebraic geometry, preprint http://arxiv.org/pdf/1312.0338.pdf. [6] Ben-Bassat, O., Kremnizer, K., A Perspective On The Foundations Of Derived Analytic Geometry, preprint. [7] Berkovich, V., Spectral Theory and Analytic Geometry Over Non-Archimedean Fields, American Math- ematical Society, 1990. [8] Berkovich, V., ´Etale cohomology for non-Archimedean analytic spaces, Publications Mathematiques de l’IHES 78.1 (1993): 5-161. [9] Bokstedt, M., and Neeman, A. Homotopy limits in triangulated categories, Compositio Mathematica 86.2 (1993): 209-234. [10] Bosch, S., Guntzer, U., Remmert, R., Non-Archimedean analysis. A systematic approach to rigid ana- lytic geometry, Springer, 1984. [11] Chiarellotto, Duality in Rigid Analysis, LNM 1454, Springer, 1991. [12] Crew, R., Finiteness theorems for the cohomology of an overconvergent isocrystal on a curve, Annales scientifiques de l’cole Normale Suprieure Volume: 31, Issue: 6, page 717-763, 1988. [13] Douady, A., Douady, R., Hubbard, J., Verdier, J-L., Seminaire de geometrie analytique, Asterisque 16, 1974. [14] Engelking, R., General topology. (1989). [15] Gruson, L., Th´eorie de Fredholm p-adique, Bulletin de la S.M.F., tome 94, 1966. [16] Grauert, H., Remmert, R., Theory of Stein Spaces, (reprint of 1979 edition) Springer, 2004. [17] Grosse-Klonne, E. Rigid analytic spaces with overconvergent structure sheaf., J. reine angew. Math 519 (2000): 73-95. [18] Schiffmann, Jacquet, Ferrier, Gruson, Houzel, Seminaire Banach, Lecture Notes in Mathematics 277, Edited by C. Houzel, Springer-Verlag, 1972. [19] Hogbe-Nlend, H., Bornologies and functional analysis: introductory course on the theory of duality topology-bornology and its use in functional analysis. Vol. 26. Elsevier, 1977. [20] Hogbe-Nlend, H., Th´eorie des bornologies et applications. Springer, 1971. [21] Hogbe-Nlend, H., and Moscatelli, V. B., Nuclear and conuclear spaces. Elsevier, 2011. 52 FEDERICO BAMBOZZI, OREN BEN-BASSAT, AND KOBI KREMNIZER [22] Higgs, D. A., and Rowe, K. A., Nuclearity in the category of complete semilattices. Journal of Pure and Applied Algebra 57.1 (1989): 67-78. [23] Kiehl, R., Theorem A und Theorem B in der nichtarchimedischen Funktionentheorie, Inventiones math- ematicae 2.4 (1967): 256-273. [24] Meyer, R., Local and Analytic Cyclic Homology, European Mathematical Society, 2007. [25] Meyer, R., Bornological versus topological analysis in metrizable spaces. Contemporary Mathematics 363 (2004): 249-278. [26] Palamodov, V. P., Homological methods in the theory of locally convex spaces, Russian Mathematical Surveys 26.1 (1971): 1. [27] Perez-Garcia, C., and Schikhof, W. H., Locally convex spaces over non-Archimedean valued fields. Cambridge University Press, 2010. [28] Pirkovskii, A. Yu., On certain homological properties of Stein algebras. Journal of Mathematical Sci- ences, Volume 95, Issue 6 (1999) 2690-2702. [29] Poineau, J., Les espaces de Berkovich sont ang´eliques [30] Prosmans, F., and Schneiders J., A homological study of bornological spaces, Laboratoire analyse, geometrie et applications, Unite mixte de recherche, Institut Galilee, Universite Paris 13, CNRS, 2000. [31] Prosmans, F. Derived categories for functional analysis., Publications of the Research Institute for Mathematical Sciences, 36.1 (2000): 19-83. [32] Prosmans, F. Derived limits in quasi-abelian categories., Bulletin de la Soci´et´e Royale des Sciences de Li´ege [En ligne], Volume 68 - Ann´ee 1999, Num´ero 5 - 6, 335 - 401. [33] Prosmans, F. Derived Projective Limits of Topological Abelian Groups., Journal of Functional Analysis 162, 135-177 (1999) [34] Retakh, V. S., On the subspace of a countable inductive limit, Dokl. Akad. Nauk SSSR 194(1970), 1277-1279. [35] Ramis, J.-P., G. Ruget. R´esidus et dualit´e., Inventiones mathematicae, 26.2 (1974): 89-131. [36] Schneiders, J.-P., Quasi-Abelian Categories and Sheaves, M´emoires de la S.M.F. deuxieme s´erie, tome 76, 1999. [37] Taylor, J. L. Homology and cohomology for topological algebras, Advances in Mathematics 9.2 (1972): 137-182. [38] Terzioglu, T. ”On Schwartz spaces.” Mathematische Annalen 182.3 (1969): 236-242. [39] Toen, Vaqui´e, Under Spec Z, J. K-Theory 3 (2009), no. 3, 437–500. arXiv:math/0509684 [40] Toen, B., Vezzosi, G., Homotopical algebraic geometry I: Topos theory, Adv. Math. 193, no. 2, (2005). [41] Toen, B., Vezzosi, G., Homotopical algebraic geometry II: Geometric stacks and applications, Mem. Amer. Math. Soc. 193, no. 902, (2008). [42] Vogt, D. Topics on projective spectra of (LB)-spaces., Advances in the theory of Frechet spaces. Springer Netherlands, 1989. 11-27. Federico Bambozzi, Fakultat fur Mathematik, Universitat Regensburg, 93040 Regens- burg, Germany E-mail address: [email protected] Oren Ben-Bassat, Mathematical Institute, Radcliffe Observatory Quarter, Woodstock Road, Oxford OX2 6GG, UK E-mail address: [email protected] Kobi Kremnizer, Mathematical Institute, Radcliffe Observatory Quarter, Woodstock Road, Oxford OX2 6GG, UK E-mail address: [email protected]
1708.01254
1
1708
2017-08-02T19:34:25
Common fixed points for $C^{*}$-algebra-valued modular metric spaces via $C_{*}$-class functions with application
[ "math.FA" ]
Based on the concept and properties of $C^{*}$-algebras, the paper introduces a concept of $C_{*}$-class functions. Then by using these functions in $C^{*}$-algebra- valued modular metric spaces of moeini et al. [14], some common fixed point theorems for self-mappings are established. Also, to support of our results an application is provided for existence and uniqueness of solution for a system of integral equations.
math.FA
math
C ∗-algebra-valued modular metric spaces with application 1 Common fixed points for C∗-algebra-valued modular metric spaces via C∗-class functions with application Bahman Moeini∗,a and Arsalan Hojat Ansarib aDepartment of Mathematics, Hidaj Branch, Islamic Azad University, Hidaj, Iran bDepartment of Mathematics, Karaj Branch, Islamic Azad University, Karaj, Iran Abstract Based on the concept and properties of C ∗-algebras, the paper introduces a concept of C∗-class functions. Then by using these functions in C ∗-algebra- valued modular metric spaces of moeini et al. [14], some common fixed point theorems for self-mappings are established. Also, to support of our results an application is provided for existence and uniqueness of solution for a system of integral equations. 1 Introduction As is well known, the Banach contraction mapping principle is a very useful, sim- ple and classical tool in modern analysis, and it has many applications in applied mathematics. In particular, it is an important tool for solving existence problems in many branches of mathematics and physics. In order to generalize this principle, many authors have introduced various types of contraction inequalities (see [3, 6, 11, 17, 19, 20]). In 2014, Ansari [1] introduced the concept of C-class functions which cover a large class of contractive conditions. Afterwards, Ansari et al. [2] defined and used concept of complex C-class functions involving C-class functions in complex valued Gb-metric spaces to obtain some fixed point results. One of the main directions in obtaining possible generalizations of fixed point results is introducing new types of spaces. In 2010 Chistyakov [5] defined the notion of modular on an arbitrary set and develop the theory of metric spaces generated by modular such that called the modular metric spaces. Recently, Mongkolkeha et ∗ Corresponding author E-mail:[email protected] E-mail:[email protected] 2010 Mathematics Subject Classification: 47H10, 46L07, 46A80. Keywords: C∗-class function, C ∗-algebra-valued modular metric space, common fixed point, occasionally weakly compatible, integral equation. 7 1 0 2 g u A 2 ] . A F h t a m [ 1 v 4 5 2 1 0 . 8 0 7 1 : v i X r a 2 Moeini and Hojat Ansari al. [15, 16] have introduced some notions and established some fixed point results in modular metric spaces. In [13], Ma et al. introduced the concept of C∗-algebra-valued metric spaces. The main idea consists in using the set of all positive elements of a unital C∗- algebra instead of the set of real numbers. This line of research was continued in [9, 10, 12, 18, 21], where several other fixed point results were obtained in the framework of C∗-algebra valued metric, as well as (more general) C∗-algebra-valued b-metric spaces. Recently, Moeini et al. [14] introduced the concept of C∗-algebra- valued modular metric spaces which is a generalization of modular metric spaces and next proved some fixed point theorems for self-mappings with contractive conditions on such spaces. In this paper, we introduce a concept of C∗-class functions on a set of unital C∗-algebra and via these functions some common fixed point results are proved for self-mappings with contractive conditions in C∗-algebra-valued modular metric spaces. Also, some examples to elaborate and illustrate of our results are con- structed. Finally, as application, existence and uniqueness of solution for a type of system of nonlinear integral equations is discussed. 2 Basic notions Let X be a non empty set, λ ∈ (0,∞) and due to the disparity of the arguments, function ω : (0,∞) × X × X → [0,∞] will be written as ωλ(x, y) = ω(λ, x, y) for all λ > 0 and x, y ∈ X. [4] Let X be a non empty set. a function ω : (0,∞) × X × X → Definition 2.1. [0,∞] is said to be a modular metric on X if it satisfies the following three axioms: (i) given x, y ∈ X, ωλ(x, y) = 0 for all λ > 0 if and only if x = y; (ii) ωλ(x, y) = ωλ(y, x) for all λ > 0 and x, y ∈ X; (iii) ωλ+µ(x, y) ≤ ωλ(x, z) + ωµ(z, y) for all λ > 0 and x, y, z ∈ X, and (X, ω) is called a modular metric space. Recall that a Banach algebra A (over the field C of complex numbers) is said to be a C∗-algebra if there is an involution ∗ in A (i.e., a mapping ∗ : A → A satisfying a∗∗ = a for each a ∈ A) such that, for all a, b ∈ A and λ, µ ∈ C, the following holds: (i) (λa + µb)∗ = ¯λa∗ + ¯µb∗; (ii) (ab)∗ = b∗a∗; (iii) ka∗ak = kak2. C ∗-algebra-valued modular metric spaces with application 3 Note that, form (iii), it easy follows that kak = ka∗k for each a ∈ A. Moreover, the pair (A,∗) is called a unital ∗-algebra if A contains the identity element 1A. A positive element of A is an element a ∈ A such that a∗ = a and its spectrum σ(a) ⊂ R+, where σ(a) = {λ ∈ R : λ1A − a is noninvertible}. The set of all positive elements will be denoted by A+. Such elements allow us to define a partial ordering '(cid:23)' on the elements of A. That is, b (cid:23) a if and only if b − a ∈ A+. If a ∈ A is positive, then we write a (cid:23) θ, where θ is the zero element of A. Each positive element a of a C∗-algebra A has a unique positive square root. From now on, by A we mean a unital C∗-algebra with identity element 1A. Further, A+ = {a ∈ A : a (cid:23) θ} and (a∗a) 2 = a. 1 Lemma 2.2. [7] Suppose that A is a unital C∗-algebra with a unit 1A. (1) For any x ∈ A+, we have x (cid:22) 1A ⇔ kxk ≤ 1. (2) If a ∈ A+ with kak < 1 (3) Suppose that a, b ∈ A with a, b (cid:23) θ and ab = ba, then ab (cid:23) θ. (4) By A′ we denote the set {a ∈ A : ab = ba,∀b ∈ A}. Let a ∈ A′ if b, c ∈ A with 2 , then 1A − a is invertible and ka(1A − a)−1k < 1. b (cid:23) c (cid:23) θ, and 1A − a ∈ A′ is an invertible operator, then (1A − a)−1b (cid:23) (1A − a)−1c. Notice that in a C∗-algebra, if θ (cid:22) a, b, one cannot conclude that θ (cid:22) ab. For −2 4 (cid:19) , example, consider the C∗-algebra M2(C) and set a =(cid:18) 3 then ab =(cid:18) −1 2 −4 8 (cid:19). Clearly a, b ∈ M2(C)+, while ab is not. 3 (cid:19), b =(cid:18) 1 − 2 2 2 [14] Let X be a non empty set. a function ω : (0,∞)×X×X → A+ Definition 2.3. is said to be a C∗-algebra-valued modular metric (briefly, C∗.m.m) on X if it satisfies the following three axioms: (i) given x, y ∈ X, ωλ(x, y) = θ for all λ > 0 if and only if x = y; (ii) ωλ(x, y) = ωλ(y, x) for all λ > 0 and x, y ∈ X; (iii) ωλ+µ(x, y) (cid:22) ωλ(x, z) + ωµ(z, y) for all λ, µ > 0 and x, y, z ∈ X. The truple (X, A, ω) is called a C∗.m.m space. 4 Moeini and Hojat Ansari If instead of (i), we have the condition (i′) ωλ(x, x) = θ for all λ > 0 and x ∈ X, then ω is said to be a C∗-algebra-valued pseudo modular metric (briefly, C∗.p.m.m) on X and if ω satisfies (i′), (iii) and (i′′) given x, y ∈ X, if there exists a number λ > 0, possibly depending on x and y, such that ωλ(x, y) = θ, then x = y, then ω is called a C∗-algebra-valued strict modular metric (briefly, C∗.s.m.m) on X. A C∗.m.m (or C∗.p.m.m, C∗.s.m.m) ω on X is said to be convex if, instead of λ+µ ωλ(x, z) + µ (iii), we replace the following condition: (iv) ωλ+µ(x, y) (cid:22) λ Clearly, if ω is a C∗.s.m.m, then ω is a C∗.m.m, which in turn implies ω is a C∗.p.m.m on X, and similar implications hold for convex ω. The essential property of a C∗.m.m ω on a set X is a following given x, y ∈ X, the function 0 < λ → ωλ(x, y) ∈ A is non increasing on (0,∞). In fact, if 0 < µ < λ, then we have λ+µ ωµ(z, y) for all λ, µ > 0 and x, y, z ∈ X. ωλ(x, y) (cid:22) ωλ−µ(x, x) + ωµ(x, y) = ωµ(x, y). (2.1) It follows that at each point λ > 0 the right limit ωλ+0(x, y) := limε→+0 ωλ+ε(x, y) and the left limit ωλ−0(x, y) := limε→+0 ωλ−ε(x, y) exist in A and the following two inequalities hold: ωλ+0(x, y) (cid:22) ωλ(x, y) (cid:22) ωλ−0(x, y). (2.2) It can be check that if x0 ∈ X, the set Xω = {x ∈ X : lim λ→∞ ωλ(x, x0) = θ}, is a C∗-algebra-valued metric space, called a C∗-algebra-valued modular space, whose d0 ω : Xω × Xω → A is given by d0 ω = inf{λ > 0 : kωλ(x, y)k ≤ λ} for all x, y ∈ Xω. Moreover, if ω is convex, the set Xω is equal to X∗ω = {x ∈ X : ∃ λ = λ(x) > 0 such that kωλ(x, x0)k < ∞}, and d∗ω : X∗ω × X∗ω → A is given by d∗ω = inf{λ > 0 : kωλ(x, y)k ≤ 1} for all x, y ∈ X∗ω. It is easy to see that if X is a real linear space, ρ : X → A and ωλ(x, y) = ρ( x − y λ ) for all λ > 0 and x, y ∈ X, (2.3) then ρ is C∗-algebra valued modular (convex C∗-algebra-valued modular) on X if and only if ω is C∗.m.m (convex C∗.m.m, respectively) on X. On the other hand, if ω satisfy the following two conditions: C ∗-algebra-valued modular metric spaces with application 5 (i) ωλ(µx, 0) = ω λ µ (x, 0) for all λ, µ > 0 and x ∈ X; (ii) ωλ(x + z, y + z) = ωλ(x, y) for all λ > 0 and x, y, z ∈ X. If we set ρ(x) = ω1(x, 0) with (2.3) holds, where x ∈ X, then (a) Xρ = Xω is a linear subspace of X and the functional kxkρ = d0 is a F -norm on Xρ; ω(x, 0), x ∈ Xρ (b) If ω is convex, X∗ρ ≡ X∗ω = Xρ is a linear subspace of X and the functional kxkρ = d∗ω(x, 0), x ∈ X∗ρ is a norm on X∗ρ . Similar assertions hold if replace C∗.m.m by C∗.p.m.m. If ω is C∗.m.m in X, we called the set Xω is C∗.m.m space. By the idea of property in C∗-algebra-valued metric spaces and C∗-algebra- valued modular spaces, we defined the following: Definition 2.4. [14] Let Xω be a C∗.m.m space. (1) The sequence (xn)n∈N in Xω is said to be ω-convergent to x ∈ Xω with respect to A if ωλ(xn, x) → θ as n → ∞ for all λ > 0. (2) The sequence (xn)n∈N in Xω is said to be ω-Cauchy with respect to A if ωλ(xm, xn) → θ as m, n → ∞ for all λ > 0. (3) A subset C of Xω is said to be ω-closed with respect to A if the limit of the ω-convergent sequence of C always belong to C. (4) Xω is said to be ω-complete if any ω-Cauchy sequence with respect to A is ω-convergent. (5) A subset C of Xω is said to be ω-bounded with respect to A if for all λ > 0 δω(C) = sup{kωλ(x, y)k; x, y ∈ C} < ∞. [14] Let Xω be a C∗.m.m space. Let f, g self-mappings of Xω. a Definition 2.5. point x in Xω is called a coincidence point of f and g iff f x = gx. We shall call w = f x = gx a point of coincidence of f and g. Definition 2.6. said to be weakly compatible if they commute at coincidence points. [14] Let Xω be a C∗.m.m space. Two maps f and g of Xω are 6 Moeini and Hojat Ansari [14] Let Xω be a C∗.m.m space. Two self-mappings f and g of Definition 2.7. Xω are occasionally weakly compatible (owc) iff there is a point x in Xω which is a coincidence point of f and g at which f and g commute. [8] Let Xω be a C∗.m.m space and f, g owc self-mappings of Xω. If f Lemma 2.8. and g have a unique point of coincidence, w = f x = gx, then w is a unique common fixed point of f and g. In 2017, Ansari et al. [2] introduced the concept of complex C-class functions as follows: Definition 2.9. Suppose S = {z ∈ C : z (cid:23) 0}, then a continuous function F : S2 → C is called a complex C-class function if for any s, t ∈ S, the following conditions hold: (1) F (s, t) (cid:22) s; (2) F (s, t) = s implies that either s = 0 or t = 0. An extra condition on F that F (0, 0) = 0 could be imposed in some cases if required. For examples of these functions see [2]. 3 Main results In this section, we introduce a C∗-class function. The main idea consists in using the set of elements of a unital C∗-algebra instead of the set of complex numbers. Definition 3.1. (C∗-class function) Suppose A is a unital C∗-algebra, then a con- tinuous function F : A+ × A+ → A is called C∗-class function if for any A, B ∈ A+, the following conditions hold: (1) F (A, B) (cid:22) A; (2) F (A, B) = A implies that either A = θ or B = θ. An extra condition on F that F (θ, θ) = θ could be imposed in some cases if required. The letter C∗ will denote the class of all C∗-class functions. Remark 3.2. The class C∗ includes the set of complex C-class functions. suffitient to take A = C in Definition 3.1. It is The following examples show that the class C∗ is nonempty: Example 3.3. Let A = M2(R), of all 2 × 2 matrices with the usual operation of addition, scalar multiplication, and matrix multiplication. Define norm on A by kAk = (cid:16)P2 , and ∗ : A → A, given by A∗ = A, for all A ∈ A, defines a convolution on A. Thus A becomes a C∗-algebra. For 1 2 i,j=1 aij2(cid:17) A =(cid:18) a11 a21 a12 a22 (cid:19) , B =(cid:18) b11 b21 b12 b22 (cid:19) ∈ A = M2(R), C ∗-algebra-valued modular metric spaces with application 7 we denote A (cid:22) B if and only if (aij − bij) ≤ 0, for all i, j = 1, 2. (1) Define F∗ : A+ × A+ → A by a22 (cid:19) ,(cid:18) b11 b22 (cid:19) (cid:17) =(cid:18) a11 − b11 a21 − b21 F∗(cid:16)(cid:18) a11 a21 a12 a22 − b22 (cid:19) a12 − b12 b12 b21 for all ai,j, bi,j ∈ R+, (i, j ∈ {1, 2}). Then F∗ is a C∗-class function. (2) Define F∗ : A+ × A+ → A by F∗(cid:16)(cid:18) a11 a21 a12 a22 (cid:19) ,(cid:18) b11 b21 b12 b22 (cid:19) (cid:17) = m(cid:18) a11 a21 a12 a22 (cid:19) for all ai,j, bi,j ∈ R+, (i, j ∈ {1, 2}), where, m ∈ (0, 1). Then F∗ is a C∗-class function. Example 3.4. Let X = L∞(E) and H = L2(E), where E is a lebesgue measurable set. By B(H) we denote the set of bounded linear operator on Hilbert space H. Clearly, B(H) is a C∗-algebra with the usual operator norm. Define F∗ : B(H)+ × B(H)+ → B(H) by F∗(U, V ) = U − ϕ(U ), where ϕ : B(H)+ → B(H)+ is is a continuous function such that ϕ(U ) = θ if and only if U = θ (θ = 0B(H)). Then F∗ is a C∗-class function. Let Φu denote the class of the functions ϕ : A+ → A+ which satisfy the following conditions: (ϕ1) ϕ is continuous and non-decreasing; (ϕ2) ϕ(T ) ≻ θ, T ≻ θ and ϕ(θ) (cid:23) θ. Let Ψ be a set of all continuous functions ψ : A+ → A+ satisfying the following conditions: (ψ1) ψ is continuous and non-decreasing; (ψ2) ψ(T ) = θ if and only if T = θ. Definition 3.5. A tripled (ψ, ϕ, F∗) where ψ ∈ Ψ, ϕ ∈ Φu and F∗ ∈ C∗ is said to be monotone if for any A, B ∈ A+ A (cid:22) B =⇒ F∗(ψ(A), ϕ(A)) (cid:22) F∗(ψ(B), ϕ(B)). We now give detailed proofs of main results of this paper. 8 Moeini and Hojat Ansari Theorem 3.6. Let Xω be a C∗.m.m space and I, J, R, S, T, U : Xω → Xω be self- mappings of Xω such that the pairs (SR, I) and (T U, J) are occasionally weakly compatible. Suppose there exist a, b, c ∈ A with 0 < kak2 +kbk2 +kck2 ≤ 1 such that the following assertion for all x, y ∈ Xω and λ > 0 hold: (3.1.1) ψ(ωλ(SRx, T U y)) (cid:22) F∗(cid:16)ψ(M (x, y)), ϕ(M (x, y)(cid:17), where M (x, y) = a∗ωλ(Ix, J y)a + b∗ωλ(SRx, J y)b + c∗ω2λ(T U y, Ix)c, ψ ∈ Ψ, ϕ ∈ Φu and F∗ ∈ C∗ such that (ψ, ϕ, F∗) is monotone; (3.1.2) kωλ(SRx, T U y)k < ∞. Then SR, T U, I and J have a common fixed point in Xω. Furthermore if the pairs (S, R), (S, I), (R, I), (T, J), (T, U ), (U, J ) are commuting pairs of mappings then I, J, R, S, T and U have a unique common fixed point in Xω. Proof. Since the pair (SR, I) and (T U, J) are occasionally weakly compatible then there exists u, v ∈ Xω : SRu = Iu and T U v = J v. Moreover; SR(Iu) = I(SRu) and T U (J v) = J(T U v). Now we can assert that SRu = T U v. By (3.1.1), we have ψ(ωλ(SRu, T U v)) (cid:22) F∗(cid:16)ψ(M (u, v)), ϕ(M (u, v)(cid:17), (3.4) where M (u, v) = a∗ωλ(Iu, J v)a + b∗ωλ(SRu, J v)b + c∗ω2λ(T U v, Iu)c = a∗ωλ(Iu, J v)a + b∗ωλ(Iu, J v)b + c∗ω2λ(J v, Iu)c. (3.5) C ∗-algebra-valued modular metric spaces with application 9 By definition of C∗.m.m space and inequalites (2.1), (3.4) and (3.5), we get ψ(ωλ(Iu, J v)) = ψ(ωλ(SRu, T U v)) (cid:22) F∗(cid:16)ψ(a∗ωλ(Iu, J v)a + b∗ωλ(Iu, J v)b + c∗(ωλ(Iu, Iu) + ωλ(Iu, J v))c), ϕ(a∗ωλ(Iu, J v)a + b∗ωλ(Iu, J v)b + c∗(ωλ(Iu, Iu) + ωλ(Iu, J v))c)(cid:17) = F∗(cid:16)ψ(a∗ωλ(Iu, J v)a + b∗ωλ(Iu, J v)b + c∗ωλ(Iu, J v)c), ϕ(a∗ωλ(Iu, J v)a + b∗ωλ(Iu, J v)b + c∗ωλ(Iu, J v)c)(cid:17) = F∗(cid:16)ψ(a∗(ωλ(Iu, J v)) 2 a + b∗(ωλ(Iu, J v)) 2 (ωλ(Iu, J v)) 2 (ωλ(Iu, J v)) 1 1 1 1 1 1 2 b 2 c), 2 a + b∗(ωλ(Iu, J v)) 1 1 2 (ωλ(Iu, J v)) 1 2 b +c∗(ωλ(Iu, J v)) ϕ(a∗(ωλ(Iu, J v)) +c∗(ωλ(Iu, J v)) 1 2 (ωλ(Iu, J v)) 2 (ωλ(Iu, J v)) 2 (ωλ(Iu, J v)) 1 1 2 c)(cid:17) 1 = F∗(cid:16)ψ((a(ωλ(Iu, J v)) 1 +(c(ωλ(Iu, J v)) ϕ((a(ωλ(Iu, J v)) +(c(ωλ(Iu, J v)) 1 2 )∗(c(ωλ(Iu, J v)) 2 )∗(a(ωλ(Iu, J v)) 2 )∗(c(ωλ(Iu, J v)) 1 1 1 2 ))(cid:17) (cid:22) F∗(cid:16)ψ(ka(ωλ(Iu, J v)) 2k21A + kb(ωλ(Iu, J v)) 2k21A + kb(ωλ(Iu, J v)) ϕ(ka(ωλ(Iu, J v)) = F∗(cid:16)ψ(kωλ(Iu, J v)k(kak2 + kbk2 + kck2)1A), ϕ(kωλ(Iu, J v)k(kak2 + kbk2 + kck2)1A)(cid:17). 1 1 1 2 )∗(a(ωλ(Iu, J v)) 1 2 ) + (b(ωλ(Iu, J v)) 1 2 )∗(b(ωλ(Iu, J v)) 1 2 ) 2 )), 2 ) + (b(ωλ(Iu, J v)) 1 1 2 )∗(b(ωλ(Iu, J v)) 1 2 ) 1 2k21A + kc(ωλ(Iu, J v)) 2k21A + kc(ωλ(Iu, J v)) 2k21A)(cid:17) 1 1 2k21A), (3.6) So, ψ(kωλ(Iu, J v)k1A) ≤ F∗(cid:16)ψ(kωλ(Iu, J v)k(kak2 + kbk2 + kck2)1A), ϕ(kωλ(Iu, J v)k(kak2 + kbk2 + kck2)1A)(cid:17) (cid:22) F∗(cid:16)ψ(kωλ(Iu, J v)k1A), ϕ(kωλ(Iu, J v)k1A)(cid:17). (3.7) Thus, ψ(kωλ(Iu, J v)k1A) = θ or ϕ(kωλ(Iu, J v)k1A) = θ, which means Iu = J v. Hence SRu = T U v and thus SRu = Iu = T U v = J v. (3.8) 10 Moeini and Hojat Ansari Moreover, if there is another point z such that SRz = Iz, and using condition (3.1.1) ψ(ωλ(SRz, T U v)) (cid:22) F∗(cid:16)ψ(a∗ωλ(Iz, J v)a + b∗ωλ(SRz, J v)b + c∗ω2λ(T U v, Iz)c), ϕ(a∗ωλ(Iz, J v)a + b∗ωλ(SRz, J v)b + c∗ω2λ(T U v, Iz)c)(cid:17) = F∗(cid:16)ψ(a∗ωλ(SRz, T U v)a + b∗ωλ(SRz, T U v)b + c∗ω2λ(SRz, T U v)c), ϕ(a∗ωλ(SRz, T U v)a + b∗ωλ(SRz, T U v)b + c∗ω2λ(SRz, T U v)c)(cid:17). (3.9) By above similar way, we conclude that ψ(kωλ(SRz, T U v)k1A) (cid:22) F∗(cid:16)ψ(kωλ(SRz, T U v)k(kak2 + kbk2 + kck2))1A, ϕ(kωλ(SRz, T U v)k(kak2 + kbk2 + kck2)1A)(cid:17) (cid:22) F∗(cid:16)ψ(kωλ(SRz, T U v)k1A, ϕ(kωλ(SRz, T U v)k1A)(cid:17). Therefore, ψ(kωλ(SRz, T U v)k1A) = θ or ϕ(kωλ(SRz, T U v)k1A) = θ, which means SRz = T U v, and so, (3.10) SRu = Iu = T U v = J v. Thus from equation (3.9) and (3.10) it follows that SRu = SRz. Hence, w = SRu = Iu for some w ∈ Xω is the unique point of coincidence of SR and I. Then by Lemma 2.8, w is a unique common fixed point of SR and I. So, SRw = Iw = w. Similarly, there is another common fixed point w′ ∈ Xω : T U w′ = J w′ = w′. For the uniqueness, by (3.1.1) we have ψ(ωλ(SRw, T U w′)) = ψ(ωλ(w, w′)) (cid:22) F∗(cid:16)ψ(a∗ωλ(Iw, J w′)a + b∗ωλ(SRw, J w′)b + c∗ω2λ(T U w, Iw′)c), ϕ(a∗ωλ(Iw, J w′)a + b∗ωλ(SRw, J w′)b + c∗ω2λ(T U w, Iw′)c)(cid:17) = F∗(cid:16)ψ(a∗ωλ(w, w′)a + b∗ωλ(w, w′)b + c∗ω2λ(w, w′)c, ϕ(a∗ωλ(w, w′)a + b∗ωλ(w, w′)b + c∗ω2λ(w, w′)c)(cid:17). (3.11) Thus, ψ(kωλ(w, w′)k1A) (cid:22) F∗(cid:16)ψ(kωλ(w, w′)k(kak2 + kbk2 + kck2)1A), ϕ(kωλ(w, w′)k(kak2 + kbk2 + kck2)1A)(cid:17) (cid:22) F∗(cid:16)ψ(kωλ(w, w′)k1A), ϕ(kωλ(w, w′)k1A)(cid:17). C ∗-algebra-valued modular metric spaces with application 11 So, ψ(kωλ(w, w′)k1A) = θ or ϕ(kωλ(w, w′)k1A) = θ. Hence w = w′. Therefore, w is a unique common fixed point of SR, T U, I and J. Furthermore, if we take pairs (S, R), (S, I), (R, I), (T, J), (T, U ), (U, J) are com- muting pairs then Sw = S(SRw) = S(RS)w = SR(Sw) Sw = S(Iw) = S(RS)w = I(Sw) Rw = R(SRw) = RS(Rw) = SR(Rw) Rw = R(Iw) = (Rw), this shows that Sw and Rw is common fixed point of (SR, I) and this gives SRw = Sw = Rw = Iw = w. Similarly, we have T U w = T w = U w = J w = w. Hence, w is a unique common fixed point of S, R, I, J, T, U . Corollary 3.7. Let Xω be a C∗.m.m space and I, J, S, T : Xω → Xω be self- mappings of Xω such that the pairs (S, I) and (T, J) are occasionally weakly com- patible. Suppose there exist a, b, c ∈ A with 0 < kak2 + kbk2 + kck2 ≤ 1 such that the following assertion for all x, y ∈ Xω and λ > 0 hold: (3.2.1) ψ(ωλ(Sx, T y)) (cid:22) F∗(cid:16)ψ(N (x, y)), ϕ(N (x, y))(cid:17) where, N (x, y) = a∗ωλ(Ix, J y)a + b∗ωλ(Sx, J y)b + c∗ω2λ(T y, Ix)c ψ ∈ Ψ, ϕ ∈ Φu and F∗ ∈ C∗ such that (ψ, ϕ, F∗) is monotone; (3.2.2) kωλ(Sx, T y)k < ∞. Then S, T, I and J have a unique common fixed point in Xω. Proof. If we put R = U := Ixω where Ixω is an identity mapping on Xω, the result follows from Theorem 3.6. Corollary 3.8. Let Xω be a C∗.m.m space and S, T : Xω → Xω be self-mappings of Xω such that S and T are occasionally weakly compatible. Suppose there exist a, b, c ∈ A with 0 < kak2 + kbk2 + kck2 ≤ 1 such that the following assertion for all x, y ∈ Xω and λ > 0 hold: (3.3.1) ψ(ωλ(T x, T y)) (cid:22) F∗(cid:16)ψ(O(x, y)), ϕ(O(x, y))(cid:17) where, O(x, y) = a∗ωλ(Sx, Sy)a + b∗ωλ(T x, Sy)b + c∗ω2λ(T y, Sx)c ψ ∈ Ψ, ϕ ∈ Φu and F∗ ∈ C∗ such that (ψ, ϕ, F∗) is monotone; (3.3.2) kωλ(T x, T y)k < ∞. Then S and T have a unique common fixed point in Xω. 12 Moeini and Hojat Ansari Proof. If we put I = J := S, and S := T in (3.2.1) and (3.2.2) the result follows from Theorem 3.6. Corollary 3.9. Let Xω be a C∗.m.m space and S, T : Xω → Xω be self-mappings of Xω such that S and T are occasionally weakly compatible. Suppose there exist a ∈ A with 0 < kak ≤ 1 such that the following assertion for all x, y ∈ Xω and λ > 0 hold: (3.4.1) ψ(ωλ(T x, T y)) (cid:22) F∗(cid:16)ψ(a∗ωλ(Sx, Sy)a), ϕ(a∗ωλ(Sx, Sy)a)(cid:17), where, ψ ∈ Ψ, ϕ ∈ (3.4.2) kωλ(T x, T y)k < ∞. Then S and T have a unique common fixed point in Xω. Φu and F∗ ∈ C∗ such that (ψ, ϕ, F∗) is monotone; Proof. If we put b = c := θ, in (3.3.1) the result follows from Corollary 3.8. 4 Examples In this section we furnish some nontrivial examples in favour of our results. Example 4.1. Let X = R and consider, A = M2(R) as in Example 3.3. Define ω : (0,∞) × X × X → A+ by ωλ(x, y) =(cid:18) x−y λ 0 λ (cid:19) , 0 x−y for all x, y ∈ X and λ > 0. It is easy to check that ω satisfies all the conditions of Definition 2.3. So, (X, A, ω) is a C∗.m.m space. Example 4.2. Let X = { 1 Define ω : (0,∞) × X × X → A+ by cn : n = 1, 2,··· } where 0 < c < 1 and A = M2(R). ωλ(x, y) =(cid:18) k x−y λ k 0 λ k (cid:19) , 0 αk x−y for all x, y ∈ X, α ≥ 0 and λ > 0. Then it is easy to check that ω is a C∗.m.m. Example 4.3. Let X = L∞(E) and H = L2(E), where E is a Lebesgue measurable set. By B(H) we denote the set of bounded linear operator on Hilbert space H. Clearly, B(H) is a C∗-algebra with the usual operator norm. Define ω : (0,∞) × X × X → B(H)+ by f −g λ (∀f, g ∈ X), where πh : H → H is the multiplication operator defined by ωλ(f, g) = π , πh(φ) = h.φ, C ∗-algebra-valued modular metric spaces with application 13 for φ ∈ H. Then ω is a C∗.m.m and (Xω, B(H), ω) is a ω-complete C∗.m.m space. It suffices to verify the completeness of Xω. For this, let {fn} be a ω-Cauchy sequence with respect to B(H), that is for an arbitrary ε > 0, there is N ∈ N such that for all m, n ≥ N , kωλ(fm, fn)k = kπ fm−fn λ k = k fm−fn λ k∞ ≤ ε, so {fn} is a Cauchy sequence in Banach space X. Hence, there is a function f ∈ X and N1 ∈ N such that It implies that k fn−f λ k∞ ≤ ε, (n ≥ N1). kωλ(fn, f )k = kπ λ k = k fn−f fn−f λ k∞ ≤ ε, (n ≥ N1). Consequently, the sequence {fn} is a ω-convergent sequence in Xω and so Xω is a ω-complete C∗.m.m space. Example 4.4. Let (X, A, ω) is C∗.m.m space defined as in Example 4.1. Define S, T, I, J : Xω → Xω by 2x 3 2 0 if x ∈ (−∞, 2), if x = 2, if x ∈ (2,∞). Suppose, Sx = T x = 2, J x = 4 − x, Ix =  (cid:26) ϕ : A+ → A+ (cid:26) ψ : A+ → A+ ψ(A) = 2A, (cid:26) F∗ : A+ × A+ → A F∗(A, B) = A − B. Then, (ψ, ϕ, F∗) is monotone. For all x, y ∈ Xω = R and λ > 0, we have ϕ(A) = A, For every a, b, c ∈ A with 0 < kak2 + kbk2 + kck2 ≤ 1, we get 0 0 0 (cid:19)(cid:13)(cid:13)(cid:13) = kωλ(Sx, T y)k < ∞. 0 =(cid:13)(cid:13)(cid:13)(cid:18) 0 0 (cid:19) = ψ(ωλ(Sx, T y)) (cid:22) F∗(cid:16)ψ(M (x, y)), ϕ(M (x, y))(cid:17), 0 (cid:18) 0 0 for all x, y ∈ Xω and λ > 0. Also clearly, the pairs (S, I) and (T, J) are occasionally weakly compatible. So all the conditions of the Corollary 3.7 are satisfied and x = 2 is a unique common fixed point of S, T, I and J. 14 5 Application Moeini and Hojat Ansari Remind that if for λ > 0 and x, y ∈ L∞(E), define ω : (0,∞) × L∞(E) × L∞(E) → B(H)+ by ωλ(x, y) = π , x−y λ where, πh : H → H be defined as in Example 4.3, then (L∞(E)ω, B(H), ω) is a ω-complete C∗.m.m space. Let E be a Lebesgue measurable set, X = L∞(E) and H = L2(E) be the Hilbert space. Consider the following system of nonlinear integral equations: x(t) = w(t) + ki(t, x(t)) + µZE n(t, s)hj(s, x(s))ds, (5.12) for all t ∈ E, where w ∈ L∞(E)ω is known, ki(t, x(t)), n(t, s), hj(s, x(s)), i, j = 1, 2 and i 6= j are real or complex valued functions that are measurable both in t and s on E and µ is real or complex number, and assume the following conditions: (a) sups∈ERE n(t, s)dt = M1 < +∞, (b) ki(s, x(s)) ∈ L∞(E)ω for all x ∈ L∞(E)ω, and there exists L1 > 1 such that for all s ∈ E, k1(s, x(s)) − k2(s, y(s)) √2 ≥ L1x(s) − y(s) for all x, y ∈ L∞(E)ω, (c) hi(s, x(s)) ∈ L∞(E)ω for all x ∈ L∞(E)ω, and there exists L2 > 0 such that for all s ∈ E, h1(s, x(s)) − h2(s, y(s)) ≤ L2x(s) − y(s) for all x, y ∈ L∞(E)ω, (d) there exists x(t) ∈ L∞(E)ω such that implies (e) there exists y(t) ∈ L∞(E)ω such that x(t) − w(t) − µRE n(t, s)h1(s, x(s))ds = k1(t, x(t)), k1(t, x(t)) − w(t) − µRE n(t, s)h1(s, k1(s, x(s)))ds = k1(t, x(t) − w(t) − µRE n(t, s)h1(s, x(s))ds). y(t) − w(t) − µRE n(t, s)h2(s, y(s))ds = k2(t, y(t)), C ∗-algebra-valued modular metric spaces with application 15 implies k2(t, y(t)) − w(t) − µRE n(t, s)hi(s, k2(s, y(s)))ds = k2(t, y(t) − w(t) − µRE n(t, s)h2(s, y(s))ds). Theorem 5.1. With the assumptions (a)-(e), the system of nonlinear integral equa- tions (5.12) has a unique solution x∗ in L∞(E)ω for each real or complex number µ with 1+µL2M1 L1 ≤ 1. Proof. Define Sx(t) = x(t) − w(t) − µZE n(t, s)h1(s, x(s))ds, T x(t) = x(t) − w(t) − µZE n(t, s)h2(s, x(s))ds, Ix(t) = k1(t, x(t)), J x(t) = k2(t, x(t)). Set a = q 1+µM1L2 kbk2 + kck2 = 1+µM1L2 L1 Define ≤ 1. L1 .1B(H), b = c = θ = 0B(H), then a ∈ B(H)+ and 0 < kak2 + (cid:26) ψ : B(H)+ → B(H)+ ψ(B) = 1 2 B, (cid:26) ϕ : B(H)+ → B(H)+ ϕ(B) = 1 4 B, ( F∗ : B(H)+ × B(H)+ → B(H) F∗(A, B) = 1√2 A. Then, (ψ, ϕ, F∗) is monotone. 16 Moeini and Hojat Ansari h, h) Sx−T y λ 2 supkhk=1(π For any h ∈ H, we have kψ(ωλ(Sx, T y))k = 1 2λ(cid:12)(cid:12)(cid:12) = supkhk=1REh 1 2λ(cid:12)(cid:12)(cid:12) ≤ supkhk=1REh 1 2λ supkhk=1RE h(t)2dthkx − yk∞ + µM1L2kx − yk∞i ≤ 1 ≤ ( 1+µM1L2 ≤ 1√2 (x − y) + µRE n(t, s)(h2(s, y(s) − h1(s, x(s))ds(cid:12)(cid:12)(cid:12)ih(t)h(t)dt (x − y) + µRE n(t, s)(h2(s, y(s) − h1(s, x(s))ds(cid:12)(cid:12)(cid:12)ih(t)2dt )k k1(t,x(t))−k2(t,y(t)) )kx − yk∞ ( 1+µM1L2 2L1 λ k∞ 2λ = 1√2 2 ( 1+µM1L2 . 1 L1 )kωλ(Ix, J y)k = 1√2 . 1 2kak2kωλ(Ix, J y)k 2(cid:16)kak2kωλ(Ix, J y)k + kbk2kωλ(Sx, J y)k + kck2kω2λ(T y, Ix)k(cid:17) . 1 = 1√2 = kF∗(cid:16)ψ(N (x, y)), ϕ(N (x, y))(cid:17)k Then, kψ(ωλ(Sx, T y))k ≤ kF∗(cid:16)ψ(N (x, y)), ϕ(N (x, y))(cid:17)k, for all x, y ∈ L∞(E)ω and λ > 0. Also by conditions (d) and (e) the pairs (S, I) and (T, J) are occasionally weakly compatible. Therefore, by the Corollary 3.7, there exists a unique common fixed point x∗ ∈ L∞(E)ω such that x∗ = Sx∗ = T x∗ = Ix∗ = J x∗, which proves the existence of unique solution of (5.12) in L∞(E)ω. This completes the proof. References [1] A.H. Ansari, Note on " ϕ-ψ -contractive type mappings and related fixed point", The 2nd Regional Conference on Mathematics And Applications, Payame Noor University, 2014, pages 377-380. [2] A.H. Ansari, O. Ege and S. Randenovi´c, Some fixed point results on complex valued Gb-metric spaces, RACSAM (2017). doi:10.1007/s13398-017-0391-x. C ∗-algebra-valued modular metric spaces with application 17 [3] A. Branciari, A fixed point theorem for mappings satisfying a general contrac- tive condition of integral type, Hindawi Publishing Corpration, Inter. J. Math. Math. Sci., 29 (2002), 531-536. [4] V.V. Chistyakov, Modular metric spaces generated by F -modulars, Folia Math., 14 (2008), 3-25. [5] V.V. Chistyakov, Modular metric spaces I basic concepts, Nonlinear Anal., 72 (2010), 1-14. [6] S. Dhompongsa, H. Yingtaweesittikul, Fixed point for multivalued mappings and the metric completeness, Fixed Point Theory and Applications, 2009, 15 pages, Article ID 972395. [7] R. Douglas, Banach Algebra Techniques in Operator Theory, Springer, Berlin (1998). [8] G. Jungck and B.E. Rhoades Fixed point theorems for occasionally weakly compatible mappings, Fixed Point Theory, 7(2) (2006), 287-296. [9] Z. Kadelburg and S. Radenovi´c, Fixed point results in C∗-algebra-valued metric spaces are direct consequences of their standard metric counterparts, Fixed Point Theory and Appl., 2016, Article ID 53 (2016). [10] T. Kamran, M. Postolache, A. Ghiura, S. Batul and R. Ali, The Banach con- traction principle in C∗-algebra-valued b-metric spaces with application, Fixed Point Theory and Appl., 2016, Article ID 10 (2016). [11] M. Kikkawa, T. Suzuki, Three fixed point theorems for generalized contractions with constants in complete metric spaces, Nonlinear Analysis, 69 (2008), 2942- 2949. [12] Z. Ma and L. Jiang, C∗-Algebra-valued b-metric spaces and related fixed point theorems, Fixed point Theory and Appl., 2015, Article ID 222 (2015). [13] Z. Ma, L. Jiang and H. Sun, C∗-Algebra-valued metric spaces and related fixed point theorems, Fixed point Theory and Appl., 2014, Article ID 206 (2014). [14] B. Moeini, A.H. Ansari and C. Park, C∗-algebra-valued modular metric spaces and related fixed point results, submitted. [15] C. Mongkolkeha, W. Sintunavarat and P. Kumam, Fixed point theorems for contraction mappings in modular metric spaces, Fixed Point Theory and Ap- plications, 2011(2011), Article ID 93. 18 Moeini and Hojat Ansari [16] C. Mongkolkeha, W. Sintunavarat and P. Kumam, Fixed point theorems for contraction mappings in modular metric spaces, Fixed Point Theory and Ap- plications, 2012(2012), Article ID 103. [17] G. Mot, A. Petru¸sel, Fixed point theory for a new type of contractive multival- ued operators, Nonlinear Analysis, 70 (2009), 3371-3377. [18] D. Shehwar and T. Kamran, C∗-Valued G-contraction and fixed points, Journal of Inequalities and Appl., 2015, Article ID 304 (2015). [19] T. Suzuki, A new type of fixed point theorem in metric spaces, Nonlinear Anal- ysis, 71 (2009), 5313-5317. [20] T. Suzuki, A generalized Banach contraction principle that characterizes metric completeness, Proc. Amer. Math. Soc., 136 (2008), 1861-1869. [21] A. Zada, S. Saifullah and Z. Ma, Common fixed point theorems for G- contraction in C∗-algebra-valued metric spaces, International Journal of Anal- ysis and Applications, Vol. 11, no. 1 (2016), 23-27.
1009.4821
1
1009
2010-09-24T12:28:18
On the two-dimensional moment problem
[ "math.FA" ]
In this paper we obtain an algorithm towards solving the two-dimensional moment problem. This algorithm gives the necessary and sufficient conditions for the solvability of the moment problem. It is shown that all solutions of the moment problem can be constructed using this algorithm. In a consequence, analogous results are obtained for the complex moment problem.
math.FA
math
ON THE TWO-DIMENSIONAL MOMENT PROBLEM SERGEY ZAGORODNYUK1 ∗ Abstract. In this paper we obtain an algorithm towards solving the two- dimensional moment problem. This algorithm gives the necessary and sufficient conditions for the solvability of the moment problem. It is shown that all solutions of the moment problem can be constructed using this algorithm. In a consequence, analogous results are obtained for the complex moment problem. 1. Introduction and preliminaries ZR2 In this paper we analyze the two-dimensional moment problem. Recall that this problem consists of finding a non-negative Borel measure µ in R2 such that 1 xn xm 2 dµ = sm,n, m, n ∈ Z+, (1.1) where {sm,n}m,n∈Z+ is a prescribed sequence of complex numbers. The two-dimensional moment problem and the (closely related to this subject) complex moment problem have an extensive literature, see books [7], [1], [3], sur- veys [5],[4] and [8]. Some conditions of solvability for this moment problem were obtained by Kilpi and by Stochel and Szafraniec, see e.g. [1] and [8]. However, these conditions are hard to check. Putinar and Vasilescu derived conditions of solvability and a description of all solutions by means of a dimensional exten- sion [6] (even for the N-dimensional moment problem). The two-dimensional moment problem is solvable if and only if the prescribed sequence of moments can be extended to a sequence {sm,n,k}m,n,k∈Z+, satisfying some easy conditions (including the positivity condition). This extended sequence is the moment se- quence for an extended moment problem: ZR2 xm 1 xn 2 1 1 + x2 (1 + x2 2)k dµ = sm,n,k, m, n, k ∈ Z+. (1.2) The unique solution of the moment problem (1.2) provides a solution of the two-dimensional moment problem. In this way all different extensions define all different solutions of the two-dimensional moment problem. However, it is not clear whether such extensions exist and what is a procedure for the construction of extensions {sm,n,k}m,n,k∈Z+. The method of our investigation uses an abstract operator approach, see [9]. Firstly, we obtain a solvability criterion for an auxiliary extended two-dimensional moment problem. This moment problem is somewhat similar to the moment Date: Received: xxxxxx; Revised: yyyyyy; Accepted: zzzzzz. ∗ Corresponding author. Key words and phrases. moment problem, Hilbert space, linear equation. 1 2 S. ZAGORODNYUK problem (1.2) but we do not see any direct relationship. It is shown that the extended two-dimensional moment problem is always determinate and its solution can be constructed explicitly. An idea of our algorithm is to extend the symmetric operators related to the two-dimensional moment problem, not "entirely", but on a discrete set of points. It is shown that all solutions of the moment problem (1.1) can be constructed on this way. Roughly speaking, the final algorithm reduces to the solving of finite and infinite linear systems of equations with parameters. In a consequence, analogous results are obtained for the complex moment prob- lem. Notations. As usual, we denote by R, C, N, Z, Z+ the sets of real numbers, complex numbers, positive integers, integers and non-negative integers, respec- tively. The real plane will be denoted by R2. For a subset S of R2 we denote by B(S) the set of all Borel subsets of S. Everywhere in this paper, all Hilbert spaces are assumed to be separable. By (·,·)H and k·kH we denote the scalar product and the norm in a Hilbert space H, respectively. The indices may be omitted in obvi- ous cases. For a set M in H, by M we mean the closure of M in the norm k· kH. For {xm,n}m,n∈Z+, xm,n ∈ H, we write Lin{xm,n}m,n∈Z+ for the set of linear combi- nations of elements {xm,n}m,n∈Z+ and span{xm,n}m,n∈Z+ = Lin{xm,n}m,n∈Z+. The identity operator in H is denoted by EH. For an arbitrary linear operator A in H, the operators A∗,A,A−1 mean its adjoint operator, its closure and its inverse (if they exist). By D(A) and R(A) we mean the domain and the range of the op- erator A. The norm of a bounded operator A is denoted by kAk. By P H H1 = PH1 we mean the operator of orthogonal projection in H on a subspace H1 in H. By L2 µ we denote the usual space of square-integrable complex functions f (x1, x2), x1, x2 ∈ R2, with respect to the Borel measure µ in R2. 2. The solution of an extended two-dimensional moment problem. Consider the following moment problem: to find a non-negative Borel measure µ in R2 such that ZR2 1 (x1 + i)k(x1 − i)lxn xm 2 (x2 + i)r(x2 − i)tdµ = um,k,l;n,r,t, (2.1) where {um,k,l;n,r,t}m,n∈Z+,k,l,r,t∈Z is a prescribed sequence of complex numbers. This problem is said to be the extended two-dimensional moment problem. We set m, n ∈ Z+, k, l, r, t ∈ Z, Ω = {(m, k, l; n, r, t) : m, n ∈ Z+, k, l, r, t ∈ Z}, Ω0 = {(m, k, l; n, r, t) : m, n ∈ Z+, k, l, r, t ∈ Z, k = l = r = t = 0}, Ω′ = Ω\Ω0. Let the moment problem (2.1) have a solution µ. Choose an arbitrary function P (x1, x2) = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)t, ON THE TWO-DIMENSIONAL MOMENT PROBLEM 3 where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Then 0 ≤ZR2 P (x1, x2)2dµ = ∗ZR2 xm+m′ 1 (x1 + i)k+l′ (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′ X (x1 − i)l+k′ αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′um+m′,k+l′,l+k′;n+n′,r+t′,t+r′. (x2 − i)t+r′ (x2 + i)r+t′ xn+n′ 2 dµ = Therefore (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω X X αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′um+m′,k+l′,l+k′;n+n′,r+t′,t+r′ ≥ 0, (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (2.2) for arbitrary complex coefficients αm,k,l;n,r,t, where all but finite number of αm,k,l;n,r,t are zeros. The latter condition on the coefficients αm,k,l;n,r,t in infinite sums will be assumed in similar situations. We shall use the following important fact (e.g. [2, pp.361-363]). Theorem 2.1. Let a sequence of complex numbers {um,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy condition (2.2). Then there exist a separable Hilbert space H with a scalar product (·,·)H and a sequence {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω in H, such that (xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω, (2.3) and span{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω = H. Proof. (We do not claim the originality of the idea of this proof). Choose an arbi- trary infinite-dimensional linear vector space V (for instance, one may choose the space of all complex sequences (un)n∈N, un ∈ C). Let X = {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω be an arbitrary infinite sequence of linear independent elements in V which is in- dexed by elements of Ω. Set LX = Lin{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω. Introduce the following functional: [x, y] = αm,k,l;n,r,tβm′,k′,l′;n′,r′,t′ (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω X ∗ um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (2.4) for x, y ∈ LX , x = X(m,k,l;n,r,t)∈Ω y = X(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,txm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′xm′,k′,l′;n′,r′,t′, where αm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′ ∈ C. Here all but finite number of indices αm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′ are zeros. The set LX with [·,·] will be a quasi-Hilbert space. Factorizing and making the completion we obtain the desired space H (e.g. [3]). (cid:3) A0 X(m,k,l;n,r,t)∈Ω B0 X(m,k,l;n,r,t)∈Ω x = X(m,k,l;n,r,t)∈Ω We may write αm,k,l;n,r,txm,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t = X(m′,k′,l′;n′,r′,t′)∈Ω  X(m,k,l;n,r,t)∈Ω = X(m,k,l;n,r,t)∈Ω = X(m,k,l;n,r,t)∈Ω  X(m′,k′,l′;n′,r′,t′)∈Ω = (x, xa+1,b,c;d,e,f )H , = (x, xa+1,b,c;d,e,f )H , αm,k,l;n,r,txm+1,k,l;n,r,t, xa,b,c;d,e,fH αm,k,l;n,r,tum+1+a,k+c,l+b;n+d,r+f,t+e αm,k,l;n,r,t(xm,k,l;n,r,t, xa+1,b,c;d,e,f )H (a, b, c; d, e, f ) ∈ Ω. αm′,k′,l′;n′,r′,t′xm′+1,k′,l′;n′,r′,t′, xa,b,c;d,e,fH (a, b, c; d, e, f ) ∈ Ω. In the same manner we obtain: 4 S. ZAGORODNYUK Let the moment problem (2.1) be given and the condition (2.2) hold. By The- orem 2.1 there exist a Hilbert space H and a sequence {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H, such that relation (2.3) holds. Set L = Lin{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω. Introduce the following operators where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Let us check that these definitions are correct. Indeed, suppose that αm,k,l;n,r,txm+1,k,l;n,r,t, αm,k,l;n,r,txm,k,l;n+1,r,t, (2.5) (2.6) βm′,k′,l′;n′,r′,t′xm′,k′,l′;n′,r′,t′. (2.7) Therefore the definition of A0 is correct. The correctness of the definition of B0 can be checked in a similar manner. Notice that (A0xm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm+1,k,l;n,r,t, xa,b,c;d,e,f )H = um+1+a,k+c,l+b;n+d,r+f,t+e = (xm,k,l;n,r,t, xa+1,b,c;d,e,f )H = (xm,k,l;n,r,t, A0xa,b,c;d,e,f )H , (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. Therefore A0 is symmetric. The same argument implies that B0 is symmetric, as well. Suppose that the following conditions hold: um+1+a,k+c,l+b;n+d,r+f,t+e + ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+1+c,l+b;n+d,r+f,t+e, um+1+a,k+c,l+b;n+d,r+f,t+e − ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+c,l+1+b;n+d,r+f,t+e, (2.8) (2.9) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 5 um+a,k+c,l+b;n+1+d,r+f,t+e + ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+c,l+b;n+d,r+1+f,t+e, (2.10) um+a,k+c,l+b;n+1+d,r+f,t+e − ium+a,k+c,l+b;n+d,r+f,t+e (2.11) for all (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. These conditions are equivalent to conditions = um+a,k+c,l+b;n+d,r+f,t+1+e, (xm+1,k,l;n,r,t + ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k+1,l;n,r,t, xa,b,c;d,e,f)H, (xm+1,k,l;n,r,t − ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k,l+1;n,r,t, xa,b,c;d,e,f)H, (xm,k,l;n+1,r,t + ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k,l;n,r+1,t, xa,b,c;d,e,f )H, (2.12) (2.13) (2.14) (xm,k,l;n+1,r,t − ixm,k,l;n,r,t, xa,b,c;d,e,f )H (2.15) for all (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. The latter conditions are equivalent to the following conditions: = (xm,k,l;n,r,t+1, xa,b,c;d,e,f)H, xm+1,k,l;n,r,t + ixm,k,l;n,r,t = xm,k+1,l;n,r,t, xm+1,k,l;n,r,t − ixm,k,l;n,r,t = xm,k,l+1;n,r,t, xm,k,l;n+1,r,t + ixm,k,l;n,r,t = xm,k,l;n,r+1,t, xm,k,l;n+1,r,t − ixm,k,l;n,r,t = xm,k,l;n,r,t+1, for all (m, k, l; n, r, t) ∈ Ω. The last conditions mean that (A0 + iEH )xm,k,l;n,r,t = xm,k+1,l;n,r,t, (A0 − iEH)xm,k,l;n,r,t = xm,k,l+1;n,r,t, (B0 + iEH )xm,k,l;n,r,t = xm,k,l;n,r+1,t, (B0 − iEH )xm,k,l;n,r,t = xm,k,l;n,r,t+1, for all (m, k, l; n, r, t) ∈ Ω. The latter conditions imply that (B0 ± iEH )L = L. (A0 ± iEH )L = L, (2.16) (2.17) (2.18) (2.19) (2.20) (2.21) (2.22) (2.23) Therefore operators A0 and B0 are essentially self-adjoint. The conditions (2.20)- (2.23) also imply that (A0 + iEH)−1xm,k,l;n,r,t = xm,k−1,l;n,r,t, (A0 − iEH )−1xm,k,l;n,r,t = xm,k,l−1;n,r,t, (B0 + iEH )−1xm,k,l;n,r,t = xm,k,l;n,r−1,t, (B0 − iEH )−1xm,k,l;n,r,t = xm,k,l;n,r,t−1, for all (m, k, l; n, r, t) ∈ Ω. Consider the Cayley transformations of A0 and B0: VA0 = (A0 − iEH )(A0 + iEH )−1, VB0 = (B0 − iEH )(B0 + iEH )−1, D(A0) = D(B0) = L. (2.24) (2.25) (2.26) (2.27) (2.28) (2.29) 6 S. ZAGORODNYUK By virtue of relations (2.21),(2.23),(2.24),(2.26) we obtain: VA0VB0xm,k,l;n,r,t = xm,k−1,l+1;n,r−1,t+1 = VB0VA0xm,k,l;n,r,t, for all (m, k, l; n, r, t) ∈ Ω. Therefore VA0VB0x = VB0VA0x, x ∈ L. x ∈ H. By continuity we extend the isometric operators VA0 and VB0 to unitary operators UA0 and VB0 in H, respectively. By continuity we conclude that UA0UB0x = UB0UA0x, Set A = A0, B = B0. The Cayley transformations of the self-adjoint operrators A and B coincide on L with UA0 and UB0, respectively. Thus, the Cayley trans- formations of A and B are UA0 and UB0, respectively. Therefore, operators A and B commute. Notice that (2.30) (2.31) (2.32) xm,k,l;n,r,t = Am(A + i)k(A − i)lBn(B + i)r(B − i)tx0,0,0;0,0,0, for all (m, k, l; n, r, t) ∈ Ω. In fact, by induction we can check that xm,k,l;n,r,t = (B − iEH )txm,k,l;n,r,0, t ∈ Z, for any fixed m, n ∈ Z+, k, l, r ∈ Z; xm,k,l;n,r,0 = (B + iEH )rxm,k,l;n,0,0, for any fixed m, n ∈ Z+, k, l ∈ Z; xm,k,l;n,0,0 = Bnxm,k,l;0,0,0, for any fixed m ∈ Z+, k, l ∈ Z; for any fixed m ∈ Z+, k ∈ Z; xm,k,l;0,0,0 = (A − iEH )lxm,k,0;0,0,0, xm,k,0;0,0,0 = (A + iEH )lxm,0,0;0,0,0, for any fixed m ∈ Z+; xm,0,0;0,0,0 = Amx0,0,0;0,0,0, n ∈ Z+, r ∈ Z, l ∈ Z, k ∈ Z, m ∈ Z+, and then by substitution of each relation into previous one we obtain rela- tion (2.32). For the commuting self-adjoint operators A and B there exists an orthogonal operator spectral measure E(x) on B(R2) such that A =ZR2 um,k,l;n,r,t = (xm,k,l;n,r,t, x0,0,0;0,0,0)H (2.33) x2dE(x). x1dE(x), B =ZR2 2 (x2 + i)r(x2 − i)tdE(x)x0,0,0;0,0,0, x0,0,0;0,0,0(cid:19)H 2 (x2 + i)r(x2 − i)td(E(x)x0,0,0;0,0,0, x0,0,0;0,0,0)H. xm 1 (x1 + i)k(x1 − i)lxn xm 1 (x1 + i)k(x1 − i)lxn Then =(cid:18)ZR2 =ZR2 ON THE TWO-DIMENSIONAL MOMENT PROBLEM 7 Hence, the Borel measure µ = (E(x)x0,0,0;0,0,0, x0,0,0;0,0,0)H, (2.34) is a solution of the moment problem (2.1). Theorem 2.2. Let the extended two-dimensional moment problem (2.1) be given. The moment problem has a solution if and only if conditions (2.2) and (2.8)- (2.11) are satisfied. If these conditions are satisfied then the solution of the mo- ment problem is unique and can be constructed by (2.34). Proof. The sufficiency of conditions (2.2) and (2.8)-(2.11) for the existence of a solution of the moment problem (2.1) was shown before the statement of the Theorem. The necessity of condition (2.2) was proved, as well. Let us check that conditions (2.8)-(2.11) are necessary for the solvability of the moment prob- lem (2.1). Let µ be a solution of the moment problem (2.1). Consider the space L2 following subsets in L2 µ: Lµ = Lin{xm We denote ym,k,l;n,r,t := xm Notice that 2 (x2 + i)r(x2− i)t}(m,k,l;n,r,t)∈Ω, Hµ = Lµ. (2.35) (m, k, l; n, r, t) ∈ Ω. (2.36) 1 (x1 + i)k(x1− i)lxn 1 (x1+i)k(x1−i)lxn 2 (x2+i)r(x2−i)t, µ and the (ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 (2.37) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Consider the operators of multipli- cation by the independent variable in L2 µ: µ = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, Aµf (x1, x2) = x1f (x1, x2), Bµf (x1, x2) = x2f (x1, x2), Notice that f ∈ L2 µ. (Aµ + iEL2 (Aµ − iEL2 (Bµ + iEL2 (Bµ − iEL2 µ)ym,k,l;n,r,t = ym,k+1,l;n,r,t, µ)ym,k,l;n,r,t = ym,k,l+1;n,r,t, µ)ym,k,l;n,r,t = ym,k,l;n,r+1,t, µ)ym,k,l;n,r,t = ym,k,l;n,r,t+1, (2.38) (2.39) (2.40) (2.41) (2.42) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Since conditions (2.2) are satisfied, by Theorem 2.1 there exist a Hilbert space H and a sequence of elements {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H, such that relation (2.3) holds. Repeating arguments after the Proof of Theorem 2.1 we construct opera- tors A0 and B0 in H. Consider the following operator: W0 X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,tym,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t, (2.43) where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Let us check that this operator is defined correctly. In fact, suppose that x = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,tym,k,l;n,r,t = X(m′,k′,l′;n′,r′,t′)∈Ω βm′,k′,l′;n′,r′,t′ym′,k′,l′;n′,r′,t′, (2.44) 8 S. ZAGORODNYUK 2 L2 µ . where βm′,k′,l′;n′,r′,t′ ∈ C. We may write ∗(αm′,k′,l′;n′,r′,t′ − βm′,k′,l′;n′,r′,t′)(ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 µ = = (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (αm,k,l;n,r,t − βm,k,l;n,r,t) (αm,k,l;n,r,t − βm,k,l;n,r,t)ym,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 0 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) X(m,k,l;n,r,t)∈Ω X X (αm,k,l;n,r,t − βm,k,l;n,r,t)xm,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) X(m,k,l;n,r,t)∈Ω Thus, the operator W0 is defined correctly. Ifex ∈ H and ex = X(m,k,l;n,r,t)∈Ω X where γm,k,l;n,r,t ∈ C, then (αm,k,l;n,r,t − βm,k,l;n,r,t) γm,k,l;n,r,tym,k,l;n,r,t, (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω ∗(αm′,k′,l′;n′,r′,t′ − βm′,k′,l′;n′,r′,t′)(xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H (W0x, W0ex)H = αm,k,l;n,r,tγm′,k′,l′;n′,r′,t′ = (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω X ∗(xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H αm,k,l;n,r,tγm′,k′,l′;n′,r′,t′(ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)H . µ = (x,ex)L2 By continuity we extend W0 to a unitary operator W which maps Hµ onto H. Observe that W −1A0W ym,k,l;n,r,t = ym+1,k,l;n,r,t = Aµym,k,l;n,r,t, W −1B0W ym,k,l;n,r,t = ym,k,l;n+1,r,t = Bµym,k,l;n,r,t, (2.46) for all (m, k, l; n, r, t) ∈ Ω. By using the last relations in relations (2.39)-(2.42) we obtain relations (2.20)-(2.23). The latter relations are equivalent to condi- tions (2.8)-(2.11). (2.45) Let us check that the solution of the moment problem is unique. Consider the following transformation T : (x1, x2) ∈ R2 7→ (ϕ, ψ) ∈ [0, 2π) × [0, 2π), eiϕ = and set x1 + i x1 − i , eiψ = x2 + i x2 − i ; ν(∆) = µ(T −1(∆)), ∆ ∈ B([0, 2π) × [0, 2π)). (2.47) (2.48) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 9 Since T is a bijective continuous transformation, then ν is a non-negative Borel measure on [0, 2π) × [0, 2π). Moreover, we have x1 − i(cid:19)k(cid:18) x2 + i x2 − i(cid:19)l u0,k,−k;0,l,−l =ZR2(cid:18) x1 + i for all k, l ∈ Z. Leteµ be another solution of the moment problem (2.1) andeν be dµ =Z[0,2π)×[0,2π) eikϕeilψdν, defined by (2.49) (2.50) By relation (2.49) we obtain that eν(∆) =eµ(T −1(∆)), ∆ ∈ B([0, 2π) × [0, 2π)). eikϕeilψdν =Z[0,2π)×[0,2π) k, l ∈ Z. Z[0,2π)×[0,2π) (2.51) By the Weierstrass theorem we can approximate ϕm and ψn, for some fixed m, n ∈ Z+, by trigonometric polynomials Pk(ϕ) and Rk(ψ), respectively: eikϕeilψdeν, ϕ∈[0,2π)ϕm − Pk(ϕ) ≤ max , max ψ∈[0,2π)ψm − Rk(ψ) ≤ , k ∈ N. (2.52) 1 k 1 k Then =(cid:12)(cid:12)(cid:12)(cid:12)Z[0,2π)×[0,2π) (cid:12)(cid:12)(cid:12)(cid:12)Z[0,2π)×[0,2π) ϕmψndν −Z[0,2π)×[0,2π) (ϕm − Pk(ϕ))ψndν +Z[0,2π)×[0,2π) ≤ max ψ∈[0,2π)ψm ν([0, 2π)) + max ϕ∈[0,2π)Pk(ϕ) 1 k Pk(ϕ)Rk(ψ)dν(cid:12)(cid:12)(cid:12)(cid:12) Pk(ϕ)(ψn − Rk(ψ))dν(cid:12)(cid:12)(cid:12)(cid:12) ν([0, 2π)) 1 k ≤ max ψ∈[0,2π)ψm k as k → ∞. In the same manner we get ν([0, 2π)) → 0, k 1 k + max ν([0, 2π)) +(cid:18) 1 ϕmψndeν −Z[0,2π)×[0,2π) ϕmψndν =Z[0,2π)×[0,2π) ϕ∈[0,2π)ϕm(cid:19) 1 Pk(ϕ)Rk(ψ)deν(cid:12)(cid:12)(cid:12)(cid:12) → 0, ϕmψndeν, m, n ∈ Z+. (cid:12)(cid:12)(cid:12)(cid:12)Z[0,2π)×[0,2π) Z[0,2π)×[0,2π) as k → ∞. Hence, we conclude that Since the two-dimensional moment problem on a rectangular has a unique solu- (cid:3) (2.53) tion, we get ν =eν and µ =eµ. 3. An algorithm towards solving the two-dimensional moment As a first application of our results on the extended two-dimensional moment problem we get the following theorem. problem. 10 S. ZAGORODNYUK Theorem 3.1. Let the two-dimensional moment problem (1.1) be given. The moment problem has a solution if and only if there exists a sequence of complex numbers {um,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, which satisfies conditions (2.2), (2.8)-(2.11) and (3.1) um,0,0;n,0,0 = sm,n, m, n ∈ Z+. The proof is obvious and left to the reader. Let the two-dimensional moment problem (1.1) be given. As it is well known (and can be checked in the same manner as for the relation (2.2)) the necessary condition for its solvability is the following: Xm,n,m′,n′∈Z+ αm,nαm′,n′sm+m′,n+n′ ≥ 0, (3.2) for arbitrary complex coefficients αm,n, where all but finite number of αm,n are zeros. We assume that the condition (3.2) holds. Repeating arguments of the proof of Theorem 2.1 we can state that there exist a Hilbert space H0 and a sequence {hm,n}m,n∈Z+ such that (hm,n, hm′,n′)H0 = sm+m′,n+n′, m, n, m′, n′ ∈ Z+. Consider the following Hilbert space: H = H0 ⊕ ∞Mj=1 Hj! , (3.3) (3.4) (3.5) (3.6) (3.7) (3.8) where Hj are arbitrary one-dimensional Hilbert spaces, j ∈ N. We shall call it the model space for the two-dimensional moment problem. Introduce an arbitrary indexation in the set Ω′ by the unique positive integer index j: j ∈ N 7→ w = w(j) = (m, k, l; n, r, t)(j) ∈ Ω′. Suppose that the two-dimensional moment problem (1.1) has a solution µ. Consider the space L2 µ and the following subsets in L2 µ: Lµ,0 = Lin{xm 2}m,n∈Z+, Hµ,0 = Lµ,0. 1 xn We denote Notice that ym,n := xm 1 xn 2 , m, n ∈ Z+. (ym,n, ym′,n′)L2 µ = sm+m′,n+n′, for all m, n, m′, n′ ∈ Z+. We shall also use the notations from (2.35),(2.36). Define the following numbers um,k,l;n,r,t :=ZR2 xm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)tdµ, (3.9) For these numbers conditions (2.2) hold and repeating arguments after the rela- tion (2.42) we construct a Hilbert space H and a sequence of elements {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, (m, k, l; n, r, t) ∈ Ω. ON THE TWO-DIMENSIONAL MOMENT PROBLEM 11 in H, such that relation (2.3) holds. We introduce the operator W as after (2.43). The operator W maps Hµ onto H. Set H0 = span{xm,0,0;n,0,0}m,n∈Z+ ⊆ H. (3.10) Let us construct a sequence of Hilbert spaces Hj, j ∈ N, in the following way. Step 1. We set f1 = xw(1) − P H H0 xw(1), H1 = span{f1}, where w(·) is the indexation in the set Ω′. Step r, with r ≥ 2. We set fr = xw(r) − P H Then we get a representation H0⊕(⊕1≤t≤r−1Ht)xw(r), Hr = span{fr}. H = H0 ⊕ ∞Mj=1 Hj! . (3.11) (3.12) (3.13) Observe that Hj is either a one-dimensional Hilbert space or Hj = {0}. We denote Λµ = {j ∈ N : Hj 6= {0}}, Λ′ µ = N\Λµ. Then (3.14) (3.15) We shall construct a unitary operator U which maps H onto the following sub- space of the model space H: Hj . H = H0 ⊕Mj∈Λµ Hj ⊂ H. bH = H0 ⊕Mj∈Λµ αm,nxm,0,0;n,0,0 +Xj∈Λµ αm,nhm,n +Xj∈Λµ βj x = Xm,n∈Z+ U x = Xm,n∈Z+ Choose an arbitrary element with αm,n, βj ∈ C. Set (3.16) , (3.17) (3.18) fj kfjkH βjej, where ej ∈ Hj, kejkH = 1, are chosen arbitrarily. Let us check that this definition is correct. Suppose that x has another represen- tation: x = Xm,n∈Z+eαm,nxm,0,0;n,0,0 +Xj∈Λµeβj fj kfjkH , (3.19) 12 S. ZAGORODNYUK 2 H . Xm,n∈Z+ = Xm,n,m′,n′∈Z+ with eαm,n,eβj ∈ C. By orthogonality we have βj = eβj, j ∈ N. Then (αm,n −eαm,n)xm,0,0;n,0,0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 0 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (αm,n −eαm,n)(αm′,n′ −eαm′,n′)(xm,0,0;n,0,0, xm′,0,0;n′,0,0)H = Xm,n,m′,n′∈Z+ (αm,n −eαm,n)(αm′,n′ −eαm′,n′)(hm,n, hm′,n′)H (αm,n −eαm,n)hm,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xm,n∈Z+ Thus, the operator U is defined correctly. Ifbx ∈ H and bx = Xm,n∈Z+bαm,nxm,0,0;n,0,0 +Xj∈Λµbβj where bαm,n,bβj ∈ C, then (x,bx)H = Xm,n,m′,n′∈Z+ = Xm,n,m′,n′∈Z+ αm,nbαm′,n′(xm,0,0;n,0,0, xm′,0,0;n′,0,0)H +Xj∈Λµ αm,nbαm′,n′(hm,n, hm′,n′)H +Xj∈Λµ βjbβj = (U x, Ubx)H. fj kfjkH , βjbβj By continuity we extend U to a unitary operator which maps H onto bH. Then the operator U W is a unitary operator which maps Hµ onto bH. We could define this operator directly, but we prefer to underline an abstract structure of the corresponding spaces and this maybe explains where the model space comes from. We set Observe that hm,k,l;n,r,t := U W ym,k,l;n,r,t, (m, k, l; n, r, t) ∈ Ω. hm,0,0;n,0,0 = hm,n, m, n ∈ Z+; U W Hµ,0 = H0. (3.20) (3.21) (3.22) (3.23) Since then xw(r) ∈ H0 ⊕ Mj∈Λµ: j≤r Hj , hw(r) ∈ H0 ⊕ Mj∈Λµ: j≤r Hj , r ∈ N. Observe that {ym,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy relations (2.16)-(2.19) (with y instead of x). Therefore {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy relations, as well. Notice that (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 µ ON THE TWO-DIMENSIONAL MOMENT PROBLEM 13 = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (3.24) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Choose an arbitrary r ∈ N. By (3.12) we may write xw(r) = fr + P H H0⊕(⊕1≤t≤r−1Ht)xw(r) = kfrk fr kfrk + Xt∈Λµ: 1≤t≤r−1 βt ft kftk + u0, (3.25) where βt ∈ C, u0 ∈ H0. By (3.20) and (3.18) we get hw(r) = U xw(r) = kfrker + Xt∈Λµ: 1≤t≤r−1 βtet + w0, (3.26) where w0 = U u0 ∈ H0. Therefore (hw(r), er) ≥ 0; and hw(r) ∈ H0 ⊕ Mt∈Λµ: 1≤t≤r−1 In particular, we may write Ht ⇔ (hw(r), er) = 0 ⇔ fr = 0 ⇔ r ∈ Λ′ µ. r ∈ Λµ ⇔ (hw(r), er) > 0, r ∈ N. (3.27) (3.28) (3.29) Theorem 3.2. Let the two-dimensional moment problem (1.1) be given and con- dition (3.2) holds. Choose an arbitrary model space H with a sequence {hm,n}m,n∈Z+, satisfying (3.3) and fix it. The moment problem has a solution if and only if there exists a sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H such that the following conditions hold: 1) hm,0,0;n,0,0 = hm,n, m, n ∈ Z+; 2) hw(r) ∈ H0 ⊕(cid:16)Lj∈Λ: 0≤j≤r Hj(cid:17), and (hw(r), er) ≥ 0, r ∈ N, for some 3) The sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfies conditions (2.16)-(2.19) (with 4) There exists a complex function ϕ(m, k, l; n, r, t), (m, k, l; n, r, t) ∈ Ω, such subset Λ ⊂ N. h instead of x). that (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = ϕ(m + m′, k + l′, l + k′; n + n′, r + t′, t + r′), for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Proof. The necessity of conditions 1)-4) for the solvability of the two-dimensional moment problem was established before the statement of the Theorem. Let conditions 1),3),4) be satisfied. Consider the extended two-dimensional mo- ment problem (2.1) with um,k,l;n,r,t := ϕ(m, k, l; n, r, t), (m, k, l; n, r, t) ∈ Ω, (3.30) 14 S. ZAGORODNYUK where ϕ is from the condition 4). Then X X (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω = (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′um+m′,k+l′,l+k′;n+n′,r+t′,t+r′ αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′(hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,thm,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 H ≥ 0, for arbitrary complex coefficients αm,k,l;n,r,t, where all but finite number of αm,k,l;n,r,t are zeros. By conditions 3) and 4) we conclude that conditions (2.8)-(2.11) hold. By The- orem 2.2 we obtain that there exists a non-negative Borel measure µ in R2 such that (2.1) holds. In particular, using conditions 4),1) we get ZR2 xm 1 xn 2 dµ = um,0,0;n,0,0 = ϕ(m, 0, 0; n, 0, 0) = (hm,0,0;n,0,0, h0,0,0;0,0,0)H = (hm,n, h0,0)H = sm,n, m, n ∈ Z+. (cid:3) Observe that condition 2) can be removed from the statement of Theorem 3.2. However, it will be used later. Denote a set of sequences {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H satisfying conditions 1)-4) by X = X(H). As we have seen in the proof of Theorem 3.2, for an arbi- trary {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H), the unique solution of the extended two- dimensional moment problem with moments (3.30) gives a solution of the two- dimensional moment problem. Observe that all solutions of the two-dimensional moment problem can be constructed in this manner. Indeed, let µ be an arbitrary solution of the two-dimensional moment problem. Repeating arguments from relation (3.6) till the statement of Theorem 3.2 we may write ZR2 xm 1 (x1 + i)k(x1 − i)lxn = (U W ym,k,l;n,r,t, U W y0,0,0;0,0,0)H = (hm,k,l;n,r,t, h0,0,0;0,0,0)H 2 (x2 + i)r(x2 − i)tdµ = (ym,k,l;n,r,t, y0,0,0;0,0,0)L2 µ = ϕ(m, k, l; n, r, t) = um,k,l;n,r,t, for all (m, k, l; n, r, t) ∈ Ω. Here the operators U,W , the sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, the function ϕ(m, k, l; n, r, t) and the moments um,k,l;n,r,t, of course, depend on the choice of µ. Notice that {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H). For such constructed parameters, the measure µ is a solution of the extended two-dimensional moment problem considered in the proof of Theorem 3.2. Since the solution of this moment problem is unique, µ will be reconstructed in the above described manner. Notice that condition 4) of Theorem 3.2 is equivalent to the following condi- tions: (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (h em,k,l;n,r,t, h em′,k′,l′;n′,r′,t′)H, (3.31) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 15 (3.32) (3.33) (3.34) (3.35) (3.36) (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;en,r,t, hm′,k′,l′;en′,r′,t′)H, (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,ek,l;n,r,t, hm′,k′,el′;n′,r′,t′)H, (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,el;n,r,t, hm′,ek′,l′;n′,r′,t′)H, if m, m′,em,em′, n, n′ ∈ Z+, k, l, r, t, k′, l′, r′, t′ ∈ Z: m + m′ = em + em′; if m, m′, n, n′ ∈ Z+, k,ek, l, r, t, k′, l′,el′, r′, t′ ∈ Z: k + l′ =ek +el′; if m, m′, n, n′ ∈ Z+, k, l,el, r, t, k′,ek′, l′, r′, t′ ∈ Z: l + k′ =el +ek′; if m, m′, n, n′,en,en′ ∈ Z+, k, l, r, t, k′, l′, r′, t′ ∈ Z: n + n′ =en +en′; if m, m′, n, n′ ∈ Z+, k, l, r,er, t, k′, l′, r′, t′,et′ ∈ Z: r + t′ =er +et′; if m, m′, n, n′ ∈ Z+, k, l, r, t,et, k′, l′, r′,er′, t′ ∈ Z: t + r′ =et +er′. (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;n,er,t, hm′,k′,l′;n′,r′,et′)H, (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;n,r,et, hm′,k′,l′;n′,er′,t′)H, Let {gn}∞ As we can see, the solving of the two-dimensional moment problem reduces to a construction of the set X(H). Let us describe an algorithm for a construction of sequences from X(H). n=1 be an arbitrary orthonormal basis in H0 obtained by the Gram- Schmidt orthogonalization procedure from the sequence {hm,n}m,n∈Z+ indexed by a unique index. If we had constructed {hm,n}m,n∈Z+ ∈ X(H), then the two-dimensional moment problem has a solution µ and Choose an arbitrary j ∈ N. Let w(j) = (m, k, l; n, r, t)(j) ∈ Ω′. dj := khw(j)k2 H = kxm =ZR2 x2m 1 (x2 1 (x1 + i)k(x1 − i)lxn 1 + 1)k(x2 1 + 1)lx2n 2 (x2 2 (x2 + i)r(x2 − i)tk2 2 + 1)r(x2 2 + 1)tdµ. L2 µ Therefore dj are bounded by some constants Mj = Mj(S) depending on the prescribed moments S := {sm,n}m,n∈Z+. (Notice that e.g. (x2 1 + 1)l ≤ 1, for l < 0, and for non-negative m, k, l; n, r, t the values of dj are determined uniquely). Step 0. We set hm,0,0;n,0,0 = hm,n, m, n ∈ Z+. (3.37) We check that conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) are satisfied for hm,0,0;n,0,0, m, n ∈ Z0. If they are not satisfied, the two-dimensional moment problem has no solution and we stop the algorithm. Step 1. We seek for hw(1) in the following form: hw(1) = α1;ngn + β1;1e1, (3.38) ∞Xn=1 with some complex coefficients α1;n, β1;1. Conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) which include hw(1) and the already constructed hm,k,l;n,r,t are equivalent to a set L1 of linear Set Finally, we set bS1 =((α1;n, n ∈ N; d1) ∈ S1 : ∞Xn=1 G1 = α1;n2! 1 ∞Xn=1 ∞Xn=1 α1;ngn + d1 − 2 The case G1 = ∅ is not excluded. Step r, with r ≥ 2. We seek for hw(r) in the following form: α1;n2 ≤ d1, d1 ≤ M1) . e1 : (α1;n, n ∈ N; d1) ∈ bS1 rXj=1 βr;jej, (3.39) (3.40) . (3.41) (3.42) 16 S. ZAGORODNYUK equations with respect to α1;n, n ∈ N, and d1 = khw(1)k2 depend on β1;1 only by d1. Denote the set of solutions of these equations by H. Notice that they S1 = {(α1;n, n ∈ N; d1) : equations from L1 are satisfied} . hw(r) = αr;ngn + ∞Xn=1 with some complex coefficients αr;n, βr;j. Conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) which include hw(r) and the already constructed hm,k,l;n,r,t are equivalent to a set Lr of linear equations with respect to αr;n, n ∈ N, βr;j, 1 ≤ j ≤ r − 1, and dr = khw(r)k2 H, and depending on parameters (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1 . Notice that these linear equations depend on βr;j only by dr. Denote the set of solutions of these equations by Sr =(cid:8)(αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) : (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1, and equations from Lr with parameters (hw(1), hw(2), ..., hw(r−1)), are satisfied(cid:9) . bSr =(cid:8)(αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ Sr : (3.43) Set βr;j2 ≤ dr, dr ≤ Mr) . Finally, we set αr;n2 + Gr =(cid:8)(cid:0)hw(1), hw(2), ..., hw(r−1), βr;jej + dr − r−1Xj=1 r−1Xj=1 αr;n2 − ∞Xn=1 αr;ngn + er : βr;j2! 1 ∞Xn=1 (αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ bSro . (The case Gr = ∅ is not excluded.) Final step. Consider a space H of sequences 2 r−1Xj=1 ∞Xn=1 (3.44) (3.45) (3.46) h = (h1, h2, h3, ...), hr ∈ H, r ∈ N, ON THE TWO-DIMENSIONAL MOMENT PROBLEM 17 with the norm given by khkH = sup r∈N 1 √MrkhrkH < ∞. (3.47) For arbitrary (h1, ..., hr) ∈ Gr, we put into correspondence elements h ∈ H of the following form h = (h1, ..., hr, gr+1, gr+2, ...) : gj ∈ H, kgjkH ≤pMj, j > r. Thus, the set Gr is mapped onto a set Gr ⊂ H. Observe that all elements of Gr has the norm less or equal to 1. Set (3.48) If Gr = ∅, we set Gr = ∅. G = Gr. (3.49) ∞\r=1 If G 6= ∅, then to each (g1, g2, ...) ∈ G, we put into correspondence a sequence H = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (3.37) holds and r ∈ N. hw(r) := gr, (3.50) We state that H ∈ X(H). In fact, conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) are satisfied for hm,0,0;n,0,0, m, n ∈ Z+, by Step 0. If one of these equations include hw(r) with r ≥ 1, then we choose the maximal appearing index r. Since (hw(1), ..., hw(r)) ∈ Gr, then this equation is satisfied. Condition 2) is satisfied by the construction. Thus, if G 6= ∅, then using H we can construct a solution of the two-dimensional moment problem in the described above manner. Theorem 3.3. Let the two-dimensional moment problem (1.1) be given and con- dition (3.2) holds. Choose an arbitrary model space H with a sequence {hm,n}m,n∈Z+, satisfying (3.3) and fix it. The moment problem has a solution if and only if conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) are satisfied for hm,0,0;n,0,0 := hm,n, m, n ∈ Z+, and G 6= ∅, (3.51) where G is constructed by (3.49) according to the algorithm. If the latter conditions are satisfied then to each (g1, g2, ...) ∈ G, we put into correspondence a sequence H = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (3.37) holds and (3.52) This sequence belongs to X(H) and the unique solution of the extended two- dimensional moment problem with moments (3.30) gives a solution µ of the two- dimensional moment problem. Moreover, all solutions of the two-dimensional moment problem can be obtained in this way. hw(r) := gr, r ∈ N. Proof. The sufficiency of the conditions in the statement of the Theorem for the solvability of the two-dimensional moment problem was shown before the statement of the Theorem. Let us show that these conditions are necessary. Let µ be a solution of the two-dimensional moment problem. By Theorem 3.2 the set X(H) is not empty. Choose an arbitrarybH = {bhm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H). By conditions 3),4) we see that conditions (2.12)-(2.15) (withbh instead of x) and 18 S. ZAGORODNYUK (3.31)-(3.36) are satisfied. In particular, they are satisfied forbhm,0,0;n,0,0 = hm,n, m, n ∈ Z+. Comparing condition 2) with Steps 1 and r for r ≥ 2, we see that Therefore elements (bhw(1), ...,bhw(r)) ∈ Gr, r ∈ N. where gj ∈ H : kgjkH ≤pMj, for j > r are arbitrary. Thus, the element (bhw(1), ...,bhw(r), gr+1, gr+2, ...) ∈ Gr, bh := (bhw(1),bhw(2),bhw(3), ...) ∈ Gr, Gr = G. r ∈ N, r ∈ N. and Therefore G 6= ∅. If the conditions of the Theorem are satisfied then to each g = (g1, g2, ...) ∈ G, we put into correspondence a sequence H = H(g) = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (3.37) and (3.52) hold. Then H ∈ X(H), as it was shown before the statement of the Theorem. The sequence H generates a solution of the extended two-dimensional moment problem and of the two-dimensional moment problem, see considerations after the proof of Theorem 3.2. It remains to show that all solutions of the two-dimensional moment problem can be obtained in this way. Since elements of X(H) generate all solutions of the two- dimensional moment problem (see considerations after the proof of Theorem 3.2), it remains to prove that {H(g) : g ∈ G} = X(H). (3.57) It was shown that X1 ⊆ X(H). Denote the set on the left-hand side by X1. On the other hand, choose an arbitrary eH = {ehm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H). Repeating the construction at the beginning of this proof we obtain that bh ∈\r∈N (3.53) (3.54) (3.55) (3.56) (3.58) (3.59) (cid:3) Observe that eh := (ehw(1),ehw(2),ehw(3), ...) ∈ G. H(eh) =eH. Therefore X(H) ⊆ X1 and relation (3.57) holds. Remark 3.4. The truncated two-dimensional moment problem can be considered in a similar manner. Moreover, the set of indices of the known elements hm,n will be finite in this case and therefore equations in the r-th step of the algorithm will form finite systems of linear equations. Thus, the r-th step could be easily performed using computer. Remark 3.5. Consider the following system of r linear equations: A1 x1 x2 ...  = f 1 1 f 1 2 ... f 1 r  , (3.60) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 19 i,j)1≤i≤r;j∈N is a given complex numerical matrix, f 1 If A1 = 0 then the algorithm stops. Conditions of solvability and the where A1 = (a1 i , 1 ≤ i ≤ r, are given complex numbers, and xj, j ∈ N, are unknown complex numbers; r ∈ N. The Gauss algorithm allows to solve this system explicitly. Let us briefly describe this. Step 1. set of solutions are obvious in this case. If A1 6= 0, let m1-th column of A1 be the first non-zero column of A1. Inter- changing equations we set the non-zero element of this column in the first row and divide this equation by this element. Then we exclude xm1 from the rest of equations. We get the following system: 1,m1+1xm1+1 + a2 1,m1+2xm1+2 + ... = f 2 1 , xm1 + a2 (3.61) A2 xm1+1 xm1+2 ...  = f 2 2 f 2 3 ... f 2 r  , 1,j, j ≥ m1 + 1, are given complex numbers. where A2 is a given complex numerical matrix with r − 1 rows, f 2 and a2 i , 1 ≤ i ≤ r, Then we repeat the same construction for the linear system (3.62). After a finite number of steps the algorithm stops. Then we exclude xmt , xmt−1, ..., xm1 from the previous equations (t ≤ r). Thus, the numbers xj: j 6= m1, m2, ..., mt can be chosen arbitrary such that the corresponding series in (3.60) converge, and xm1, xm2 , ..., xmt are defined uniquely. If t < r, we additionally have the solvability conditions which follow from (3.62) in the last step. Observe that if we have an infinite number of equations in (3.60), we can choose an increasing number of equations and then construct the intersection of the solution sets. (3.62) (3.63) A modified algorithm. Notice that in Theorem 3.3 the correspondence between the parameters set G and solutions of the two-dimensional moment problem is not necessarily bijective. The algorithm may be modified to make this correspondence one-to-one. The following modified algorithm is more complicated. If we only need to check the solvability or the bijection is not necessary for our purposes, we can use the original algorithm. First, condition 2) of Theorem 3.2 may be replaced by the following more precise condition: 2) Set Λ := {r ∈ N : (hw(r), er) > 0}. Then hw(r) ∈ H0 ⊕ Mj∈Λ: 1≤j≤r Hj! , and (hw(r), er) ≥ 0, r ∈ N. The necessity of this condition was shown before Theorem 3.2, while condition 2) was not used in the proof of the sufficiency of Theorem 3.2. As before, we denote a set of sequences {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H satisfying conditions 1)-4) of Theorem 3.2 by X = X(H). Observe that the modified X 20 S. ZAGORODNYUK is a subset of the original one. The same arguments show that the new X(H) generates all solutions of the two-dimensional moment problem, as well. Step 0 and Step 1 of the algorithm will be the same as before. We set Hk := H0 ⊕ M1≤j≤k Hj! , Hk + := {h ∈ Hk : (h, ek) > 0}, k ∈ N. (3.64) In the r-th step we shall proceed in the following way (r ≥ 2). Choose an arbitrary (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1. Observe that by the con- struction in the (r − 1)-th step we have hw(j) ∈ Hk−1 or hw(j) ∈ Hk +, 1 ≤ k ≤ r − 1. Set Sr := {~s = (s1, s2, ..., sr−1) : sj = 1 or sj = 0, 1 ≤ j ≤ r − 1}; and ~s := {(h1, h2, ..., hr−1) : hj ∈ Hj−1 if sj = 0; hj ∈ Hj Hr + if sj = 1; 1 ≤ j ≤ r−1}, (3.67) Observe that Sr is a finite set of 2r−1 binary numbers. By (3.65) we obtain that ~s ∈ Sr. (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1 ∩ Hr ~s, for some ~s ∈ Sr. (3.65) (3.66) (3.68) (3.69) (3.70) (3.71) Set Notice that and Γr−1,~s := Gr−1 ∩ Hr ~s, Γr−1,~s1 ∩ Γr−1,~s2 = ∅, ~s ∈ Sr. ~s1, ~s2 ∈ Sr, Γr−1,~s. Gr−1 = [~s∈Sr αr;ngn + X1≤j≤r−1: sj=1 ∞Xn=1 Choose an arbitrary ~s ∈ Sr. We seek for hw(r) in the following form: hw(r) = βr;jej + βr;rer, (3.72) with some complex coefficients αr;n, βr;j: βr;r ≥ 0. Conditions (2.12)-(2.15) (with h instead of x) and (3.31)-(3.36) which include hw(r) and the already constructed hm,k,l;n,r,t are equivalent to a set Lr(~s) of linear equations with respect to αr;n, n ∈ N, βr;j, 1 ≤ j ≤ r − 1 : sj = 1, and dr = khw(r)k2 Notice that these linear equations depend on βr;j only by dr. Denote the set of solutions of these equations by H, and depending on parameters (hw(1), hw(2), ..., hw(r−1)) ∈ Γr−1,~s . Sr(~s) =(cid:8)(αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1 : sj = 1; dr; hw(1), hw(2), ..., hw(r−1)) : and equations from Lr(~s) with parameters (hw(1), hw(2), ..., hw(r−1)), are satisfied(cid:9) . (hw(1), hw(2), ..., hw(r−1)) ∈ Γr−1;~s, (3.73) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 21 Set ∞Xn=1 Finally, we set . (3.74) bSr(~s) =(cid:8)(αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1 : sj = 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ Sr(~s) : αr;n2 + X1≤j≤r−1: sj=1 Gr(~s) =(cid:8)(cid:0)hw(1), hw(2), ..., hw(r−1), βr;jej +dr − βr;j2 ≤ dr, dr ≤ Mr ∞Xn=1 αr;n2 − X1≤j≤r−1: sj=1 er : βr;j2 ∞Xn=1 αr;ngn + X1≤j≤r−1: sj =1 (αr;n, n ∈ N; βr;j, 1 ≤ j ≤ r − 1 : sj = 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ bSr(~s)o . 1 2 (3.75) The case Gr(~s) = ∅ is not excluded. We set Gr := [~s∈Sr Gr(~s). (3.76) The final step is the same as for the original algorithm. Thus, we obtain the set G. To each g = (g1, g2, ...) ∈ G, we put into correspondence a sequence H = H(g) = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (3.37),(3.50) hold. We state that H(g) ∈ X(H). Observe that the set Gr in (3.76) is a subset of Gr for the original algorithm. Therefore the set G is a subset of G for the original algorithm. Thus, H(g) belongs to the old X(H). To show that it belongs to the modified X(H) it remains to verify (3.63). Observe that (hw(1), hw(2), ..., hw(r)) ∈ Gr(~s), for some ~s ∈ Sr. The condition (hw(r), er) ≥ 0 follows from the construction of hw(r) in the r-th step. Set By (3.72) we get {1 ≤ j ≤ r − 1 : sj = 1}, Λ0(r) :=(cid:26) {1 ≤ j ≤ r − 1 : sj = 1} ∪ r, hw(r) ∈ H0 ⊕ Mj∈Λ0(r) Hj . if (hw(r), er) > 0 if (hw(r), er) = 0 . (3.77) (3.78) Thus, it remains to verify that Λ0(r) = {j ∈ Λ : 1 ≤ j ≤ r} =: Λ(r). But for 1 ≤ j ≤ r − 1, conditions sj = 1 and (hw(j), ej) > 0 are equivalent. Consequently, we obtain H(g) ∈ X(H). Theorem 3.3 remains true if we replace words "according to the algorithm" by the words "according to the modified algorithm", and add the following sentence: "The correspondence between elements of G and solutions of the two-dimensional moment problem is bijective". Let us check this last assertion (and the rest of the proof is similar). 22 S. ZAGORODNYUK The correspondence between G and X(H) is obviously bijective. Let Hj = {hj m,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H), j = 1, 2, be different: H1 6= H2. They pro- duce solutions µ1 and µ2 of the two-dimensional moment problem, respectively. Suppose that µ1 = µ2 = µ. Recall that µj is constructed as a solution of the corre- sponding extended two-dimensional moment problem with moments uj m,k,l;n,r,t = ϕj(m, k, l; n, r, t), j = 1, 2 (see the proof of Theorem 3.2). Here ϕj(m, k, l; n, r, t) is from Condition 4) for Hj, j = 1, 2. Therefore ϕ1(m, k, l; n, r, t) = ϕ2(m, k, l; n, r, t). By condition 4) of Theorem 3.2 this means that (h1 m,k,l;n,r,t, h1 m′,k′,l′;n′,r′,t′)H = (h2 for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Choose the minimal r, r ∈ N, such that w(r) 6= h2 h1 w(r). By (3.37),(3.79) we obtain m,k,l;n,r,t, h2 m′,k′,l′;n′,r′,t′)H, H0h1 P H By condition (3.63) we may write w(r) = P H H0h2 w(r) =: h0. (3.79) (3.80) (3.81) h1 w(r) = h0 + h2 w(r) = h0 + X1≤j≤r−1: (h1 X1≤j≤r−1: (h2 w(j) w(j) ,ej)>0 γ1 r;jej + γ1 r;rer, γ1 r;j ∈ C, γ1 r;r ≥ 0; ,ej)>0 γ2 r;jej + γ2 r;rer, r;j ∈ C, γ2 γ2 r;r ≥ 0. (3.82) w(j) =: hw(j), 1 ≤ j ≤ r − 1, we get w(j), ej) > 0} = {j : 1 ≤ j ≤ r − 1, (h2 w(j), ej) > 0} =:bΛ. (3.83) γa r;jej + γa r;rer, γa r;j ∈ C, γa r;r ≥ 0, a = 1, 2. (3.84) Suppose that there exists j ∈bΛ such that γ1 index j. Since j0 ∈bΛ, we get ζj0 := (hw(j0), ej0) = (ha w(j0), ej0) > 0, a = 1, 2, r;j 6= γ2 r;j. Let j0 be the minimal such w(j) = h2 Since h1 {j : 1 ≤ j ≤ r − 1, (h1 Therefore w(r) = h0 +Xj∈ bΛ ha and Then and hw(j0) = ζj0ej0 + uj0−1, uj0−1 ∈ Hj0−1. ej0 = 1 ζj0 (hw(j0) − uj0−1); (3.85) (3.86) γa r;j0 = (ha (ha w(r), ej0) = 1 ζj0 w(r), hw(j0)) − (ha w(r), hw(j0) − uj0−1) 1 ζj0 w(r), uj0−1) (ha = 1 ζj0 r;r ≥ 0, a = 1, 2. (3.88) ON THE TWO-DIMENSIONAL MOMENT PROBLEM 23 = 1 ζj0 (ha w(r), hw(j0)) − 1 ζj0 (h0 + Xj∈ bΛ:j<j0 γa r;jej, uj0−1). (3.87) By (3.79) and our assumption about j0 we obtain γ1 means that γ1 r;j = γ2 r;j0 = γ2 r;j0. This contradiction Observe that ha r;rer, r;j, ∀j ∈bΛ. Therefore w(r) =bh + γa bh ∈ Hr−1, γa w(r)k2 = kbhk2 + γa r;r2, r;r = γ2 kha a = 1, 2. By (3.79) we conclude that γ1 w(r). We obtained a contradiction with (3.80). The proof of the last assertion for the modified Theorem 3.3 is complete. r;r. Therefore h1 w(r) = h2 4. On a connection with the complex moment problem. In this section we shall analyze the complex moment problem: to find a non- negative Borel measure σ in the complex plane such that ZC zmzndσ = am,n, m, n ∈ Z+, (4.1) where {am,n}m,n∈Z+ is a prescribed sequence of complex numbers. Recall the canonical identification of C with R2: z = x1 + x2i, x1 = Re z, x2 = Im z, z ∈ C, (x1, x2) ∈ R2. (4.2) Let σ be a solution of the complex moment problem (4.1). The measure σ, viewed as a measure in R2, we shall denote by µσ. Then where C n k = n! k!(n−k)!. Then k C n j ak+j,m−k+n−j, (4.3) 2 (cid:19)m(cid:18) z − z 2i (cid:19)n dσ zk+jzm−k+n−jdσ 1 = 1 = C m k C n xm 1 xn 2m(2i)n 2m(2i)n sm,n :=ZR2 mXk=0 2 dµσ =ZC(cid:18) z + z j (−1)n−jZC nXj=0 nXj=0 mXk=0 (−1)n−jC m zmzndσ =ZR2 am,n =ZC nXl=0 mXr=0 nXl=0 mXr=0 r C n r C n C m C m = = (x1 + ix2)m(x1 − ix2)ndµσ xr+l 1 (ix2)m−r+n−ldµσ l (−1)n−lZR2 l (−1)n−lim−r+n−lsr+l,m−r+n−l; C m r C n l (−1)n−lim−r+n−lsr+l,m−r+n−l, m, n ∈ Z+, 24 S. ZAGORODNYUK and therefore where am,n = mXr=0 nXl=0 sm,n = 1 2m(2i)n mXk=0 nXj=0 (−1)n−jC m k C n j ak+j,m−k+n−j, m, n ∈ Z+. Since µσ is a solution of the two-dimensional moment problem, then conditions of Theorem 3.3 hold. Theorem 4.1. Let the complex moment problem (4.1) be given. This problem has a solution if an only if conditions of Theorem 3.3 and (4.4) with sm,n defined by (4.5) hold. Proof. It remains to prove the sufficiency. Suppose that for the complex moment problem (4.1) conditions of Theorem 3.3 and (4.4) with sm,n defined by (4.5) hold. By Theorem 3.3 we obtain that there exists a solution µ of the two-dimensional moment problem with moments sm,n. The measure µ, viewed as a measure in C, we shall denote by σµ. Then (4.4) (4.5) (cid:3) (4.6) = ZC zmzndσµ =ZR2 nXl=0 mXr=0 mXr=0 nXl=0 r C n r C n C m C m = (x1 + ix2)m(x1 − ix2)ndµ xr+l 1 (ix2)m−r+n−ldµ l (−1)n−lZR2 l (−1)n−lim−r+n−lsr+l,m−r+n−l = am,n, where the last equality follows from (4.4). Theorem 4.2. Let the complex moment problem (4.1) be given and conditions of Theorem 3.3 and (4.4) with sm,n defined by (4.5) hold. Let Ψ be a set of all solutions of the complex moment problem (4.1) and Φ be a set of all solutions of the two-dimensional moment problem (1.1) with sm,n defined by (4.5). Then Ψ = {σµ : µ ∈ Φ}. Therefore all solutions of the moment problem (4.1) are described by Theorem 3.3. Proof. The proof is straightforward. (cid:3) Of course, this Theorem holds for the modified version of Theorem 3.3, as well. References 1. N.I. Akhiezer, Classical Moment Problem, Fizmatlit., Moscow, 1961. (in Russian). 2. N.I. Akhiezer, I.M. Glazman, Theory of Linear Operators in a Hilbert Space, Gos. izdat. teh.-teor. lit., Moscow, Leningrad, 1950. (in Russian). 3. Ju.M. Berezanskii, Expansions in Eigenfunctions of Selfadjoint Operators, Amer. Math. Soc., Providence, RI, 1968. (Russian edition: Naukova Dumka, Kiev, 1965). ON THE TWO-DIMENSIONAL MOMENT PROBLEM 25 4. Yu.M. Berezansky, Spectral theory of the infinite block Jacobi type normal matrices, or- thogonal polynomials on the complex domain, and the complex moment problem, Operator Theory: Advances and Applications 191 (2009), 37 -- 50. 5. B. Fuglede, The multidimensional moment problem, Expo. Math. 1 (1983), no. 4, 47 -- 65. 6. M. Putinar, F.-H. Vasilescu, Solving moment problems by dimensional extension, Ann. Math. 149 (1999), 1087 -- 1107. 7. J.A. Shohat, J.D. Tamarkin, The Problem of Moments, Amer. Math. Soc., New York City, 1943. 8. J. Stochel, F.H. Szafraniec, The complex moment problem and subnormality: a polar de- composition approach, J. Funct. Anal. 159 (1998), 432 -- 491. 9. S.M. Zagorodnyuk, Positive definite kernels satisfying difference equations, Methods Funct. Anal. Topol. 16 (2010), 83 -- 100.
1102.3850
1
1102
2011-02-18T15:04:00
$M$-structures in vector-valued polynomial spaces
[ "math.FA" ]
This paper is concerned with the study of $M$-structures in spaces of polynomials. More precisely, we discuss for $E$ and $F$ Banach spaces, whether the class of weakly continuous on bounded sets $n$-homogeneous polynomials, $\mathcal P_w(^n E, F)$, is an $M$-ideal in the space of continuous $n$-homogeneous polynomials $\mathcal P(^n E, F)$. We show that there is some hope for this to happen only for a finite range of values of $n$. We establish sufficient conditions under which the problem has positive and negative answers and use the obtained results to study the particular cases when $E=\ell_p$ and $F=\ell_q$ or $F$ is a Lorentz sequence space $d(w,q)$. We extend to our setting the notion of property $(M)$ introduced by Kalton which allows us to lift $M$-structures from the linear to the vector-valued polynomial context. Also, when $\mathcal P_w(^n E, F)$ is an $M$-ideal in $\mathcal P(^n E, F)$ we prove a Bishop-Phelps type result for vector-valued polynomials and relate norm-attaining polynomials with farthest points and remotal sets.
math.FA
math
M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES VERÓNICA DIMANT AND SILVIA LASSALLE Abstract. This paper is concerned with the study of M-structures in spaces of polynomials. More precisely, we discuss for E and F Banach spaces, whether the class of weakly continuous on bounded sets n-homogeneous polynomials, Pw(nE, F ), is an M-ideal in the space of continuous n-homogeneous polynomials P(nE, F ). We show that there is some hope for this to happen only for a finite range of values of n. We establish sufficient conditions under which the problem has positive and negative answers and use the obtained results to study the particular cases when E = ℓp and F = ℓq or F is a Lorentz sequence space d(w, q). We extend to our setting the notion of property (M ) introduced by Kalton which allows us to lift M-structures from the linear to the vector-valued polynomial context. Also, when Pw(nE, F ) is an M-ideal in P(nE, F ) we prove a Bishop-Phelps type result for vector-valued polynomials and relate norm-attaining polynomials with farthest points and remotal sets. 1 1 0 2 b e F 8 1 ] . A F h t a m [ 1 v 0 5 8 3 . 2 0 1 1 : v i X r a Introduction M -ideals emerged in the geometric theory of Banach spaces as a generalization, to the Banach space setting, of the closed two-sided ideals in a C ∗-algebra. This notion, introduced by Alfsen and Effros in their seminal article [4] of 1972, leads us to a better understanding of the isometric structure of a Banach space in terms of geometric and analytic properties of the closed unit ball of the dual space. To be more precise, a closed subspace J of a Banach space X is an M -ideal in X, if its annihilator, J ⊥, is the kernel of a projection P on the dual space X ∗ such that kx∗k = kP (x∗)k + kx∗ − P (x∗)k, for all x∗ ∈ X ∗. When J is an M -ideal in X, the canonical complement of J ⊥ in X ∗ is (isometrically) identified with J ∗. Then, we may write X ∗ = J ⊥ ⊕1 J ∗, which in some sense tells us that there is a maximum norm structure underlying the geometry of the unit ball of X and this structure is closely related to J. If it is possible to decompose X as J ⊕∞ eJ, for some closed subspace eJ of X, we say that J is an M -summand of X. Clearly, M -summands are M -ideals, but there exist subtle differences. For instance, c0 is an M -ideal in ℓ∞ and it is not an M -summand. Since M -ideals appeared, they have been intensively studied. A comprehensive exposition of the main developments in this subject can be found in the outstanding book by Hardmand, Werner and Werner [24]. The Gelfand-Naimark theorem states that any arbitrary C ∗-algebra is isometrically ∗-isomorphic to a C ∗-algebra of bounded operators on a Hilbert space. Here the only norm closed two-sided ∗-ideal is the subspace of compact operators. Then, it is natural to investigate under which conditions the closed subspace J of compact operators between Banach spaces E and F , J = K(E, F ), results an M - ideal in X = L(E, F ), the space of linear and bounded operators, endowed with the supremum norm. During the last thirty years a number of papers have been devoted to this question (see, for example [24, 25, 26, 27, 28, 30, 31]), where the case E = F is of special interest. 2010 Mathematics Subject Classification. 47H60,46B04,47L22,46B20. Key words and phrases. M-ideals, homogeneous polynomials, weakly continuous on bounded sets polynomials. Partially supported by conicet-pip 11220090100624. The second author was also partially supported by UBACyT X218 and UBACyT X038. 1 2 VERÓNICA DIMANT AND SILVIA LASSALLE In this paper we focus our study in determining the presence of an M -structure in the space of contin- uous n-homogeneous polynomials between Banach spaces E and F , denoted by P(nE, F ). Here the lack of linearity and, more specifically, the degree of homogeneity will play a crucial role. In the polynomial setting, the space of compact operators is usually replaced by the space of homogeneous polynomials which are weakly continuous on bounded sets, denoted by Pw(nE, F ). Recall that a polynomial P ∈ P(nE, F ) is compact if maps the unit ball of E into a relatively compact set in F and that P is in Pw(nE, F ) if maps bounded weak convergent nets into convergent nets. For linear operators both properties, to be compact and to be weakly continuous on bounded sets, produce the same subspace. For n-homogeneous polynomials with n > 1, that coincidence is no longer true. Although any polynomial in Pw(nE, F ) is compact (as it can be derived from results in [9] and [8]), the reverse inclusion fails. This is due to the fact that continuous polynomials are not, in general, weak-to-weak continuous. Then, every scalar-valued continuous polynomial is compact but it is not necessarily weakly continuous on bounded sets, as the k, for all x = (xk)k ∈ ℓ2, shows. With this in mind, our main purpose is to discuss whether Pw(nE, F ) is an M -ideal in P(nE, F ). In [16], the first author studied the analogous question when F is the scalar field. We will see that the vector-valued case is not a mere generalization of the scalar-valued case. standard example P (x) =Pk x2 The problem of stating if Pw(nE, F ) is a proper subspace of P(nE, F ) is nontrivial at all. However, when this is not the situation our question is trivially answered. We refer the reader to [3, 13, 22, 23], where the equality Pw(nE, F ) = P(nE, F ) is studied. As it happens for n-homogeneous polynomials in the scalar-valued case, the value of n for which Pw(nE, F ) has the chance to be a nontrivial M -ideal in P(nE, F ) cannot be chosen arbitrarily. Thus, our firsts efforts are focused to discuss this matter. In order to do so, following [24] and [16], we define the essential norm of a vector-valued polynomial P as the distance from P to the space Pw(nE, F ). Also we describe the extreme points of the ball of the dual space of Pw(nE, F ). Then, combining this with properties of the essential norm we obtain the range within we may expect to find an M -structure. When Pw(nE, F ) is an M -ideal in P(nE, F ), the essential norm allows us to obtain a Bishop-Phelps type theorem. We use this result to study the existence of farthest points and densely remotal sets. These concepts are related to geometric properties such us the existence of exposed points, the Mazur intersection property and norm attaining functions, see [11, 20]. These results appear in Section 1. Section 2 is dedicated to give sufficient conditions on E and F so that Pw(nE, F ) is an M -ideal in P(nE, F ). The main requirement stays around the concept of shrinking approximations of the identity. When F is an M∞-space, without any further assumption on the space E, we prove that Pw(nE, F ) is a nontrivial M -ideal in P(nE, F ) for all but one possible value of n in the range of interest. For the remaining value of n, we obtain the result when E satisfies some additional conditions, see Propositions 2.9 and 2.11. In Section 3, we focus our attention on classical sequence spaces E and F , for E = ℓp (1 ≤ p < ∞) and F = ℓq or F = d(w, q) a Lorentz sequence space, (1 ≤ q < ∞). The questions of whether K(ℓp, ℓq) is an M -ideal in L(ℓp, ℓq) and K(ℓp, d(w, q)) is an M -ideal in L(ℓp, d(w, q)) were previously addressed in [24] and [30]. In [16], it was studied when Pw(nℓp) is an M -ideal in P(nℓp). We analyze here when Pw(nℓp, ℓq) is an M -ideal in P(nℓp, ℓq) and when Pw(nℓp, d(w, q)) is an M -ideal in P(nℓp, d(w, q)). Giving conditions on n, p, q and w we solve the problem for all the possible situations. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 3 In the last section we study the property (M ), introduced by Kalton in [26] for Banach spaces, developed later for operators by Kalton and Werner in [27] and finally generalized to the scalar-valued polynomial setting in [16]. Here, we present a natural extension to the vector-valued polynomial setting of the notions mentioned before and establish the connection this property has with our main problem. We apply the results obtained to give examples of M -ideals in vector-valued polynomial spaces defined on Bergman spaces. Before proceeding, we fix some notation and give basic definitions. Every time we write X, E or F we will be considering Banach spaces over the real or complex field, K. The closed unit ball of X will be noted by BX and the unit sphere by SX . Also, if x ∈ X and r > 0, B(x, r) will stand for the closed ball in X with center at x and radius r. As usual, X ∗ and X ∗∗ will be the notations for the dual and bidual of X, respectively. The space of linear bounded operators from E to E will be noted by L(E) and its subspace of compact mappings will be noted by K(E). A function P : E → F is said to be an n-homogeneous polynomial if there exists a (unique) symmetric n-linear form ∨ P : E × · · · × E } → F such that P (x) = ∨ P (x, . . . , x), n {z for all x ∈ E. For scalar-valued mappings we will write P(nE) instead of P(nE, F ) to denote the space of all continuous n-homogeneous polynomials from E to K. The space P(nE, F ) endowed with the supremum norm kP k = sup{kP (x)kF : x ∈ BE}, is a Banach space. We may write kP (x)k instead of kP (x)kF unless we prefer to emphasize the space where the norm is taken. Every polynomial P ∈ P(nE, F ) has two natural mappings associated: the linear adjoint or transpose P ∗ ∈ L(F ∗, P(nE)) which is given by (P ∗(y∗))(x) = y∗(P (x)), for every x ∈ E and y∗ ∈ F ∗, and the polynomial P ∈ P(nE∗∗, F ∗∗), the canonical extension of P from E to E∗∗ obtained by weak-star density, known as the Aron-Berner extension of P [5]. For each z ∈ E∗∗, ez will refer to the application given by ez(P ) = P (z); for x ∈ E, ex denotes the evaluation map. Besides the subspace of weakly continuous n-homogeneous polynomials on bounded sets which was already introduced, we will consider the following classes. The first one is the space of n-homogeneous polynomials that are weakly continuous on bounded sets at 0, which consists on those polynomials mapping bounded weakly null nets into null nets. This space will be denoted by Pw0(nE, F ). We also have the j ∈ E∗, yj ∈ F for all j = 1, . . . , N and N ∈ N. The space of finite type n-homogeneous polynomials will be denoted by Pf (nE, F ). Its closure (in the supremum norm) is the space of approximable n-homogeneous polynomials which will be noted by PA(nE, F ). When F is K we omit F and write Pw0(nE), Pf (nE) or PA(nE) for instance. subspace formed by polynomials of finite type, which are of the formPN j )n · yj, with x∗ j=1(x∗ Recall that if E does not contain a subspace isomorphic to ℓ1, then, for any Banach space F , Pw(nE, F ) coincides with the space of weakly sequentially continuous polynomials Pwsc(nE, F ) [8, Proposition 2.12]. 4 VERÓNICA DIMANT AND SILVIA LASSALLE The space of n-homogeneous polynomials that are weakly sequentially continuous at 0 will be denoted Pwsc0(nE, F ). We refer to [18, 29] for the necessary background on polynomials on Banach spaces. Related to the study of M -structures there are two relevant geometric properties that we will use repeatedly. The first one is a well-known characterization, called the 3-ball property, given by Alfsen and Effros in [4, Theorem A] to which the main part of their article is dedicated, see also [24, Theorem I.2.2 (iv)]: Theorem A. Suppose that J is a closed subspace of X. The following are equivalent: (i) J is an M -ideal. (ii) J satisfies the 3-ball property: for every x1, x2, x3 ∈ X and positive numbers r1, r2, r3 such that 3\j=1 it holds that B(xj, rj) 6= ∅ and B(xj, rj) ∩ J 6= ∅, j = 1, 2, 3, 3\j=1 B(xj, rj + ε) ∩ J 6= ∅ f or all ε > 0. (iii) J satisfies the (restricted) 3-ball property: for every y1, y2, y3 ∈ BJ , x ∈ BX and ε > 0, there exists y ∈ J satisfying kx + yj − yk ≤ 1 + ε, j = 1, 2, 3. Note that one of the benefits of having the 3-ball property is that we have a criterium to decide if a closed subspace of a Banach space X is an M -ideal in terms of an intersection of balls in X. Thus, there is no need to appeal to the dual space to determine the existence of an M -structure. The 2-ball property is not sufficient to this end, see [24]. When a closed subspace of X satisfies the 2-ball property we say that we are in presence of a semi M -ideal structure. The second property we referred, provides us with a nice description of the extreme points of the unit ball of X ∗ in terms of the sets of the extreme points of the unit balls of J ⊥ and J ∗, if J is an M -ideal in X, see [24, Lemma 1.5]. As usual Ext(BX ) denotes the set of extreme points of the unit ball of a Banach space X. Theorem B. Suppose that J is an M -ideal in X. Then, the extreme points of the unit ball of X ∗ satisfy Ext(BX ∗ ) = Ext(BJ ⊥) ∪ Ext(BJ ∗ ). Many authors investigated M -structures on Banach spaces. Hardmand, Werner and Werner summa- rized the main results on this topic in their monograph [24]. The reader will find out that it is a very clear and well-organized survey on M -ideals. Along this paper, we will recourse to the ideas and results in it. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 5 1. General results It is natural to begin our research with vector-valued polynomial versions of basic results stated for linear operators in [24, Propositions VI.4.2 and VI.4.3] and for scalar-valued polynomials in [16, Propositions 1.1 and 1.2]. We omit the proofs since they are straightforward. Proposition 1.1. (a) If Pw(nE, F ) is an M -summand in P(nE, F ), then Pw(nE, F ) = P(nE, F ). (b) If Pw(nE, F ) is an M -ideal in P(nE, F ) and E1 ⊂ E, F1 ⊂ F are 1-complemented subspaces, then Pw(nE1, F1) is an M -ideal in P(nE1, F1). (c) The class of Banach spaces E and F for which Pw(nE, F ) is an M -ideal in P(nE, F ) is closed with respect to the Banach-Mazur distance. The knowledge of the extreme points of the unit ball of a Banach space provides a crucial tool in the geometric study of the space. We borrow some ideas of [24] and [16] to examine the extreme points of the unit ball of the dual spaces: P(nE, F )∗ and Pw(nE, F )∗. Note that if J is a subspace of P(nE, F ) that contains Pf (nE, F ), then ex ⊗ y∗ ∈ J ∗ is a norm one element, for all x ∈ SE and y∗ ∈ SF ∗. Indeed, the application ex ⊗ y∗ belongs to BJ ∗ and since J contains all finite type n-homogeneous polynomials, it contains the elements of the form (x∗)n ·y, for every x∗ ∈ E∗ and y ∈ F , thus kex ⊗ y∗k = 1. Proposition 1.2. (a) If J is a subspace of P(nE, F ) that contains all finite type n-homogeneous polynomials, then where w∗ designates the topology σ(J ∗, J). ExtBJ ∗ ⊂(cid:8)ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗(cid:9)w∗ , (b) For the particular case J = Pw(nE, F ) we can be more precise: ExtBPw(nE,F )∗ ⊂ {ez ⊗ y∗ : z ∈ SE ∗∗, y∗ ∈ SF ∗(cid:9). Proof. (a) Through Hahn-Banach theorem and the comment made above, it easily follows that BJ ∗ = Γ(cid:8)ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗(cid:9)w∗ . Now, by Milman's theorem [21, Theorem 3.41] we derive the desired inclusion: (b) Suppose that J = Pw(nE, F ). Let us see that {ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗} w∗ ExtBJ ∗ ⊂(cid:8)ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗(cid:9)w∗ . BE ∗∗, y∗ ∈ BF ∗(cid:9). If Φ ∈ {ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗} in SF ∗ such that exα ⊗ y∗ α convergent to an element z in BE ∗∗ and {y∗ ⊂ {ez ⊗ y∗ : z ∈ α}α w∗ → Φ. Without loss of generality, we may assume that {xα}α is σ(E∗∗, E∗)- , then there exist nets {xα}α in SE and {y∗ w∗ α}α is σ(F ∗, F )-convergent to an element y∗ in BF ∗. Note that for any P ∈ Pw(nE, F ), its Aron-Berner extension P belongs to P(nE∗∗, F ) (see for instance [14, Proposition 2.5]) and the compacity of P implies that P is w∗-continuous on bounded sets. Then, we w∗ → ez ⊗ y∗ and therefore, have that y∗ Φ = ez ⊗ y∗. When Φ is a norm one element we have that both z and y are elements in the respective unit spheres SE ∗∗ and SF ∗. Now, the result follows. (cid:3) α(cid:0)P (xα)(cid:1) → y∗(cid:0)P (z)(cid:1), for every P ∈ Pw(nE, F ). Thus, exα ⊗ y∗ α 6 VERÓNICA DIMANT AND SILVIA LASSALLE In [16], the notion of the essential norm was extended from operators to scalar-valued polynomials and was used to determine that Pw(nE) may be a nontrivial M -ideal of P(nE) for at most only one value of n. For vector-valued polynomials, also through the essential norm, we obtain a finite range of possible values of n for which Pw(nE, F ) has the chance to be a nontrivial M -ideal of P(nE, F ). Recall that the essential norm of a linear operator T is the distance from T to the subspace of compact operators. When K(E, F ) is an M -ideal in L(E, F ), there is an explicit alternative formula to compute this essential norm [24, Proposition VI.4.7]. Now we proceed to discuss de degrees of homogeneity for which our problem might have a nontrivial solution. Definition 1.3. Let P be an n-homogeneous polynomial P ∈ P(nE, F ). The essential norm of P is defined by kP kes = d(P, Pw(nE, F )) = inf{kP − Qk : Q ∈ Pw(nE, F )}. In order to obtain a good description of the essential norm, we will make use of the transpose of a πs E → F the linearization of P , where polynomial. Note that if P ∈ P(nE, F ) and we denote by LP :Nn,s πs is the projective symmetric tensor norm; then P ∗ is the usual adjoint of LP . Lemma 1.4. If P ∈ Pw(nE, F ) then P ∗ belongs to L(F ∗, Pw(nE)) and it is w∗-continuous on bounded sets. Proof. If P ∈ Pw(nE, F ) then P is compact and P ∗ ∈ L(F ∗, Pw(nE)). By [10, Proposition 3.2], P ∗ is a compact operator. Since P ∗ = L∗ P it follows that LP is compact and its adjoint P ∗ is w∗-continuous. (cid:3) Now we can obtain an alternative formula for the essential norm in the case that there is an M -structure. Proposition 1.5. Suppose Pw(nE, F ) is an M -ideal in P(nE, F ). Then, for any P ∈ P(nE, F ), where kP kes = max{w(P ), w∗(P )}, w(P ) = sup(cid:8) lim sup kP (xα)k : kxαk = 1, xα w∗(P ) = sup(cid:8) lim sup kP ∗(y∗ αk = 1, y∗ α)k : ky∗ α w→ 0(cid:9) and → 0(cid:9). w∗ Proof. Let P ∈ P(nE, F ). For any Q ∈ Pw(nE, F ) and for any normalized weak-star null net {y∗ holds α}α, it The other inequality follows analogously. Thus, kP kes ≥ max{w(P ), w∗(P )}. Now suppose that Pw(nE, F ) is an M -ideal in P(nE, F ). Then we have ExtBP(nE,F )∗ = ExtBPw(nE,F )⊥ ∪ ExtBPw(nE,F )∗. The essential norm of P , kP kes, is the norm of the class of P in the quotient space P(nE, F )/Pw(nE, F ) and the dual of this quotient can be isometrically identified with Pw(nE, F )⊥. Then, there exists Φ ∈ Since, by Lemma 1.4,(cid:13)(cid:13)Q∗(y∗ kP − Qk = kP ∗ − Q∗k ≥(cid:13)(cid:13)(P ∗ − Q∗)(y∗ α)(cid:13)(cid:13) → 0 it follows that kP − Qk ≥ lim sup kP ∗(y∗ α)(cid:13)(cid:13) ≥(cid:13)(cid:13)P ∗(y∗ α)(cid:13)(cid:13) −(cid:13)(cid:13)Q∗(y∗ α)(cid:13)(cid:13). α)k and thus kP kes ≥ w∗(P ). M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 7 ExtBPw(nE,F )⊥ such that Φ(P ) = kP kes. So, Φ ∈ ExtBP(nE,F )∗ and, by Proposition 1.2 (a), Φ ∈ (cid:8)ex ⊗ y∗ : x ∈ SE, y∗ ∈ SF ∗(cid:9)w∗ . Chose nets {xα}α in SE and {y∗ α}α in SF ∗ such that exα ⊗ y∗ α w∗ → Φ, where w∗ means the topology σ(P(nE, F )∗, P(nE, F )). In passing to appropriate subnets, we can suppose that {xα}α is σ(E∗∗, E∗)- convergent to an element z in BE ∗∗ and {y∗ α}α is σ(F ∗, F )-convergent to an element y in BF ∗. For any x∗ ∈ E∗ and y ∈ F , the polynomial (x∗)n · y belongs to Pw(nE, F ). This gives 0 = Φ((x∗)n · y) = lim α x∗(xα)ny∗ α(y) = z(x∗)ny∗(y). So it should be z = 0 or y∗ = 0. In the first case, {xα}α is weakly null and kP kes = Φ(P ) = lim α y∗ α (P (xα)) ≤ lim sup kP (xα)k ≤ w(P ). In the second case, {y∗ α}α is weak-star null and it follows similarly that kP kes ≤ w∗(P ). (cid:3) As in the scalar-valued polynomial case this result enable us to narrow the possible values of n for which Pw(nE, F ) could be an M -ideal in P(nE, F ). To see this, first note that the arguments from [13] and [6] used in the comments before Remark 1.8 of [16] also work for vector-valued polynomials. Thus, we obtain: Remark 1.6. For Banach spaces E and F , either Pw(kE, F ) = Pw0(kE, F ) = P(kE, F ), for all k, or there exists n ∈ N such that: • Pw(kE, F ) = Pw0(kE, F ) = P(kE, F ), for all k < n. • Pw(nE, F ) = Pw0(nE, F ) $ P(nE, F ). • Pw(kE, F ) $ Pw0(kE, F ) ⊂ P(kE, F ), for all k > n. When this value of n does exist, we call it the critical degree of (E, F ) and denote n = cd(E, F ). For the case F = K we write cd(E) instead of cd(E, K). Therefore, if there exists a polynomial from E to F which is not weakly continuous on bounded sets, the critical degree is the minimum of all k such that Pw(kE, F ) 6= P(kE, F ). Observe that if a scalar-valued polynomial P ∈ P(nE) is not weakly continuous on bounded sets then, for any y ∈ F , y 6= 0, the polynomial x 7→ P (x)y belongs to P(nE, F ) and it is not weakly continuous on bounded sets. This says that, for any Banach space F , cd(E, F ) ≤ cd(E). Note also that cd(E, F ) could be much smaller than cd(E). For instance, cd(ℓp, c0) = 1 while cd(ℓp) is the integer number satisfying p ≤ cd(ℓp) < p + 1. Lemma 1.7. Let P ∈ P(nE, F ) be a compact polynomial. (a) If n < cd(E) then P is weakly continuous on bounded sets. (b) w∗(P ) = 0. (a) If n < cd(E), then every scalar-valued n-homogeneous polynomial on E is weakly continuous Proof. on bounded sets. Then, P is weak-to-weak continuous on bounded sets. So, for any bounded net {xα}α 8 VERÓNICA DIMANT AND SILVIA LASSALLE w→ x, we have P (xα) w→ P (x). Being P compact, the bounded net {P (xα)}α should in E such that xα have a convergent subnet. By a canonical argument we derive that P (xα)→P (x) and thus P is weakly continuous on bounded sets. (b) This is a consequence of the proof of Lemma 1.4. (cid:3) Proposition 1.8. Every polynomial in P(nE, F ) which is weakly continuous on bounded sets at 0 and compact is weakly continuous on bounded sets if and only if n ≤ cd(E). Proof. If n > cd(E), there exists a polynomial p ∈ Pw0(nE) \ Pw(nE). Then, for a fixed y ∈ F the polynomial P (x) = p(x)y is weakly continuous on bounded sets at 0 and compact but it is not weakly continuous on bounded sets. Reciprocally, let n ≤ cd(E) and let P ∈ P(nE, F ) be a polynomial weakly continuous on bounded sets at 0 and compact. We know from [10, Proposition 3.4] that, for 0 < k < n, any derivative dkP (x) is compact. Thus, by Lemma 1.7 (a), we obtain that dkP (x) is weakly continuous on bounded sets, for all 0 < k < n. This fact together with the hypothesis of P being weakly continuous on bounded sets at 0 implies that P is weakly continuous on bounded sets. (cid:3) By Lemma 1.7 (b), if P ∈ P(nE, F ) is weakly continuous on bounded sets at 0 and compact then w(P ) = w∗(P ) = 0. If, in addition, Pw(nE, F ) is an M -ideal in P(nE, F ), Proposition 1.5 states that kP kes = 0 and so P is weakly continuous on bounded sets. Thus, by Proposition 1.8, it should be n ≤ cd(E). Therefore, we have: Corollary 1.9. If Pw(nE, F ) is an M -ideal in P(nE, F ), then n ≤ cd(E). Clearly, if n < cd(E, F ), Pw(nE, F ) is a trivial M -ideal in P(nE, F ). So the problem proposed is worth being studied for polynomials of degree n, with cd(E, F ) ≤ n ≤ cd(E). The fact that Pw(nE, F ) is an M -ideal in P(nE, F ) has some incidence in the set of polynomials whose Aron-Berner extension attains the norm. As we have for scalar-valued polynomials [16, Proposition 1.10], the following version of [24, Proposition VI.4.8] is a Bishop-Phelps type result for vector-valued polynomials. The proof is omitted since it can be obtained as a slight modification of the proof given in [16]. Proposition 1.10. Let E and F be Banach spaces and suppose that Pw(nE, F ) is an M -ideal in P(nE, F ). (a) If P ∈ P(nE, F ) is such that its Aron-Berner extension P does not attain its norm at BE ∗∗, then kP k = kP kes. (b) The set of polynomials in P(nE, F ) whose Aron-Berner extension does not attain the norm is nowhere dense in P(nE, F ). We finish this section relating norm attaining polynomials with farthest points and remotal sets. The study of the existence of farthest points in a set of a Banach space can be traced to the articles of Asplund [11] and Edelstein [20]. This concept is related to several geometric properties of the space, like the existence of exposed points and the Mazur intersection property. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 9 Perhaps some definitions are in order. Let J be a subspace of a Banach space X. Fix x ∈ X, the farthest distance from x to the unit ball of J is given by ρ(x, BJ ) = sup{kx − yk : y ∈ BJ }. A point x ∈ X has a farthest point in BJ if there exists y ∈ BJ such that kx − yk = ρ(x, BJ ). The set of points in X having farthest points in BJ is denoted by R(BJ ). Then we have: R(BJ ) = {x ∈ X : ∃ y ∈ BJ such that kx − yk = ρ(x, BJ )} . It is said that BJ is densely remotal in X if R(BJ ) is dense in X and it is almost remotal in X if R(BJ ) contains a dense Gδ set. In [12], Bandyopadhyay, Lin and Rao studied dense remotality of the ball of K(E, F ) in the space L(E, F ). Adapting some of their ideas and applying the previous proposition, in Corollary 1.14, we obtain a result about almost remotality of BPw(nE,F ) in P(nE, F ). First, observe that for any P ∈ P(nE, F ) we have that ρ(cid:0)P, BPw(nE,F )(cid:1) = kP k + 1. Indeed, it is clear that ρ(cid:0)P, BPw(nE,F )(cid:1) ≤ kP k + 1, for every P ∈ P(nE, F ) and the equality is obvious for the polynomial P ≡ 0. For the reverse inequality, given P ∈ P(nE, F ), P 6≡ 0, and ε > 0, fix x ∈ SE and y∗ ∈ SF ∗ such that y∗(P (x)) > (1 − ε)kP k. Now, take y ∈ SF and x∗ ∈ SE ∗ satisfying y∗(y) > 1 − ε and x∗(x) = 1 and consider the polynomial Q = −(x∗)n · y ∈ BPw(nE,F ). Then, we have kP − Qk = kP + (x∗)n · yk ≥ y∗(P (x)) + x∗(x)ny∗(y) = y∗(P (x)) + y∗(y) > (1 − ε) (kP k + 1) , for all ε > 0, which proves the claim. The relation between norm attaining linear functions and the sets of operators which admit farthest points in the unit ball of the space of compact operators was studied in [12]. To simplify our statements let us introduce the following notations: N A (P(nE, F )) = {P ∈ P(nE, F ) : P attains its norm at BE}, AB − N A (P(nE, F )) = {P ∈ P(nE, F ) : P attains its norm at BE ∗∗}. Proposition 1.11. N A (P(nE, F )) ⊂ R(cid:0)BPw(nE,F )(cid:1). Proof. By the previous observation, it is plain that the polynomial P ≡ 0 belongs to R(cid:0)BPw(nE,F )(cid:1). Now, if P ∈ N A (P(nE, F )), P 6≡ 0, there exists x ∈ SE such that kP (x)k = kP k. Let x∗ ∈ SE ∗ satisfying x∗(x) = 1. Consider the polynomial Q = −(x∗)n · P (x) kP k ∈ BPw(nE,F ). So Q is a farthest point for P because kP − Qk =(cid:13)(cid:13)(cid:13)(cid:13)P + (x∗)n · P (x) kP k(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:13)(cid:13)(cid:13)(cid:13)P (x) + P (x) kP k(cid:13)(cid:13)(cid:13)(cid:13) = kP k + 1. (cid:3) 10 VERÓNICA DIMANT AND SILVIA LASSALLE In [15], Choi and Kim proved that if E has the Radon-Nykodým property, then the set of norm attaining polynomials of P(nE, F ) is dense in P(nE, F ). As a consequence of this result we obtain: Corollary 1.12. If E has the Radon-Nykodým property, then BPw(nE,F ) is densely remotal in P(nE, F ). When Pw(nE, F ) is an M -ideal in P(nE, F ), the set R(cid:0)BPw(nE,F )(cid:1) does not only contain the set of norm attaining polynomials but it is also contained in the set of all the polynomials whose Aron-Berner extension is norm attaining. Proposition 1.13. If Pw(nE, F ) is an M -ideal in P(nE, F ), then R(cid:0)BPw(nE,F )(cid:1) ⊂ AB−N A (P(nE, F )). Proof. Let P ∈ R(cid:0)BPw(nE,F )(cid:1). So, there exists Q ∈ BPw(nE,F ) such that kP − Qk = kP k + 1. Take Φ ∈ ExtBP(nE,F )∗ satisfying Being Pw(nE, F ) an M -ideal in P(nE, F ), we should have that Φ(P − Q) = kP − Qk = kP k + 1. Φ ∈ ExtBPw(nE,F )∗ or Φ ∈ ExtBPw(nE,F )⊥. If Φ ∈ ExtBPw(nE,F )⊥, we obtain that Φ(P − Q) = Φ(P ) and so Φ(P ) = kP k + 1, which is not possible. Hence, it should be Φ ∈ ExtBPw(nE,F )∗ and, by Proposition 1.2 (b), Φ = ez ⊗ y∗, for certain z ∈ SE ∗∗ and y∗ ∈ SF ∗. Therefore, kP k + 1 = Φ(P − Q) = y∗(P (z)) − y∗(Q(z)) ≤ kP k + kQk = kP k + 1. It follows that y∗(P (z)) = kP k and so kP (z)k = kP k, meaning that P ∈ AB − N A (P(nE, F )). (cid:3) As a consequence of Propositions 1.10, 1.11 and 1.13, we obtain: Corollary 1.14. If E is reflexive and Pw(nE, F ) is an M -ideal in P(nE, F ), then and thus, P(nE, F ) \ R(cid:0)BPw(nE,F )(cid:1) is nowhere dense. This implies that BPw(nE,F ) is almost remotal in P(nE, F ). R(cid:0)BPw(nE,F )(cid:1) = N A (P(nE, F )) , 2. Sufficient conditions In this section we present several sets of sufficient conditions which enable us to ensure that Pw(nE, F ) is an M -ideal in P(nE, F ). All of them involve bounded nets of compact operators on E. The following lemma and proposition are the vector-valued versions of [16, Lemma 2.1 and Proposition 2.2], the proofs of which are analogous to those in [16]. Lemma 2.1. Let E and F be Banach spaces and suppose that there exists a bounded net {Sα}α of linear operators from E to E satisfying S∗ α(x∗) → x∗, for all x∗ ∈ E∗. Then, for all P ∈ Pw(nE, F ), we have that kP − P ◦ Sαk → 0. Proposition 2.2. Let E and F be Banach spaces and let n = cd(E, F ). Suppose that there exists a bounded net {Kα}α of compact operators from E to E satisfying the following two conditions: • K ∗ α(x∗) → x∗, for all x∗ ∈ E∗. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 11 • For all ε > 0 and all α0 there exists α > α0 such that for every x ∈ E, kKα(x)kn + kx − Kα(x)kn ≤ (1 + ε)kxkn. Then, Pw(nE, F ) is an M -ideal in P(nE, F ). Remark 2.3. A Banach space E is an (Mp)-space (1 ≤ p ≤ ∞) if K(E ⊕p E) is an M -ideal in L(E ⊕p E). This concept was introduced by Oja and Werner in [31]. By [24, Theorem VI.5.3], if E is an (Mp)-space with p ≤ n, then there exists a bounded net {Kα}α of compact operators from E to E satisfying both conditions of Proposition 2.2. Recall that a Banach space E has a finite dimensional decomposition {Ej}j if each Ej is a finite dimensional subspace of E and every x ∈ E has a unique representation of the form x = xj, with xj ∈ Ej, for every j. ∞Xj=1 Associated to the decomposition there is a bounded sequence of projections {πm}m, given by πm(cid:16)P∞ Pm j=1 xj. The decomposition is called shrinking if π∗ m(x∗) → x∗, for all x∗ ∈ E∗. It is clear that in this case {πm}m is a bounded sequence of compact operators that satisfies the first item of the previous proposition. Thus, for spaces with shrinking finite dimensional decompositions we state the following simpler version of Proposition 2.2. j=1 xj(cid:17) = Corollary 2.4. Let E and F be Banach spaces and let n = cd(E, F ). Suppose that E has a shrinking finite dimensional decomposition with associate projections {πm}m such that: • For all ε > 0 and all m0 ∈ N there exists m > m0 such that for every x ∈ E, kπm(x)kn + kx − πm(x)kn ≤ (1 + ε)kxkn. Then, Pw(nE, F ) is an M -ideal in P(nE, F ). Example 2.5. Let E =Lℓp Ym, where Xm and Ym are finite dimensional spaces and 1 < p, q < ∞. From [23] we can derive that the critical degree is the integer number cd(E, F ) satisfying p q ≤ cd(E, F ) < p q + 1. The hypothesis of Corollary 2.4 are fulfilled if cd(E, F ) ≥ p. When this is the case we thus obtain that Pw(nE, F ) is an M -ideal in P(nE, F ). Xm and F =Lℓq The conditions in the following theorem were inspired by those of [24, Lemma VI.6.7]. They concern bounded nets of compact operators both in E and in F . As an example of this result (see Example 2.8 below) we can consider the values of n = cd(ℓp, ℓq) uncovered by Example 2.5. Theorem 2.6. Let E and F be Banach spaces and suppose that there exist bounded nets of compact operators {Kα}α ⊂ K(E) and {Lβ}β ⊂ K(F ) and numbers 1 < p, q < ∞ such that: α(x∗) → x∗, for all x∗ ∈ E∗ and Lβ(y) → y, for all y ∈ F . • K ∗ • For all ε > 0 and all α0 there exists α > α0 such that for every x ∈ E, kKα(x)kp + kx − Kα(x)kp ≤ (1 + ε)pkxkp. • For all ε > 0 and all β0 there exists β > β0 such that for every y1, y2 ∈ F , kLβ(y1) + (Id − Lβ)(y2)kq ≤ (1 + ε)q(cid:16)ky1kq + ky2kq(cid:17). 12 VERÓNICA DIMANT AND SILVIA LASSALLE Suppose also that n = cd(E, F ) satisfies that p ≤ nq and n < cd(E). Then, Pw(nE, F ) is an M -ideal in P(nE, F ). Proof. We prove that the 3-ball property holds. Let P1, P2, P3 ∈ BPw(nE,F ), Q ∈ BP(nE,F ) and ε > 0. Define P = Q − (Id − Lβ)Q(Id − Kα). We want to show that P is weakly continuous on bounded sets and kQ + Pj − P k ≤ 1 + ε, for j = 1, 2, 3, for some convenient choice of α and β. To see that P is weakly continuous on bounded sets, we write P = Q − Q(Id − Kα) + LβQ(Id − Kα). The proof of [16, Proposition 2.2] shows that Q − Q(Id − Kα) is weakly continuous on bounded sets at 0 and since n = cd(E, F ), we have that Q − Q(Id − Kα) belongs to Pw(nE, F ). Also, as LβQ(Id − Kα) is a compact polynomial and n < cd(E), Lemma 1.7 (a) says that it is in Pw(nE, F ). Now, to show the (1 + ε)-bound, consider the inequality kQ + Pj − P k ≤ kQ + LβPjKα − P k + kPj − LβPjKαk. On the one hand, we have: kPj − LβPjKαk ≤ kPj − PjKαk + kPj Kα − LβPjKαk ≤ kPj − PjKαk + kPj − LβPjkkKαkn. By Lemma 2.1, kPj − PjKαk → 0 with α. Also, since Lβ approximates the identity on compact sets and the Pj's are compact polynomials, we have that kPj − LβPjkkKαkn → 0 with β, for all α. Furthermore, we can find α and β such that: kQ + LβPjKα − P k = sup x∈BE k(Id − Lβ)Q(Id − Kα)(x) + LβPjKα(x)k ≤ sup x∈BE (1 + ε) (kQ(Id − Kα)(x)kq + kPjKα(x)kq) 1 q ≤ (1 + ε) sup x∈BE ≤ (1 + ε) sup x∈BE (k(Id − Kα)(x)knq + kKα(x)knq) 1 q (k(Id − Kα)(x)kp + kKα(x)kp) n p ≤ (1 + ε)(1 + ε)n = (1 + ε)n+1, and the result follows. (cid:3) Remark 2.7. If E is an (Mp)-space and F is an (Mq)-space the conditions about the nets of compact operators of the previous theorem are fulfilled. Example 2.8. Let E =Lℓp Xm and F =Lℓq Ym, where Xm and Ym are finite dimensional spaces and 1 < p, q < ∞. As we note in Example 2.5, cd(E, F ) is the integer such that p q + 1. Also we know that cd(E) is the integer satisfying p ≤ cd(E) < p + 1. Now we obtain a result for the cases uncovered by Example 2.5, since if n = cd(E, F ) < p all the hypothesis of the above theorem hold. So, Pw(nE, F ) is an M -ideal in P(nE, F ). q ≤ cd(E, F ) < p In particular, we derive from this example and Example 2.5, that for every 1 < p, q < ∞, if n = cd(ℓp, ℓq), then Pw(nℓp, ℓq) is an M -ideal in P(nℓp, ℓq). M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 13 In all the previous results (Proposition 2.2, Corollary 2.4 and Theorem 2.6) the M -structure is obtained only in the case n = cd(E, F ). On the other hand, by the comments after Corollary 1.9, for Pw(nE, F ) to be a nontrivial M -ideal in P(nE, F ), it is necessary that cd(E, F ) ≤ n ≤ cd(E). Let us show now some positive results for values of n greater than cd(E, F ). Proposition 2.9. Let E be a Banach space and F be an (M∞)-space. If n < cd(E), then Pw(nE, F ) is an M -ideal in P(nE, F ). Proof. Being F an (M∞)-space, by [24, Theorem VI.5.3], there exists a net {Lβ}β contained in the unit ball of K(F ) satisfying Lβ(y) → y for all y ∈ F such that for any ε > 0, there exists β0 with (1) kLβ(y1) + (Id − Lβ)(y2)k ≤(cid:16)1 + ε 2(cid:17) max{ky1k, ky2k}, for all β ≥ β0 and for any y1, y2 ∈ F . Let P1, P2, P3 ∈ BPw(nE,F ) and Q ∈ BP(nE,F ), we show that with P = LβQ, choosing β properly, the 3-ball property is satisfied. First, note that by Lemma 1.7 (a), P is weakly continuous on bounded sets. Also, kQ + Pj − P k ≤ 2 for β large kQ + LβPj − P k + kPj − LβPjk. Reasoning as in Theorem 2.6, we have that kPj − LβPjk < ε enough. Now, from (1) we obtain kQ + LβPj − P k = k(Id − Lβ)Q + LβPjk ≤(cid:16)1 + ε 2(cid:17) max{kQk, kPj k} =(cid:16)1 + ε 2(cid:17) , and the result follows. (cid:3) Remark 2.10. Let E be a Banach space such that cd(E) > 2 and let F be an infinite dimensional (M∞)- space. Then, for any degree n, with 1 ≤ n < cd(E), Pw(nE, F ) is a nontrivial M -ideal in P(nE, F ). This is a simple consequence of the above proposition and the fact that cd(E, F ) = 1. The next proposition somehow complements Proposition 2.9. It states that if F is an (M∞)-space, with an additional hypothesis on E, then Pw(nE, F ) is an M -ideal in P(nE, F ) also in the case n = cd(E). Proposition 2.11. Let F be an (M∞)-space and let E be a Banach space. If n = cd(E) and there exists a bounded net of compact operators {Kα}α ⊂ K(E) satisfying both conditions: α(x∗) → x∗, for all x∗ ∈ E∗. • K ∗ • For all ε > 0 and all α0 there exists α > α0 such that for every x ∈ E, kKα(x)kn + kx − Kα(x)kn ≤ (1 + ε)kxkn, then Pw(nE, F ) is an M -ideal in P(nE, F ). Proof. Let P1, P2, P3 ∈ BPw(nE,F ), Q ∈ BP(nE,F ) and ε > 0. We will find P ∈ Pw(nE, F ) such that the 3-ball property is satisfied. Reasoning as in Theorem 2.6, we find α and β so that kPj − LβPjKαk < ε 2 . Moreover, α and β may be chosen to satisfy at the same time kKα(x)kn + kx − Kα(x)kn ≤ (1 + ε)kxkn and kLβ(y1) + (Id − Lβ)(y2)k ≤ (1 + ε) max{ky1k, ky2k} for all y1, y2 ∈ F ; where {Lβ}β is a net in K(F ), associated to the (M∞)-space F , and ε is such that (1 + ε)2 ≤ 1 + ε 2 . Let P be the polynomial P = Lβ(Q − Q(Id − Kα)). As in the proof of [16, Proposition 2.2], we can see that P is weakly continuous on bounded sets at 0. Since n = cd(E) and P is compact, we may appeal to Proposition 1.8 to derive that P is weakly continuous on bounded sets. 14 VERÓNICA DIMANT AND SILVIA LASSALLE Also we have, kQ + Pj − P k ≤ kQ + LβPjKα − P k + kPj − LβPjKαk ≤ k(Id − Lβ)Q + Lβ(PjKα + Q(Id − Kα))k + ε 2 ≤ (1 + ε) sup x∈BE max{kQ(x)k, kPj Kα(x) + Q(x − Kα(x))k} + ε 2 . Now, the hypothesis on E gives us kPj Kα(x) + Q(x − Kα(x))k ≤ kKα(x)kn + kx − Kα(x)kn ≤ (1 + ε), for all x ∈ BE, and the result follows. (cid:3) Example 2.12. Let E = ℓp, with 1 < p < ∞, and let F be an (M∞)-space. As a consequence of the previous propositions, since p ≤ cd(ℓp), Pw(nℓp, F ) is an M -ideal in P(nℓp, F ) for all 1 ≤ n ≤ cd(ℓp). 3. Polynomials between classical sequence spaces This section is devoted to study whether Pw(nE, F ) is an M -ideal in P(nE, F ), for all the values of n between cd(E, F ) and cd(E), in the cases E = ℓp and F = ℓq or F the Lorentz sequence space F = d(w, q), 1 < p, q < ∞. Recall that given a non increasing sequence w = (wj)j of positive real numbers satisfying w ∈ c0 \ ℓ1, the Lorentz sequence space d(w, q) is the space of all sequences x = (xj)j ⊂ K, such that ∞Xj=1 sup σ wjxσ(j)q < ∞, (where σ varies on the set of permutations of N) endowed with the norm kxkd(w,q) = sup σ (cid:16) ∞Xj=1 wjxσ(j)q(cid:17) 1 q . We will consider weights w = (wj)j so that w1 = 1, which implies that the canonical vectors of d(w, q) form a basis of norm 1 elements. projections associated to the decomposition; that is πm(y) = Pm We begin our study with a result about polynomials from a general Banach space E to a Banach space F having a finite dimensional decomposition (FDD) {Fn}n. As usual, {πm}m denotes the sequence of j=1 yj, with yj ∈ Fj. Also, we denote by πm = Id − πm. When the FDD is unconditional with unconditional constant 1, we have that kπmk ≤ 1 and kπm + πkk ≤ 1, for all k ≥ m. In the sequel, we will use, without further mentioning, that for any Banach space E and any Q ∈ Pw(nE, F ), kπmQ − Qk → 0, or equivalently, kπmQk → 0, both claims can be derived from the fact that Q is compact. j=1 yj for all y =P∞ The following proposition gives conditions under which, if F is a Banach space with 1-unconditional FDD, Pw(nE, F ) is not a semi M -ideal in P(nE, F ). This is a polynomial generalization of [30, Proposition 2] and our proof is modeled on the proof given in that article. From this, it is obviously inferred that Pw(nE, F ) is not an M -ideal in P(nE, F ). Proposition 3.1. Let E and F be Banach spaces such that F has an unconditional FDD with un- conditional constant equal to 1 and associated projections {πm}m. Suppose that there exist polynomials P ∈ P(nE, F ) and Q ∈ Pw(nE, F ) and numbers δ > 0 and m0 ∈ N such that: • 0 < kQk ≤ kP k < δ, M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 15 • kπmP + Qk ≥ δ, for all m ≥ m0. Then, Pw(nE, F ) is not a semi M -ideal in P(nE, F ). Proof. Fix ε > 0 so that ε < δ−kP k 3 , for all m ≥ m0. Now, fix m ≥ m0 and consider the following two closed balls of radius kP k: B1 = B(πmP + Q, kP k) and B2 = B(πmP − Q, kP k). Note that πmP ∈ B1 ∩ B2, Q ∈ B1 ∩ Pw(nE, F ) and −Q ∈ B2 ∩ Pw(nE, F ). . Since kπmQk → 0, we may assume that kπmQk < ε 2 If Pw(kE, F ) is a semi M -ideal, then for any r > kP k, the intersection B(πmP + Q, r) ∩ B(πmP − Q, r) ∩ Pw(nE, F ) is non void. Take r = kP k+δ 2 − ε and suppose that there exists R ∈ B(πmP + Q, r) ∩ B(πmP − Q, r) ∩ Pw(nE, F ). Since kπkRk → 0, we may choose k ≥ m such that kπkRk < ε/3. To get a contradiction we estimate kπkP + Qk. Note that (2) 2kπkP + Qk ≤ kπkP + πmQ − πmRk + kπkP + πmQ + πmRk + 2kπmQk. From the equality (πm + πk)(πmP + Q − R) = πkP + πmQ − πmR + πkQ − πkR, we obtain: kπkP + πmQ − πmRk ≤ kπm + πkkkπmP + Q − Rk + kπkQk + kπkRk < r + 2ε 3 . Also, we have that kπkP + πmQ + πmRk = kπkP − πmQ − πmRk, since F has 1-unconditional finite dimensional decomposition. Proceeding as before, we obtain: kπkP − πmQ − πmRk ≤ kπm + πkkkπmP − Q − Rk + 2ε 3 < r + 2ε 3 . Finally, using (2), we have that 2δ ≤ 2kπkP + Qk < 2r + 2ε = kP k + δ < 2δ. Thus, we conclude that Pw(nE, F ) is not a semi M -ideal in P(nE, F ). (cid:3) Now we can complete the case E = ℓp and F = ℓq. Theorem 3.2. Let n = cd(ℓp, ℓq). (a) Pw(nℓp, ℓq) is an M -ideal in P(nℓp, ℓq). (b) Pw(kℓp, ℓq) is not a semi M -ideal in P(kℓp, ℓq), for all k > n. Proof. Statement (a) follows from Example 2.5 and Example 2.8. To prove statement (b) take k > n. We will construct polynomials P ∈ P(kℓp, ℓq) and Q ∈ Pw(kℓp, ℓq) satisfying: kP k = kQk and kπmP + Qk ≥ δ > kP k, for some δ > 0, where {πm}m is the sequence of projections associated to the canonical basis of ℓq and πm = Id − πm, for all m ∈ N. q , as shown in Example 2.5, so we may define the k-homogeneous )j≥2. To compute the norm of P , we look, for each x ∈ ℓp, at 1(x)(xk−1 j We have that k − 1 ≥ cd(ℓp, ℓq) ≥ p continuous polynomial P (x) = e∗ the inequality kP (x)kℓq = x1(cid:16) ∞Xj=2 xj(k−1)q(cid:17) 1 q ≤ x1(cid:16) ∞Xj=2 xjp(cid:17) k−1 p . 16 VERÓNICA DIMANT AND SILVIA LASSALLE k )k−1(cid:3) 1 Then, kP k ≤ max{abk−1 : ap + bp = 1, a, b ≥ 0} =(cid:2) 1 p e1 +(cid:0)1 − 1 k(cid:1) 1 k(cid:1) 1 x =(cid:0) 1 we obtain a norm one element where P attains the bound (cid:2) 1 Let Q ∈ Pw(kℓp, ℓq) be the polynomial Q(x) = kP ke∗ k (1 − 1 p e2, k (1 − 1 and x = ( 1 k ) 1 p e1 + (1 − 1 k ) 1 p em+2, then kxkℓp = 1 and k )k−1(cid:3) 1 p . p . Now, considering 1(x)ke1. It is clear that kP k = kQk. Take m ≥ 1, kπmP + Qk ≥ k(πmP + Q)(x)kℓq =(cid:13)(cid:13)(cid:13)( 1 Then, with δ =(cid:16)1 + ( 1 for. And the theorem is proved. p (cid:17) 1 k ) kq q 1 p (1 − 1 k ) k−1 p em+1 + kP k( 1 k ) k ) = kP k(cid:16)1 + ( 1 k ) kq q p (cid:17) 1 . k p e1(cid:13)(cid:13)(cid:13)ℓq > 1, which is independent of m, we obtain the inequality we were looking (cid:3) Now we focus our attention on spaces of polynomials from ℓp to d(w, q), 1 < p, q < ∞. We study whether Pw(kℓp, d(w, q)) is an M -ideal in P(kℓp, d(w, q)) for k ≥ cd(ℓp, d(w, q)). To this end we extend to the vector-valued case a couple of results of [17] about polynomials from spaces with finite dimensional decompositions. Lemma 3.3. Let E be a Banach space which has an unconditional FDD with associated projections {πm}m. For any fixed subsequence {mj}j of N, let σj = πmj − πmj−1 , for all j. Given P ∈ P(nE, F ), the application P (x) = P (σj(x)), f or all x ∈ E, ∞Xj=1 defines a continuous n-homogeneous polynomial from E to F . Proof. We first show that the series P∞ Proposition 1.3], there exists C > 0 such that j=1 P (σj(x)) is convergent for every x ∈ E. Indeed, by [17, MXj=N (cid:13)(cid:13)(cid:13) P (σj(x))(cid:13)(cid:13)(cid:13) ≤ sup y∗∈BF ∗ = sup y∗∈BF ∗ MXj=N MXj=N y∗ ◦ P (σj(x)) y∗ ◦ P (σj(πmM (x) − πmN −1 (x))) ≤ CkP kkπmM (x) − πmN −1(x)kn, which converges to 0 with M and N . Then, P (x) is well defined and k P k ≤ CkP k. (cid:3) Recall that whenever a Banach space E has a shrinking FDD, by [8], Pw(nE, F ) = Pwsc(nE, F ). This allows us to work with sequences instead of nets. Proposition 3.4. Let E be a Banach space with an unconditional FDD and let F be a Banach space. For any n ∈ N, the following are equivalent: (i) P(nE, F ) = Pwsc(nE, F ). (ii) P(nE, F ) = Pwsc0(nE, F ). M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 17 Proof. By means of the previous lemma, the scalar valued result given in [17, Corollary 1.7] (see also [13]) can be easily modified to obtain this vector valued version. (cid:3) In [30], Eve Oja studies when K(ℓp, d(w, q)) is an M -ideal in L(ℓp, d(w, q)). In Proposition 1 of that ar- ticle, she establishes a criterium to ensure that every continuous linear operator is compact. A polynomial version of this result can be stated as follows. Proposition 3.5. Let {ej}j and {fj}j be sequences in Banach spaces E and F , respectively, satisfying: • For any semi-normalized weakly null sequence {xm}m ⊂ E, there exists a subsequence {xmj }j and an operator T ∈ L(E) such that T (ej) = xmj , for all j. • For any semi-normalized weakly null sequence {ym}m ⊂ F , there exists a subsequence {ymj }j and an operator S ∈ L(F ) such that S(ymj ) = fj, for all j. • For any subsequence {ejl}l of {ej }j, there exists an operator R ∈ L(E) such that R(el) = ejl, for all l. Take n < cd(E) and suppose that it does not exist a polynomial P ∈ P(nE, F ) such that P (ej) = fj, for every j. Then, P(nE, F ) = Pwsc0(nE, F ). Proof. Suppose there exists P ∈ P(nE, F ) which is not in Pwsc0(nE, F ). Then, there exists a weakly null sequence (xm)m such that kP (xm)k > ε, for some ε > 0 and all m. As n < cd(E), (P (xm))m is weakly null. Now, we may find a subsequence (xmj )j and operators R, T ∈ L(E) and S ∈ L(F ) satisfying: ej T ◦R−→ xmj P−→ P (xmj ) S−→ fj, which is a contradiction since S ◦ P ◦ T ◦ R belongs to P(nE, F ). (cid:3) Remark 3.6. If the Banach space E has an unconditional basis {ej}j with coordinate functionals {e∗ j }j and {fj}j is a sequence in the Banach space F, we derive from Lemma 3.3 that the existence of a polynomial P ∈ P(nE, F ) such that P (ej) = fj, for all j, is equivalent to the existence of the polynomial eP ∈ P(nE, F ) given by eP (x) = the polynomial above as eP (x) = (xn j )j. ∞Xj=1(cid:0)e∗ j (x)(cid:1)n fj, for all x ∈ E. When E and F are Banach sequence spaces with canonical bases {ej}j and {fj}j respectively, we write Let 1 < p, q < ∞. To study whether Pw(nℓp, d(w, q)) is an M -ideal in P(nℓp, d(w, q)) for n ≥ cd(ℓp, d(w, q)), we need first to establish the value of the critical degree, cd(ℓp, d(w, q)). To this end and in view of the previous remark and proposition, the point is to determine the values of n, p and q such that the polynomial x 7→ (xn j )j, from ℓp to d(w, q), is well defined. For 1 ≤ r < ∞ we use the standard notation s = r∗ to denote de conjugate number of r: 1 r + 1 s = 1. Proposition 3.7. The polynomial P (x) = (xn following two conditions holds: j )j belongs to P(nℓp, d(w, q)) if and only if one of the (a) n ≥ p (b) n < p q . In this case, kP k = 1. q and w ∈ ℓs, for s = ( p 1 nq )∗. In this case, kP k = kwk q ℓs . 18 VERÓNICA DIMANT AND SILVIA LASSALLE Proof. Let (ej)j and (fj)j be the canonical bases of ℓp and d(w, q), respectively. Suppose that n ≥ p kwk∞ = 1, we have q , as kP (x)kd(w,q) = sup σ (cid:16) ∞Xj=1 wjxσ(j)nq(cid:17) 1 q ≤ kxkn ℓp. Then, P is a well defined polynomial with norm less than or equal to 1. Also, P (ej) = fj implies kP k = 1. Now, suppose that n < p q and w ∈ ℓs, with s = ( p nq )∗. Put W = kwkℓs, by Hölder inequality, we have kP (x)kd(w,q) = sup σ (cid:16) ∞Xj=1 wjxσ(j)nq(cid:17) 1 q ≤ W 1 q kxkn ℓp. Thus, kP k ≤ W 1 q and considering x = W − s p (w s p j )j ∈ Sℓp, we obtain that kP k = W 1 q = kwk 1 q ℓs . Finally, suppose that n < p q and w 6∈ ℓs. Then, there exists (bj)j ∈ ℓ p nq with b1 ≥ b2 ≥ b3 ≥ · · · ≥ 0 such that the seriesP∞ d(w, q). Now, the proof is complete. j=1 wjbj does not converge. Taking x ∈ ℓp, x = (b 1 nq j )j we have that P (x) = (b 1 q j )j 6∈ (cid:3) Proposition 3.8. Pw(nℓp, d(w, q)) = P(nℓp, d(w, q)) if and only if n < p q and w 6∈ ℓs, with s = ( p nq )∗. Proof. By the previous proposition, whenever n ≥ p P (x) = (xn j )j belongs to P(nℓp, d(w, q)) and fails to be weakly continuous on bounded sets. q or n < p q and w ∈ ℓs, s = ( p nq )∗, the polynomial For the converse, we have that E = ℓp and F = d(w, q) satisfy the three conditions of Proposition 3.5, with (ej)j and (fj)j the respective canonical basis of ℓp and d(w, q), see [30, Corollary 2]. Also, by Re- mark 3.6 and Proposition 3.7, we have that it does not exist P ∈ P(nℓp, d(w, q)) such that P (ej) = fj, for every j. Finally, as n < p q ≤ p ≤ cd(ℓp), all the hypothesis of Proposition 3.5 are fulfilled. There- fore, P(nℓp, d(w, q)) = Pwsc0(nℓp, d(w, q)). Now, by Proposition 3.4, P(nℓp, d(w, q)) = Pwsc(nℓp, d(w, q)) and the result follows from [8], since weakly sequentially continuous polynomials and weakly continuous polynomials on bounded sets coincide on ℓp. (cid:3) Let n = cd(ℓp, d(w, q)). Taking into account that for every k < n, any polynomial in P(kℓp, d(w, q)) is weakly continuous on bounded sets, from the last proposition we derive that there are two possible values for n: (I) p q ≤ n < p (II) n < p q + 1 and w 6∈ ℓ(cid:16) nq(cid:17)∗ \ ℓ(cid:16) q and w ∈ ℓ(cid:16) p p (n−1)q (cid:17)∗, or (n−1)q (cid:17)∗. p Theorem 3.9. Let n = cd(ℓp, d(w, q)). (a) If n and w satisfy condition (I) above, then • Pw(nℓp, d(w, q)) is an M -ideal in P(nℓp, d(w, q)), and • Pw(kℓp, d(w, q)) is not a semi M -ideal in P(kℓp, d(w, q)), for all k > n. (b) If n and w satisfy condition (II) above, then Pw(kℓp, d(w, q)) is not a semi M -ideal in P(kℓp, d(w, q)), for all k ≥ n. Proof. Suppose n and w satisfy condition (I) above. Then, n = cd(ℓp, d(w, q)) ≥ p q and cd(ℓp) is the integer number satisfying p ≤ cd(ℓp) < p + 1. If n < cd(ℓp), the hypothesis of Theorem 2.6 are fulfilled. If n = cd(ℓp) we may apply Proposition 2.2. In both cases the conclusion follows. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 19 Now, take k > cd(ℓp, d(w, q)). According to Proposition 3.1, the result is proven if we find polynomials P ∈ P(kℓp, d(w, q)) and Q ∈ Pw(kℓp, d(w, q)) such that there exists δ > 0 with kP k = kQk and kπmP + Qk ≥ δ > kP k, for all m. By Proposition 3.7, as k − 1 ≥ p q , the mapping R(x) = (xk−1 j Then, P (x) = e∗ kxkℓp = 1, 1(x)R(x) belongs to P(kℓp, d(w, q)). )j≥2 is a well defined norm one polynomial. In order to compute its norm, take x so that kP (x)kd(w,q) = x1kR(x)kd(w,q) ≤ x1k(xj )j≥2kk−1 ℓp k(cid:17) 1 p(cid:16)1 − 1 ≤(cid:16) 1 k(cid:17) k−1 p , where the last inequality was shown in the proof of Theorem 3.2. Now, with x = ( 1 k ) we have that P (x) = ( 1 k ) p e1, whence kP k = ( 1 k ) p (1 − 1 k ) p (1 − 1 k ) k−1 p . k−1 1 1 1 p e1 +(1− 1 k ) 1 p e2 ∈ Sℓp Let Q be the weakly continuous on bounded sets polynomial given by Q = kP k(e∗ kQk = kP k and x = ( 1 k ) 1 p e1 + (1 − 1 k ) 1 p em+2, for m ≥ 1, is a norm one vector so that 1)k · e1. Then, k(πmP + Q)(x)kd(w,q) = kP k(cid:13)(cid:13)(cid:13)em+1 + ( 1 k ) which completes the proof of (i). k p e1(cid:13)(cid:13)(cid:13)d(w,q) = kP k(cid:16)1 + w2( 1 k ) kq p (cid:17) 1 q > kP k, To prove (ii), take n and w satisfying condition (II) and take k ≥ n. Let us denote s = ( p j )j satisfies and W = kwkℓs . By Proposition 3.7 (b), the n-homogeneous polynomial R(x) = (xn nq )∗ = p p−nq kR(x)kd(w,q) = kRk = W 1 q , where x = W − s p (w s p j )j. Observe that x∗ = (w s p∗ j )j belongs to ℓp∗ and, as a continuous functional, it also attains its norm at x: x∗(x) = kx∗k = W s p∗ . Now we are ready to construct two polynomials P and Q fulfilling the statement of Proposition 3.1. Let P ∈ P(kℓp, d(w, q)) and Q ∈ Pw(kℓp, d(w, q)) be given by P (x) = x∗(x)k−nR(x) and Q(x) = W 1 q − sn p∗ x∗(x)ke1. It is easy to see that kP (x)kd(w,q) = kP k = W r = kQ(x)kd(w,q) = kQk, where r = s(k − n) p∗ + 1 q . Finally, kπmP + Qk ≥ k(πmP + Q)(x)kd(w,q) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) = W r ∞Xj=2 1 q wjxm−1+jnq  1 + W −1 W re1 + W s(k−n) p∗ ∞Xj=m+1 xn j ej(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)d(w,q) > W r = kP k. This completes the proof of the theorem. (cid:3) 20 VERÓNICA DIMANT AND SILVIA LASSALLE 4. Polynomial property (M ) Property (M ) was introduced by Kalton in [26]. It is a geometric property relating the norm of the traslation by a weakly null net of any two elements of the space. Namely, a Banach space X has property (M ) if for any x, x ∈ X such that kxk ≤ kxk, and any bounded weakly null net (xα)α in X, it holds that lim sup kx + xαk ≤ lim sup kx + xαk. An operator version of this property was given in [27]. Later on, in [16], it is extended to the scalar-valued polynomial context. In all these cases, these properties have incidence in the correspondent M -ideal problems. To study M -structures in spaces of vector-valued polynomials, we consider a suitable property (M ), which is the result of a natural combination of the definitions given for operators and scalar-valued polynomials. Before going on, let us state the vector- valued versions of [16, Lemma 3.1 and Theorem 3.2]. Lemma 4.1. If Pw(nE, F ) is an M -ideal in P(nE, F ) then, for each P ∈ P(nE, F ) there exists a bounded net {Pα}α ⊂ Pw(nE, F ) such that Pα(x) → P (x), for all x ∈ E. Proof. Fix P ∈ P(nE, F ). By [24, Remark I.1.13], we may consider {Qα}α a bounded net in Pw(nE, F ) such that Qα → P in the topology σ(cid:16)P(nE, F ), Pw(nE, F )∗(cid:17). Since ex ⊗ y∗ belongs to Pw(nE, F )∗, y∗(Qα(x)) = hex ⊗ y∗, Qαi → hex ⊗ y∗, P i = y∗(P (x)), for all x ∈ E and all y∗ ∈ F ∗. This says that Qα(x) w→ P (x), for all x ∈ E, which can be described, in analogy to the operator setting, as Qα → P in the WPT, the "weak polynomial topology". We can also consider on P(nE, F ) the "strong polynomial topology", SPT, naturally meaning pointwise convergence of nets. Both topologies, the WPT and the SPT, are locally convex and have the same continuous functionals (the proof of [19, Theorem VI.1.4] works also for polynomials). Thus, as in the linear case, we derive that the closure of any convex set in the strong polynomial topology coincides with its closure in the weak polynomial topology. Then, we may find Pα, a convex combination of Qα, converging pointwise to P . (cid:3) As a consequence of [32, Proposition 2.3] and the previous lemma, we have the following result which can be proved analogously to [32, Theorem 3.1]: Theorem 4.2. Let E and F be Banach spaces. The following are equivalent: (i) Pw(nE, F ) is an M -ideal in P(nE, F ). (ii) For all P ∈ P(nE, F ) there exists a net {Pα}α ⊂ Pw(nE, F ) such that Pα(x) → P (x), for all x ∈ E and lim sup kQ + P − Pαk ≤ max{kQk, kQkes + kP k}, for all Q ∈ P(nE, F ). (iii) For all P ∈ P(nE, F ) there exists a net {Pα}α ⊂ Pw(nE, F ) such that Pα(x) → P (x), for all x ∈ E and lim sup kQ + P − Pαk ≤ max{kQk, kP k}, for all Q ∈ Pw(nE, F ). Now we state the property (M ) for a vector-valued polynomial. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 21 Definition 4.3. Let P ∈ P(nE, F ) with kP k ≤ 1. We say that P has property (M ) if for all u ∈ E, v ∈ F with kvk ≤ kukn and for every bounded weakly null net {xα}α ⊂ E, it holds that lim sup α kv + P (xα)k ≤ lim sup α ku + xαkn. Note that every P ∈ Pw(nE, F ) with kP k ≤ 1 has property (M ). Analogously to [27, Lemma 6.2], we can prove: Lemma 4.4. Let P ∈ P(nE, F ) with kP k ≤ 1. If P has property (M ) then for all nets {uα}α and {vα}α contained in compact sets of E and F respectively, with kvαk ≤ kuαkn and for every bounded weakly null net {xα}α ⊂ E, it holds that lim sup α kvα + P (xα)k ≤ lim sup α kuα + xαkn. Definition 4.5. We say that a pair of Banach spaces (E, F ) has the n-polynomial property (M ) if every P ∈ P(nE, F ) with kP k ≤ 1 has property (M ). The next two results can be proved mimicking the proofs of Proposition 3.7 and Theorem 3.9 of [16]. Proposition 4.6. If Pw(nE, F ) is an M -ideal in P(nE, F ) and n = cd(E, F ) then (E, F ) has the n- polynomial property (M ). Theorem 4.7. Let E and F be Banach spaces and suppose that there exists a net of compact operators {Kα}α ∈ K(E) satisfying the following two conditions: • Kα(x) → x, for all x ∈ E and K ∗ • kId − 2Kαk −→ 1. α(x∗) → x∗, for all x∗ ∈ E∗. α Suppose also that n = cd(E, F ). Then, Pw(nE, F ) is an M -ideal in P(nE, F ) if and only if (E, F ) has the n-polynomial property (M ). Sometimes it is possible to infer M -structures in the space of linear continuous operators from the existence of geometric structures on the underlying space. For instance, it is proved in [24, Theorem VI.4.17] that K(E) is an M -ideal in L(E) if and only if E has property (M ) and satisfies both conditions of the theorem above. A similar result [16, Theorem 3.9] is obtained in the scalar-valued polynomial setting for n = cd(E) using the polynomial property (M). The following proposition (which is the vector- valued polynomial version of [24, Lemma VI.4.14] and [16, Proposition 3.10]) paves the way to connect the linear M -structure with M -ideals in vector valued polynomial spaces. Proposition 4.8. Let E and F be Banach spaces and n = cd(E, F ) < cd(E). If E and F have the property (M ), then (E, F ) has the n-polynomial property (M ). Proof. Let P ∈ P(nE, F ) with kP k = 1. Fix u ∈ E, v ∈ F with kvk ≤ kukn and a bounded weakly null net {xα}α ⊂ E. We want to prove that lim sup α kv + P (xα)k ≤ lim sup α ku + xαkn. 1 n x. Then, (1 − ε)kvk < kP (x)k ≤ Given ε > 0, take x ∈ SE such that kP (x)k > 1 − ε and x = kvk kxk ≤ kuk. As n < cd(E), every scalar valued polynomial in P(nE) is weakly continuous on bounded sets. Then, P is weak-to-weak continuous and P (xα) w→ 0. Therefore, since F has property (M ), 22 VERÓNICA DIMANT AND SILVIA LASSALLE lim sup α k(1 − ε)v + P (xα)k ≤ lim sup α kP (x) + P (xα)k = lim sup α ≤ lim sup α ≤ lim sup α kP (x + xα)k kx + xαkn ku + xαkn, where the last inequality holds since E has property (M ). Now, letting ε → 0 we obtain the desired inequality. If kP k < 1, the result follows from the previous case through the following convex combination v + P (xα) = 1 + kP k 2 (cid:18)v − (xα)(cid:19) . P kP k (cid:18)v + (xα)(cid:19) + P kP k 1 − kP k 2 (cid:3) Now we can lift M -structures from the linear to the vector-valued polynomial context. This is done for the particular case of n-homogeneous polynomials when n is the critical degree of the pair (E, F ) and it is strictly less than the critical degree of the domain space E. We do not know if the result remains true even for the case n = cd(E, F ) = cd(E). Corollary 4.9. Let E and F be Banach spaces and n = cd(E, F ) < cd(E). If K(E) is an M -ideal in L(E) and F has property (M ), then Pw(nE, F ) is an M -ideal in P(nE, F ). Proof. If K(E) is an M -ideal in L(E), appealing to [24, Theorem VI.4.17], E has property (M ) and we may find {Kα}α ⊂ K(E) a net of compact operators satisfying both conditions of Theorem 4.7. By Proposition 4.8, (E, F ) has the n-polynomial property (M ). Now, we may apply Theorem 4.7 to derive the result. (cid:3) We finish this section applying the previous result to give some examples of M -ideals of polynomials between Bergman and ℓp spaces. Example 4.10. The Bergman space Bp is the space of all holomorphic functions in Lp(D, dxdy), where D is the complex disc. If 1 < p < ∞, Bp is isomorphic to ℓp [33, Theorem III.A.11] and so, for 1 < p, q < ∞, cd(ℓp, ℓq) = cd(ℓp, Bq) = cd(Bp, ℓq) = cd(Bp, Bq). Since, by [27, Corollary 4.8], K(Bp) is an M -ideal in L(Bp), we obtain from Corollary 4.9, that, if n = cd(ℓp, ℓq) < cd(ℓp), then: • Pw(nℓp, Bq) is an M -ideal in P(nℓp, Bq). • Pw(nBp, ℓq) is an M -ideal in P(nBp, ℓq). • Pw(nBp, Bq) is an M -ideal in P(nBp, Bq). References [1] Acosta, María D. Denseness of norm attaining mappings. RACSAM Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. 100 (2006), no. 1-2, 9 -- 30. [2] Acosta, María D.; Aguirre, Francisco J.; Payá, Rafael. There is no bilinear Bishop-Phelps theorem. Israel J. Math. 93 (1996), 221 -- 227. M -STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES 23 [3] Alencar, Raymundo; Floret, Klaus. Weak-strong continuity of multilinear mappings and the Pelczynski-Pitt theorem. J. Math. Anal. Appl. 206 (1997), no. 2, 532 -- 546. [4] Alfsen, Erik; Effros, Edward. Structure in real Banach spaces. I, II. Ann. of Math. (2) 96 (1972), 98 -- 128; ibid. (2) 96 (1972), 129 -- 173. [5] Aron, Richard; Berner, Paul. A Hahn-Banach extension theorem for analytic mappings, Bull. Math. Soc. France, 106 (1978), 3 -- 24. [6] Aron, Richard; Dimant, Verónica. Sets of weak sequential continuity for polynomials. Indag. Math. (N.S.) 13 (2002), no. 3, 287 -- 299. [7] Aron, Richard; García, Domingo; Maestre, Manuel. On norm attaining polynomials. Publ. Res. Inst. Math. Sci. 39 (2003), no. 1, 165 -- 172. [8] Aron, Richard; Hervés, Carlos; Valdivia, Manuel. Weakly continuous mappings on Banach spaces. J. Funct. Anal. 52 (1983), 189-204. [9] Aron, Richard; Prolla, J. B. Polynomial approximation of differentiable functions on Banach spaces. J. Reine Angew. Math. 313 (1980), 195 -- 216. [10] Aron, Richard M.; Schottenloher, Martin. Compact holomorphic mappings on Banach spaces and the approximation property. J. Funct. Anal. 21 (1976), no. 1, 7 -- 30. [11] Asplund, Edgar. Farthest points in reflexive locally uniformly rotund Banach spaces. Israel J. Math. 4 (1966), 213 -- 216. [12] Bandyopadhyay, Pradipta; Lin, Bor-Luh; Rao, T. S. S. R. K. Ball remotal subspaces of Banach spaces. Colloq. Math. 114 (2009), no. 1, 119 -- 133. [13] Boyd, Christopher; Ryan, Raymond. A. Bounded weak continuity of homogeneous polynomials at the origin. Arch. Math. (Basel) 71 (1998), no. 3, 211 -- 218. [14] Carando, Daniel; Lassalle, Silvia. E ′ and its relation with vector-valued functions on E. Ark. Mat. 42 (2004), no. 2, 283 -- 300. [15] Choi, Yun Sung; Kim, Sung Guen. Norm or numerical radius attaining multilinear mappings and polynomials. J. London Math. Soc. (2) 54 (1996), no. 1, 135 -- 147. [16] Dimant, Verónica. M-ideals of homogeneous polynomials. Studia Math. 202 (2011), 81-104. [17] Dimant, Verónica; Gonzalo, Raquel. Block diagonal polynomials. Trans. Amer. Math. Soc. 353 (2001), no. 2, 733 -- 747. [18] Dineen, Seán. Complex analysis on infinite-dimensional spaces. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 1999. [19] Dunford, Nelson; Schwartz, Jacob T. Linear operators. Part I. General theory. Interscience Publishers, New York, 1958. [20] Edelstein, Michael. Farthest points of sets in uniformly convex Banach spaces. Israel J. Math. 4 (1966), 171 -- 176. [21] Fabian, Marián; Habala, Petr; Hájek, Petr; Montesinos Santalucía, Vicente; Pelant, Jan; Zizler, Václav. Functional analysis and infinite-dimensional geometry. CMS Books in Mathematics/Ouvrages de Mathámatiques de la SMC, 8. Springer-Verlag, New York, 2001. [22] González, Manuel; Gutiérrez, Joaquín M. The polynomial property (V). Arch. Math. (Basel) 75 (2000), no. 4, 299 -- 306. [23] Gonzalo, Raquel; Jaramillo, Jesús Angel. Compact polynomials between Banach spaces. Proc. Roy. Irish Acad. Sect. A 95 (1995), no. 2, 213 -- 226. [24] Harmand, Peter; Werner, Dirk; Werner, Wend. M-ideals in Banach spaces and Banach algebras. Lecture Notes in Mathematics, 1547. Springer-Verlag, Berlin, 1993. [25] Hennefeld, Julien. M-ideals, HB-subspaces, and compact operators. Indiana Univ. Math. J. 28 (1979), no. 6, 927 -- 934. [26] Kalton, Nigel J. M-ideals of compact operators. Illinois J. Math. 37 (1993), no. 1, 147 -- 169. [27] Kalton, Nigel J.; Werner, Dirk. Property (M ), M-ideals, and almost isometric structure of Banach spaces. J. Reine Angew. Math. 461 (1995), 137 -- 178. Asvald. M-ideals of compact operators in classical Banach spaces. Math. Scand. 44 (1979), no. 1, 207 -- 217. [28] Lima, [29] Mujica, Jorge. Complex analysis in Banach spaces. North-Holland Mathematics Studies, 120. North-Holland Publishing Co., Amsterdam, 1986. [30] Oja, Eve. On M-ideals of compact operators and Lorentz sequence spaces. Proc. Estonian Acad. Sci. Phys. Math. 40 (1991), no. 1, 31 -- 36. [31] Oja, Eve; Werner, Dirk. Remarks on M-ideals of compact operators on X ⊕p X. Math. Nachr. 152 (1991), 101 -- 111. [32] Werner, Dirk. M-ideals and the "basic inequality". J. Approx. Theory 76 (1994), no. 1, 21 -- 30. 24 VERÓNICA DIMANT AND SILVIA LASSALLE [33] Wojtaszczyk, P. Banach spaces for analysts. Cambridge Studies in Advanced Mathematics, 25. Cambridge University Press, Cambridge, 1991. Departamento de Matemática, Universidad de San Andrés, Vito Dumas 284, (B1644BID) Victoria, Buenos Aires, Argentina and CONICET. E-mail address: [email protected] Departamento de Matemática - Pab I, Facultad de Cs. Exactas y Naturales, Universidad de Buenos Aires, (1428) Buenos Aires, Argentina and CONICET. E-mail address: [email protected]
1304.0376
1
1304
2013-04-01T15:23:57
The Bishop-Phelps-Bollob\'as moduli of a Banach space
[ "math.FA" ]
We introduce two Bishop-Phelps-Bollob\'as moduli which measure, for a given Banach space, what is the best possible Bishop-Phelps-Bollob\'as theorem in this space. We show that there is a common upper bound for these moduli for all Banach spaces and we present an example showing that this bound is sharp. We prove the continuity of these moduli and an inequality with respect to duality. We calculate the two moduli for Hilbert spaces and also present many examples for which the moduli have the maximum possible value (among them, there are $C(K)$ spaces and $L_1(\mu)$ spaces). Finally, we show that if a Banach space has the maximum possible value of any of the moduli, then it contains almost isometric copies of the real space $\ell_\infty^{(2)}$ and present an example showing that this condition is not sufficient.
math.FA
math
BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE MARIO CHICA, VLADIMIR KADETS, MIGUEL MART´IN, SOLEDAD MORENO-PULIDO, AND FERNANDO RAMBLA-BARRENO Dedicated to the memory of Robert R. Phelps (1926 -- 2013) Abstract. We introduce two Bishop-Phelps-Bollob´as moduli of a Banach space which measure, for a given Banach space, what is the best possible Bishop-Phelps-Bollob´as theorem in this space. We show that there is a common upper bound for these moduli for all Banach spaces and we present an example showing that this bound is sharp. We prove the continuity of these moduli and an inequality with respect to duality. We calculate the two moduli for Hilbert spaces and also present many examples for which the moduli have the maximum possible value (among them, there are C(K) spaces and L1(µ) spaces). Finally, we show that if a Banach space has the maximum possible value of any of the moduli, then it contains almost isometric copies of the real space (cid:96)(2)∞ and present an example showing that this condition is not sufficient. 1. Introduction The classical Bishop-Phelps theorem of 1961 [4] states that the set of norm attaining functionals on a Banach space is norm dense in the dual space. Few years later, B. Bollob´as [5] gave a sharper version of this theorem allowing to approximate at the same time a functional and a vector in which it almost attains the norm (see the result bellow). The main aim of this paper is to study the best possible approximation of this kind that one may have in each Banach space, measuring it by using two moduli which we define. Before going further, we first present the original result by Bollob´as which nowadays is known as the Bishop-Phelps-Bollob´as theorem. We need to fix some notation. Given a (real or complex) Banach space X, we write BX and SX to denote the closed unit ball and the unit sphere of the space, and X∗ denotes the (topological) dual of X. We will also use the notation Π(X) :=(cid:8)(x, x∗) ∈ X × X∗ : (cid:107)x(cid:107) = (cid:107)x∗(cid:107) = x∗(x) = 1(cid:9). Theorem 1.1 (Bishop-Phelps-Bollob´as theorem [5]). Let X be a Banach space. Suppose x ∈ SX and x∗ ∈ SX∗ satisfy 1 − x∗(x) (cid:54) ε2/2 (0 < ε < 1/2). Then there exists (y, y∗) ∈ Π(X) such that (cid:107)x − y(cid:107) < ε + ε2 and (cid:107)x∗ − y∗(cid:107) (cid:54) ε. So the idea is that given (x, x∗) ∈ SX × SX∗ such that x∗(x) ∼ 1, there exist y ∈ SX close to x and y∗ ∈ SX∗ close to x∗ for which y∗(y) = 1. This result has many applications, especially for the theory of numerical ranges, see [5, 6]. Our objective is to introduce two moduli which measures, for a given Banach space, what is the best possible Bollob´as theorem in this space, that is, how close can be y to x and y∗ to x∗ in the result above depending on how close is x∗(x) to 1. In the first modulus, we allow the vector and the functional to Date: March 27th, 2013. 2010 Mathematics Subject Classification. Primary: 46B04. Key words and phrases. Banach space; approximation; uniformly non-square spaces. First author partially supported by Spanish MINECO and FEDER project no. MTM2012-31755 and by Junta de Andaluc´ıa and FEDER grant FQM-185. Second author partially supported by Junta de Andaluc´ıa and FEDER grants FQM-185 and P09-FQM-4911 and by the program GENIL-PRIE of the CEI of the University of Granada. Third author partially supported by Spanish MINECO and FEDER project no. MTM2012-31755, and by Junta de Andaluc´ıa and FEDER grants FQM-185 and P09-FQM-4911. Fourth and Fifth authors partially supported by Junta de Andaluc´ıa and FEDER grant FQM-257. 2 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO have norm less than or equal to one, whereas in the second modulus we only consider norm-one vectors and functionals. Definitions 1.2 (Bishop-Phelps-Bollob´as modulus). Let X be a Banach space. The Bishop-Phelps-Bollob´as modulus of X is the function ΦX : (0, 2) −→ R+ such that given δ ∈ (0, 2), ΦX (δ) is the infimum of those ε > 0 satisfying that for every (x, x∗) ∈ BX×BX∗ with Re x∗(x) > 1 − δ, there is (y, y∗) ∈ Π(X) with (cid:107)x − y(cid:107) < ε and (cid:107)x∗ − y∗(cid:107) < ε. The spherical Bishop-Phelps-Bollob´as modulus of X is the function ΦS given δ ∈ (0, 2), ΦS Re x∗(x) > 1 − δ, there is (y, y∗) ∈ Π(X) with (cid:107)x − y(cid:107) < ε and (cid:107)x∗ − y∗(cid:107) < ε. X : (0, 2) −→ R+ such that X (δ) is the infimum of those ε > 0 satisfying that for every (x, x∗) ∈ SX × SX∗ with Evidently, ΦS X (δ) (cid:54) ΦX (δ), so any estimation from above for ΦX (δ) is also valid for ΦS X (δ) and, viceversa, any estimation from below for ΦS X (δ) is also valid for ΦX (δ). Recall that the dual of a complex Banach space X is isometric (taking real parts) to the dual of the real subjacent space XR. Also, Π(X) does not change if we consider X as a real Banach space (indeed, if (x, x∗) ∈ Π(X) then x∗ ∈ SX∗ and x ∈ SX satisfies x∗(x) = 1 so, obviously, Re x∗(x) = 1 and (x, Re x∗) ∈ Π(XR)). Therefore, only the real structure of the space is playing a role in the above definitions. We could then suppose that we are only dealing with real Banach spaces and any result would apply automatically to complex spaces. Nevertheless, we are not going to do so, mainly because for classical sequence or function spaces, the real space underlying the complex version of the space is not equal, in general, to the real version of the space. We then prefer to develop the theory for real and complex spaces which, actually, does not suppose much more effort. Unless otherwise is stated, the (arbitrary or concrete) spaces we are dealing with will be real or complex and the results work in both cases. Some notation will help to the understanding and further use of Definitions 1.2. Let X be a Banach space and fix 0 < δ < 2. Writing AX (δ) :=(cid:8)(x, x∗) ∈ BX × BX∗ : Re x∗(x) > 1 − δ(cid:9), X (δ) :=(cid:8)(x, x∗) ∈ SX × SX∗ : Re x∗(x) > 1 − δ(cid:9), AS it is clear that ΦX (δ) = sup (x,x∗)∈AX (δ) inf (y,y∗)∈Π(X) max{(cid:107)x − y(cid:107),(cid:107)x∗ − y∗(cid:107)}, max{(cid:107)x − y(cid:107),(cid:107)x∗ − y∗(cid:107)}. ΦS X (δ) = sup (x,x∗)∈AS inf (y,y∗)∈Π(X) the (cid:96)∞-distance d∞ in X×X∗ (that is, d∞(cid:0)(x, x∗), (y, y∗)(cid:1) = max{(cid:107)x−y(cid:107),(cid:107)x∗−y∗(cid:107)} for (x, x∗), (y, y∗) ∈ Therefore, if we write dH (A, B) to denote the Hausdorff distance between A, B ⊂ X × X∗ associated to X × X∗, and X (δ) (cid:26) (cid:0)AX (δ), Π(X)(cid:1) sup a∈A for A, B ⊂ X × X∗), then we clearly have that dH (A, B) = max inf b∈B d∞(a, b) , sup b∈B inf a∈A ΦX (δ) = dH and ΦS X (δ) = dH for every 0 < δ < 2 (observe that Π(X) ⊂ AX (δ) and Π(X) ⊂ AS X (δ) for every δ). d∞(a, b) (cid:27) (cid:0)AS X (δ), Π(X)(cid:1) The following result is immediate. Remark 1.3. Let X be a Banach space. Given δ1, δ2 ∈ (0, 2) with δ1 < δ2, one has Therefore, the functions ΦX (·) and ΦS AX (δ1) ⊂ AX (δ2) and AS X (·) are increasing. X (δ1) ⊂ AS X (δ2). Routine computations and the fact that the Hausdorff distance does not change if we take closure in one of the sets, provide the following observations. Remark 1.4. Let X be a Banach space. Then, for every δ ∈ (0, 2), one has 3 BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE ΦX (δ) := inf(cid:8)ε > 0 : ∀(x, x∗) ∈ BX × BX∗ with Re x∗(x) > 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) < ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ BX × BX∗ with Re x∗(x) (cid:62) 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) < ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ BX × BX∗ with Re x∗(x) > 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) (cid:54) ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ BX × BX∗ with Re x∗(x) (cid:62) 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) (cid:54) ε(cid:9), X (δ) := inf(cid:8)ε > 0 : ∀(x, x∗) ∈ SX × SX∗ with Re x∗(x) > 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) < ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ SX × SX∗ with Re x∗(x) (cid:62) 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) < ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ SX × SX∗ with Re x∗(x) > 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) (cid:54) ε(cid:9) = inf(cid:8)ε > 0 : ∀(x, x∗) ∈ SX × SX∗ with Re x∗(x) (cid:62) 1 − δ, ∃(y, y∗) ∈ Π(X) with d∞((x, x∗), (y, y∗)) (cid:54) ε(cid:9). ΦS and Observe that the smaller are the functions ΦX (·) and ΦS X (·), the better is the approximation on the space. It can be deduced from the Bishop-Phelps-Bollob´as theorem that there is a common upper bound X (·) for all Banach spaces X. Our first result in the next section will be to present the for ΦX (·) and ΦS best possible upper bound, namely we will show that (cid:0)0 < δ < 2, X Banach space(cid:1). X (δ) (cid:54) ΦX (δ) (cid:54) √ ΦS 2δ (1) This will follow from a result by R. Phelps [13]. A version for ΦS X (δ) for small δ's can be also deduced from the Brøndsted-Rockafellar variational principle [14, Theorem 3.17], as claimed in [7]. The sharpness of (1) can be verified by considering the real space X = (cid:96)(2)∞ . This is the content of section 2. Next, we prove in section 3 that for every Banach space X, the moduli ΦX (δ) and ΦS X∗ (δ). Finally, we show that ΦX (δ) = X (δ) are continuous 2δ if and √ in δ. We prove that ΦX (δ) (cid:54) ΦX∗ (δ) and ΦS only if ΦS √ 2δ. X (δ) = X (δ) (cid:54) ΦS Examples of spaces for which the two moduli are computed are presented in section 4. Among other results, the moduli of R and of every real or complex Hilbert space of (real)-dimension greater than one 2δ are calculated, and there are presented a number of spaces for which the value of both moduli are (i.e. the maximal possible value) for small δ's: namely c0, (cid:96)1 and, more in general, L1(µ), C0(L), unital C∗-algebras with non-trivial centralizer. . . √ √ 2δ0 (equivalently, The main result of section 5 states that if a Banach space X satisfies ΦX (δ0) = 2δ0) for some δ0 ∈ (0, 1/2), then X contains almost isometric copies of the real space (cid:96)(2)∞ . ΦS X (δ0) = We finish presenting, for every δ ∈ (0, 1/2), an example of a three dimensional real space Z containing an isometric copy of (cid:96)(2)∞ for which ΦZ(δ) < 2δ. This is the content of section 6. √ √ 4 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO 2. The upper bound of the moduli Our first result is the promised best upper bound of the Bishop-Phelps-Bollob´as moduli. Theorem 2.1. For every Banach space X and every δ ∈ (0, 2), ΦX (δ) (cid:54) √ 2δ and so, ΦS X (δ) (cid:54) √ 2δ We deduce the above result from [13, Corollary 2.2], which was stated for general bounded convex sets on real Banach spaces. Particularizing the result to the case of the unit ball of a Banach space, using a routine argument to change non-strict inequalities to strict inequalities, and taking into account that the dual of a complex Banach space is isometric (taking real parts) to the dual of the real subjacent space, we get the following result. Proposition 2.2 (Particular case of [13, Corollary 2.2]). Let X be Banach space. Suppose that z∗ ∈ SX∗ , z ∈ BX and η > 0 are given such that Re z∗(z) > 1 − η. Then, for any k ∈ (0, 1) there exist y∗ ∈ X∗ and y ∈ SX such that (cid:107)y∗(cid:107) = y∗(y), (cid:107)z − y(cid:107) < (cid:107)z∗ − y∗(cid:107) < k. η k , Proof of Theorem 2.1. We have to show that given (x, x∗) ∈ BX × BX∗ with Re x∗(x) > 1 − δ, there exists (y, y∗) ∈ Π(X) such that (cid:107)x − y(cid:107) < 2δ. Let us first prove the result for the more interesting case of δ ∈ (0, 1). In this case, 2δ and (cid:107)x∗ − y∗(cid:107) < √ √ (cid:107)y∗(cid:107) = y∗(y), (cid:107)z − y(cid:107) < 2δ, As k < 1, we get y∗ (cid:54)= 0 and we may write y∗ = y∗ have that (cid:107)x − y(cid:107) < 2δ. On the other hand, we have √ = (cid:13)(cid:13)(cid:13)(cid:13)x∗ − y∗ (cid:107)y∗(cid:107) (cid:107)x∗ − y∗(cid:107) = (cid:107)y∗(cid:107) , y = y, to get that (y, y∗) ∈ Π(X). We already . η k √ (cid:107)x∗(cid:107) − 1 + δ (cid:107)x∗(cid:107)√ 2δ (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − y∗(cid:13)(cid:13)(cid:13)(cid:13) < k = (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) (cid:54)(cid:13)(cid:13)(cid:13)x∗ − (cid:107)x∗(cid:107)y∗(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13)(cid:13)(cid:107)x∗(cid:107)y∗ − y∗ (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − y∗(cid:13)(cid:13)(cid:13)(cid:13) +(cid:12)(cid:12)(cid:107)x∗(cid:107)(cid:107)y∗(cid:107) − 1(cid:12)(cid:12) (cid:107)x∗(cid:107) − y∗(cid:13)(cid:13)(cid:13)(cid:13) +(cid:12)(cid:12)(cid:107)x∗(cid:107)(cid:107)y∗(cid:107) − (cid:107)x∗(cid:107)(cid:12)(cid:12) +(cid:12)(cid:12)1 − (cid:107)x∗(cid:107)(cid:12)(cid:12) (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − y∗(cid:13)(cid:13)(cid:13)(cid:13) +(cid:12)(cid:12)(cid:107)y∗(cid:107) − 1(cid:12)(cid:12)(cid:21) (cid:20)(cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − y∗(cid:13)(cid:13)(cid:13)(cid:13) + 1 − (cid:107)x∗(cid:107) (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:0)(cid:107)x∗(cid:107) − 1 + δ(cid:1) + 1 − (cid:107)x∗(cid:107). + 1 − (cid:107)x∗(cid:107) (cid:54) 2(cid:107)x∗(cid:107) (cid:54) (cid:107)x∗(cid:107) (cid:54) (cid:107)x∗(cid:107) (cid:54) (cid:107)x∗(cid:107) (cid:107)y∗(cid:107) < 2√ 2δ (cid:107)x∗(cid:107) − 1 + δ so, if we write η = (cid:107)x∗(cid:107) √ Next, we consider k = η/ 0 < 1 − δ < (cid:107)x∗(cid:107) (cid:54) 1, > 0, z∗ = x∗/(cid:107)x∗(cid:107) and z = x, one has Re z∗(z) > 1 − η. 2δ and claim that 0 < k < 1. Indeed, as the function (2) is strictly increasing, k = ϕ((cid:107)x∗(cid:107)) and 1 − δ < (cid:107)x∗(cid:107) (cid:54) 1, we have that ϕ(t) = t − 1 + δ √ 2δ t (t ∈ R+) 0 = ϕ(1 − δ) < k (cid:54) ϕ(1) = < 1, √ δ√ 2 as desired. Therefore, we may apply Proposition 2.2 with z∗ ∈ SX∗ , z ∈ BX , η > 0 and 0 < k < 1 to obtain y∗ ∈ X∗ and y ∈ SX satisfying so, writing we get < δ − 1√ 2δ − 1 (cid:18) δ − 1√ 1 2 √ 2δ − 1 √ + 2δ − 1 √ 2δ − 1 2δ − 1 (cid:19) (cid:0)δ ∈ [1, 2)(cid:1). ψ(δ) = δ − 1√ 2δ − 1 Now, as the function γ(t) = 2√ 2δ BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE (cid:0)t − 1 + δ(cid:1) + 1 − t (cid:0)t ∈ [0, 1](cid:1) is strictly increasing (for this we only need 0 < δ < 2), we get γ((cid:107)x∗(cid:107)) (cid:54) γ(1) = 2δ√ that (cid:107)x∗ − y∗(cid:107) < 2δ, as desired. √ Let us now prove the case when δ ∈ [1, 2). Here, it can be routinely verified that 2δ 5 √ = 2δ. It follows (3) Now, we have to distinguish two situations. Let first suppose that (cid:107)x∗(cid:107) (cid:54) ψ(δ). Then, we take any y ∈ SX √ such that (cid:107)x − y(cid:107) (cid:54) 1 and take y∗ ∈ SX∗ such that y∗(y) = 1. Then, (y, y∗) ∈ Π(X), (cid:107)x − y(cid:107) (cid:54) 1 < 2δ and < ψ(δ) < (cid:107)x∗ − y∗(cid:107) (cid:54) 1 + (cid:107)x∗(cid:107) (cid:54) 1 + ψ(δ) < by (3). Otherwise, suppose (cid:107)x∗(cid:107) > ψ(δ). We then write η = previous case, and we have to show that k < 1. This is trivial for the case δ = 1 and for δ > 1, we use that the function ϕ defined in (2) is now strictly decreasing to get that 2δ as in the (cid:107)x∗(cid:107) √ > 0 and k = η/ k = ϕ((cid:107)x∗(cid:107)) < ϕ(cid:0)ψ(δ)(cid:1) < ϕ √ 2δ (cid:107)x∗(cid:107) − 1 + δ (cid:19) (cid:18) δ − 1√ 2δ − 1 = 1. Then, the rest of the proof follows the same lines of the case when δ ∈ (0, 1) since this hypothesis is not (cid:3) longer used. Let us comment that the above proof is much simpler if we restrict to x∗ ∈ SX∗ (in particular, to the spherical modulus ΦS X (δ)), but the result for non-unital functionals is stronger. Actually, the following stronger version can be deduced by conveniently modifying the election of k in the proof of Theorem 2.1. Remark 2.3. For every 0 < θ < 1 and every 0 < δ < 2, there is ρ = ρ(δ, θ) > 0 such that for every Banach space X, if x∗ ∈ BX∗ with (cid:107)x∗(cid:107) (cid:54) θ, x ∈ BX satisfy that Re x∗(x) > 1 − δ, then there is a pair (y, y∗) ∈ Π(X) satisfying √ (cid:107)x − y(cid:107) < 2δ − ρ and (cid:107)x∗ − y∗(cid:107) < 2δ − ρ. √ Let us observe that, given 0 < θ < 1, the hypothesis above is not empty only when 1 − θ < δ. On the other hand, in the proof it is sufficient to consider only the case of δ < 1 + θ, because, otherwise, the evident inequality Re x∗(x) > −θ = 1 − (1 + θ) implies that there is a pair (y, y∗) ∈ Π(X) satisfying (cid:107)x − y(cid:107) < (cid:112)2(1 + θ) and (cid:107)x∗ − y∗(cid:107) < (cid:112)2(1 + θ), so the statement of our remark holds true with 2δ −(cid:112)2(1 + θ). √ ρ := Next, we rewrite Theorem 2.1 in two equivalent ways. Corollary 2.4. Let X be a Banach space. (a) Let 0 < ε < 2 and suppose that x ∈ BX and x∗ ∈ BX∗ satisfy Re x∗(x) > 1 − ε2/2. Then, there exists (y, y∗) ∈ Π(X) such that and (cid:107)x − y(cid:107) < ε (cid:107)x∗ − y∗(cid:107) < ε. 6 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO (b) Let 0 < δ < 2 and suppose that x ∈ BX and x∗ ∈ BX∗ satisfy Then, there exists (y, y∗) ∈ Π(X) such that and (cid:107)x − y(cid:107) < 2δ (cid:107)x∗ − y∗(cid:107) < √ 2δ. Re x∗(x) > 1 − δ. √ As the last result of this section, we present an example of a Banach space for which the estimate in Theorem 2.1 is sharp. Example 2.5. Let X be the real space (cid:96)(2)∞ . Then, ΦS X (δ) = ΦX (δ) = √ 2δ for all δ ∈ (0, 2). Proof. Fix 0 < δ < 2. We consider √ z = (1 − 2δ, 1) ∈ SX and z∗ = (cid:32)√ (cid:33) √ 2δ 2 , 1 − 2δ 2 ∈ SX∗ , √ and observe that z∗(z) = 1 − δ. Now, suppose we may find (y, y∗) ∈ Π(X) such that (cid:107)z − y(cid:107) < 2δ and (cid:107)z∗ − y∗(cid:107) < 2δ. By the shape of BX , we only have two possibilities: either y is an extreme point of BX or y∗ is an extreme point of BX∗ (this is actually true for all two-dimensional real spaces). Suppose first that y is an extreme point of BX , which has the form y = (a, b) with a, b ∈ {−1, 1}. As we are forced to have b = 1 and a = −1. Now, we have y∗ = (−t, 1 − t) for some 0 (cid:54) t (cid:54) 1 and (cid:107)z − y(cid:107) = max{1 − 2δ − a,1 − b} < √ (cid:107)z∗ − y∗(cid:107) = + t + √ 2δ 2 √ 2δ, (cid:111) (cid:62) √ 2δ, 2t 2δ, (cid:110)√ √ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)t − (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)√ 2δ 2 √ 2δ 2 + √ 2δ 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = max (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + 1 − (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1 − 2δ 2 √ √ √ < 2δ, 2δ 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < − b √ 2δ, (cid:107)z∗ − y∗(cid:107) = − a (cid:107)z∗ − y∗(cid:107) = a contradiction. On the other hand, if y∗ is an extreme point of BX∗ , then either y∗ = (a, 0) or y∗ = (0, b) for suitable a, b ∈ {−1, 1}. In the first case, as we are forced to have a = 1 and so, y = (1, s) for suitable s ∈ [−1, 1]. But then (cid:107)z − y(cid:107) (cid:62) √ impossible. In case y∗ = (0, b) with b = ±1, we have 2δ, which is so b = −1 and therefore, y = (s,−1) for suitable s ∈ [−1, 1], giving (cid:107)z − y(cid:107) (cid:62) 2, a contradiction. (cid:3) 3. Basic properties of the moduli Our first result is the continuity of the Bishop-Phelps-Bollob´as moduli. Proposition 3.1. Let X be a Banach space. Then, the functions δ (cid:55)−→ ΦS δ (cid:55)−→ ΦX (δ) and X (δ) are continuous in (0, 2). We need the following three lemmata which could be of independent interest. Lemma 3.2. For every pair (x0, x∗ 0) ∈ BX × BX∗ there is a pair (y, y∗) ∈ Π(X) with Moreover, if actually Re x∗ 0(x0) > 0 then (y, y∗) ∈ Π(X) can be selected to satisfy Re(cid:2)y∗(x0) + x∗ 0(y)(cid:3) (cid:62) 0. (cid:113) Re(cid:0)x∗ 0(y)(cid:3) (cid:62) 2 Re(cid:2)y∗(x0) + x∗ 0(x0)(cid:1). BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 7 Proof. 1. Take y0 ∈ SX ∩ ker x∗ 0 and let y∗ Re(cid:2)y∗ 0(y0)(cid:3) = Re y∗ 0 be a supporting functional at y0. Then 0(x0) + x∗ 0(x0) If the right hand side is positive we can take y = y0, y∗ = y∗ 0, in the opposite case take y = −y0, y∗ = −y∗ 0. 2. Take y = x0(cid:107)x0(cid:107) and let y∗ be a supporting functional at y. Then, since for a fixed a > 0 the minimum of f (t) := t + a √ t for t > 0 equals 2 Re(cid:2)y∗(x0) + x∗ a, we get 0(y)(cid:3) = (cid:107)x0(cid:107) + (cid:113) 1 (cid:107)x0(cid:107) Re x∗ 0(x0) (cid:62) 2 Re x∗ 0(x0). (cid:3) The above lemma allows us to prove the following result which we will use to show the continuity of the Bishop-Phelps-Bollob´as modulus. Lemma 3.3. Let X be a Banach space. Suppose (x0, x∗ Case 1: If δ, δ0 ∈]0, 1] then 0), AX (δ)(cid:1) (cid:54) 2 dist(cid:0)(x0, x∗ 0), AX (δ)(cid:1) (cid:54) 2 dist(cid:0)(x0, x∗ Case 2: If δ, δ0 ∈ [1, 2) then 0) ∈ AX (δ0) with 0 < δ < δ0 < 2. Then: √ 1 − δ − √ 1 − √ 1 − δ0 1 − δ0 . 2 − δ0 δ0 · δ0 − δ √ 1 − 2δ + δδ0 . δ0 − 1 + 0(x0). Let (y, y∗) ∈ Π(X) be from the previous lemma (in case 1 we use part Proof. Denote t = Re x∗ 2 of the lemma, in case 2 we use part 1). For every λ ∈ [0, 1] we define xλ = (1 − λ)x0 + λy and λ = (1− λ)x∗ λ)) (cid:54) 2λ. x∗ We have: λ belong to corresponding balls, and dist ∞ ((x0, x∗ 0 + λy∗. Both xλ and x∗ 0), (xλ, x∗ λ(xλ) = (1 − λ)2t + λ(1 − λ) Re(cid:2)y∗(x0) + x∗ Re x∗ Now we are looking for a possibly small value of λ, for which (xλ, x∗ λ = 0 is already ok and dist ∞ ((x0, x∗ √ λ(xλ) (cid:62) (1 − λ)2t + 2λ(1 − λ) √ (1 − λ) λ) ∈ AX (δ). If δ (cid:62) 1 − t, the value 0), AX (δ)) = 0. If 0 < δ < 1 − t then the positive solution in λ of t + λ2 = Re x∗ t + λ . the equation(cid:0)(1 − λ) t + λ(cid:1)2 √ = 1 − δ is 0(y)(cid:3) + λ2, (cid:17)2 (cid:16) (4) so in case 1 √ λt = 1 − δ − √ 1 − √ t t . Evidently, λt ∈ [0, 1], so (xλt, x∗ λt ) ∈ AX (δ). Since λt decreases in t, √ 0), AX (δ)) (cid:54) 2λt (cid:54) 2λ1−δ0 = 2 dist ∞ ((x0, x∗ 1 − δ − √ 1 − √ 1 − δ0 1 − δ0 . This completes the proof of case 1. In the case 2 we may assume t (cid:54) 1− δ (otherwise the corresponding distance is 0 and the job is done), so t (cid:54) 0. By part 1 of the previous lemma and (4) Re x∗ λ(xλ) (cid:62) (1 − λ)2t + λ2, so we are solving in λ the equation (1 − λ)2t + λ2 − 1 + δ = 0, i.e. (1 + t)λ2 − 2tλ + (t − 1 + δ) = 0. The discriminant of this equation is D = −tδ − δ + 1. Remark that D (cid:62) −(1− δ)δ − δ + 1 = (1− δ)2 (cid:62) 0 and t − 1 + δ (cid:54) 0, so there is a positive solution of our equation given by 1 − tδ − δ). D) = (t + (t + √ √ 1 1 λt = 1 + t 1 + t 8 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO This λt decreases in t, so λt (cid:54) λ1−δ0 = (cid:112) (1 − δ0 + 1 δ0 1 − 2δ + δδ0) = 2 + δ0 δ0 · δ0 − 1 + δ0 − δ √ 1 − 2δ + δδ0 . (cid:3) For the continuity of the spherical modulus, we need the following result. Lemma 3.4. Let X be a Banach space. Suppose (x0, x∗ Case 1: If δ < 1 then 0) ∈ AS X (δ0) with 0 < δ < δ0 < 2. Then: Case 2: If δ ∈ [1, 2) and 2 − √ dist ∞(cid:0)(x0, x∗ dist ∞(cid:0)(x0, x∗ 0), AS 2 − δ0 < δ < δ0, then 0), AS X (δ)(cid:1) (cid:54) 4(δ0 − δ) X (δ)(cid:1) (cid:54) 2(δ0 − δ) δ0 2 − δ . . Proof. Let us start with case 1. Fix ξ ∈ (0, δ). As (cid:107)x∗ For every λ ∈ [0, 1] we define 0(cid:107) = 1, we may find yξ ∈ SX satisfying x∗ 0(yξ) > 1−ξ. x(λ, ξ) = λx0 + (1 − λ)yξ. Consider λξ = δ−ξ δ0−ξ ∈ [0, 1] and write xξ = x(λξ, ξ). An straightforward verification shows that and so, as 1 − δ (cid:62) 0, we have that xξ (cid:54)= 0 and also that Re x∗ 0(xξ) > 1 − δ X (δ). It remains to estimate (cid:16) xξ(cid:107)xξ(cid:107) , x∗ 0 0 Re x∗ (cid:17) ∈ AS (cid:13)(cid:13)(cid:13)(cid:13)x0 − xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) (cid:107)x0 − xξ(cid:107) + (cid:19) (cid:18) δ0 − δ (cid:54) 2 δ0 − ξ + (cid:107)xξ(cid:107) − (cid:107)x0(cid:107) (cid:54) 2 > 1 − δ. (cid:19) (cid:18) xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13)x0 − xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13) as follows: (cid:13)(cid:13)(cid:13)(cid:13)xξ − xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) 2 (cid:18) δ0 − δ (cid:19) (cid:18) δ0 − δ (cid:19) δ0 − ξ + (cid:107)xξ − x0(cid:107) (cid:54) 4 δ0 − ξ + (cid:107)xξ(cid:107) − 1 (cid:54) (cid:19) (cid:18) δ0 − δ δ0 − ξ . Therefore, We get the result by just letting ξ −→ 0. Let us prove case 2. We have to distinguish the values of Re x∗ 0(x0). If Re x∗ 0(x0) > 1 − δ, then the proof is done. Suppose otherwise that Fix ξ ∈(cid:16) 0, min{2 − δ0, 4δ−2−δ0−δ2 we may find yξ ∈ SX satisfying x∗ δ−1 }(cid:17) 1 − δ (cid:62) Re x∗ (observe that 4δ−2−δ0−δ2 0(x0) > 1 − δ0. δ−1 0(yξ) > 1 − ξ. Now, we consider λξ = δ0 − δ 2 − δ − ξ and xξ = x0 + λξyξ. Notice that λξ ∈ (0, 1) (since δ < δ0 and ξ < 2 − δ0) and (cid:107)xξ(cid:107) (cid:62) (cid:107)x0(cid:107) − λ(cid:107)yξ(cid:107) = 1 − λξ > 0. > 0 by the conditions on δ). As (cid:107)x∗ 0(cid:107) = 1, Also, observe that so, Re x∗ Re x∗ 0(xξ) (cid:54) 1 − δ + λξ = (cid:19) 0(xξ) (cid:54) 0 since ξ (cid:54) 4δ−2−δ0−δ2 (cid:18) xξ . Now, δ−1 (cid:18) xξ(cid:107)xξ(cid:107) (cid:62) Re x∗ 0 1 − λξ Re x∗ 0 (cid:19) > (1 − δ)(2 − δ − ξ) + δ0 − δ 2 − δ − ξ 1 − δ0 + λξ(1 − ξ) 1 − λξ = 1 − δ. Therefore, (cid:16) xξ(cid:107)xξ(cid:107) , x∗ (cid:13)(cid:13)(cid:13)(cid:13)x0 − xξ(cid:107)xξ(cid:107) 0 (cid:17) ∈ AS (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) (cid:107)x0 − xξ(cid:107) + (cid:54) δ0 − δ 2 − δ − ξ Consequently, letting ξ −→ 0, we get BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE X (δ). It remains to estimate (cid:13)(cid:13)(cid:13) as follows: +(cid:12)(cid:12)(cid:107)xξ(cid:107) − 1(cid:12)(cid:12) (cid:54) (cid:13)(cid:13)(cid:13)x0 − xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13)(cid:13)xξ − xξ(cid:107)xξ(cid:107) (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) δ0 − δ +(cid:12)(cid:12)(cid:107)xξ(cid:107) − (cid:107)x0(cid:107)(cid:12)(cid:12) (cid:54) δ0 − δ X (δ)(cid:1) (cid:54) 2(δ0 − δ) dist ∞(cid:0)(x0, x∗ 2 − δ − ξ 2 − δ − ξ 0), AS 2 − δ . + (cid:107)xξ − x0(cid:107) (cid:54) 2 (cid:19) (cid:18) δ0 − δ 2 − δ − ξ . 9 (cid:3) Proof of Proposition 3.1. Let us give the proof for ΦX (δ). Observe that for δ1, δ2 ∈ (0, 2) with δ1 < δ2, one has 0 < ΦX (δ2) − ΦX (δ1) = dH (AX (δ2), Π(X)) − dH (AX (δ1), Π(X)) (cid:54) dH (AX (δ2), AX (δ1)) . Now, the continuity follows routinely from Lemma 3.3. An analogous argument allows to prove the continuity of ΦS X (δ) from Lemma 3.4. (cid:3) The following lemma will be used to show that the approximation in the space is not worse than the approximation in the dual. It is actually an easy application of the Principle of Local Reflexivity. Lemma 3.5. For ε > 0, let (x, x∗) ∈ BX × BX∗ and let (y∗, y∗∗) ∈ Π(Y ∗) such that (cid:107)x∗ − y∗(cid:107) < ε and (cid:107)x − y∗∗(cid:107) < ε. Then there is a pair (y, y∗) ∈ Π(X) such that (cid:107)x − y(cid:107) < ε and (cid:107)x∗ − y∗(cid:107) < ε. Proof. First chose ε(cid:48) < ε such that still (cid:107)x∗ − y∗(cid:107) < ε(cid:48) Now, we consider ξ > 0 such that and (cid:107)x − y∗∗(cid:107) < ε(cid:48). (cid:115) (1 + ξ)ε(cid:48) + ξ + 2ξ 1 + ξ < ε, and use the Principle of Local Reflexivity (see [1, Theorem 11.2.4], for instance) to get an operator T : Lin{x, y∗∗} −→ X satisfying Next, we consider x = (cid:107)T(cid:107),(cid:107)T −1(cid:107) (cid:54) 1 + ξ, T (x) = x, T (y∗∗) (cid:107)T (y∗∗)(cid:107) ∈ SX and x∗ = y∗ ∈ SX∗ , observe that y∗(T (y∗∗)) = y∗∗(y∗) = 1. and we use Corollary 2.4 to get (y, y∗) ∈ Π(X) satisfying that Re x∗(x) > 1 1 + ξ = 1 − ξ 1 + ξ , (cid:115) (cid:107)x − y(cid:107) < 2ξ 1 + ξ and (cid:107)x∗ − y∗(cid:107) < (cid:115) 2ξ 1 + ξ . Let us show that (y, y∗) ∈ Π(X) fulfill our requirements: (cid:107)x − y(cid:107) (cid:54) (cid:107)T (x) − T (y∗∗)(cid:107) + (cid:107)T (y∗∗) − x(cid:107) + (cid:107)x − y(cid:107) (cid:115) < (1 + ξ)ε(cid:48) + ξ + 2ξ 1 + ξ < ε 10 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO and, analogously, (cid:107)x∗ − y∗(cid:107) (cid:54) (cid:107)x∗ − y∗(cid:107) + (cid:107)y∗ − y∗(cid:107) < ε(cid:48) + (cid:115) 2ξ 1 + ξ < ε. (cid:3) Proposition 3.6. Let X be a Banach space. Then for every δ ∈ (0, 2). ΦX (δ) (cid:54) ΦX∗ (δ) and ΦS X (δ) (cid:54) ΦS X∗ (δ) Proof. The proof is the same for both moduli, so we are only giving the case of ΦX (δ). Fix δ ∈ (0, 2). We consider any ε > 0 such that ΦX∗ (δ) < ε and for a given (x, x∗) ∈ AX (δ) consider (x∗, x) ∈ AX∗ (δ) (we identify X as a subspace of X∗∗) and so we may find (y∗, y∗∗) ∈ Π(Y ∗) such that Now, an application of the previous lemma gives us a (y, y∗) ∈ Π(X) such that (cid:107)x∗ − y∗(cid:107) < ε and (cid:107)x − y∗∗(cid:107) < ε. (cid:107)x − y(cid:107) < ε and(cid:107)x∗ − y∗(cid:107) < ε. This means that ΦX (δ) (cid:54) ε and, therefore, ΦX (δ) (cid:54) ΦX∗ (δ), as desired. (cid:3) We do not know whether the inequalities in Proposition 3.6 can be strict. Of course, this can not be the case when the space is reflexive. Corollary 3.7. For every reflexive Banach space X, one has ΦX (δ) = ΦX∗ (δ) and ΦS every 0 < δ < 2. X (δ) = ΦS X∗ (δ) for Our last result in this section states that when the Bishop-Phelps-Bollob´as modulus is the worst possible, then the spherical Bishop-Phelps-Bollob´as modulus is also the worst possible. Proposition 3.8. Let X be a Banach space. For every δ ∈ (0, 2), the condition ΦX (δ) = equivalent to the condition ΦS √ 2δ. √ 2δ is 2δ(cid:3) ⇒(cid:2)ΦX (δ) = √ 2δ(cid:3) is evident. Let us n) ∈ BX × BX∗ 2δ. Then there is a sequence of pairs (xn, x∗ X (δ) = 2δ, the implication(cid:2)ΦS Proof. Since ΦS prove the inverse implication. Let ΦX (δ) = such that Re x∗ X (δ) (cid:54) ΦX (δ) (cid:54) √ n(xn) > 1 − δ but for every (y, y∗) ∈ Π(X) we have n − y∗(cid:107) (cid:62) √ (cid:107)xn − y(cid:107) (cid:62) √ or (cid:107)x∗ X (δ) = √ √ . 2δ − 1 n An application of Remark 2.3 gives us that (cid:107)x∗ n(cid:107) −→ 1 as n → ∞. As the duality argument given in Lemma 3.5 implies the dual version of Remark 2.3, we also have (cid:107)xn(cid:107) −→ 1 as n → ∞. Denote xn = xn(cid:107)xn(cid:107) , n = x∗ n(cid:107) . In the case when δ ∈ (0, 1], we have Re x∗ x∗ n(cid:107)x∗ 2δ − 1 n (cid:107)xn − y(cid:107) (cid:62) √ n − y∗(cid:107) (cid:62) √ √ 2δ, we get the condition ΦS In the case of δ ∈ (1, 2), we no longer know that Re x∗ n(xn) > 1 − δ, but what we do know is that 2δ thanks to the continuity (cid:3) n(xn) (cid:62) 1 − δ, and that gives us the desired condition ΦS lim inf Re x∗ of the spherical modulus (Proposition 3.1). Since the right-hand sides of the above inequalities go to − (cid:107)xn − xn(cid:107) n(xn) > 1 − δ but for every (y, y∗) ∈ Π(X) or (cid:107)x∗ − (cid:107)x∗ 2δ − 1 n 2δ − 1 n n − x∗ n(cid:107). X (δ) = X (δ) = √ √ 2δ. BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 11 4. Examples We start with the simplest example of X = R. (cid:40) Example 4.1. ΦR(δ) = √ δ δ − 1 + 1 if 0 < δ (cid:54) 1 if 1 < δ < 2 , ΦSR(δ) = 0 for every δ ∈ (0, 2). Proof. We first fix δ ∈ (0, 1]. First observe that taking x = 1 − δ, x∗ = 1, it is evident that ΦR(δ) (cid:62) δ. For the other inequality, we fix x, x∗ ∈ [−1, 1] with x∗x > 1 − δ. Then, x and x∗ have the same sign and we have that x > 1− δ and x∗ > 1− δ. Indeed, if x < 1− δ, as x∗ (cid:54) 1, one has x∗x = x∗x < 1− δ, a contradiction; the other inequality follows in the same manner. Finally, one deduces that x−sign(x) < δ and x∗ − sign(x∗) < δ, as desired. δ − 1 + 1 and x∗ − 1 = Second, fix δ ∈ (1, 2). On the one hand, taking x = δ − 1, one has x∗x = 1 − δ. δ − 1 + 1. For the other As x + 1 = inequality, we fix x, x∗ ∈ [−1, 1] with x∗x > 1 − δ. If x and x∗ have the same sign, which we may and do suppose positive, then x − 1 (cid:54) 1 < δ and x∗ − 1 (cid:54) 1 < δ and the same is true if one of them is null. Therefore, to prove the last case we may and do suppose that x > 0 and x∗ < 0. Now, if we suppose, for the sake of contradiction, that δ − 1 + 1, it follows that ΦR(δ) (cid:62) √ δ − 1, x∗ = −√ √ √ √ x − (−1) (cid:62) √ δ − 1 + 1 and x∗ − 1 (cid:62) √ δ − 1 + 1, δ − 1, so −x∗x (cid:62) δ − 1 or, equivalently, x∗x (cid:54) 1 − δ, a contradiction. δ − 1 + 1 and x∗ − (−1) < 1 < δ − 1 + 1 and δ − 1 + 1 or x∗ − 1 < √ √ δ − 1 and −x∗ (cid:62) √ we get x (cid:62) √ √ Therefore, either x − (−1) < √ δ − 1 + 1. x − 1 < 1 < The result for ΦSR is an obvious consequence of the fact that SR = {−1, 1}. (cid:3) Let us observe that the above proof gives actually a lower bound for ΦX (δ) for every Banach space X when δ ∈ (0, 1]. Remark 4.2. Let X be a Banach space. Then ΦX (δ) (cid:62) δ for every δ ∈ (0, 1]. Indeed, consider x0 ∈ SX 0. Then Re x∗(x) = 1 − δ and and x∗ dist (x, SX ) = δ. 0(x0) = 1 and write x = (1 − δ)x0 and x∗ = x∗ 0 ∈ SX∗ with x∗ We do not know a result giving a lower bound for ΦX (δ) when δ > 1, outside of the trivial one ΦX (δ) (cid:62) 1. Also, we do not know if the lower bound for the behavior of ΦX (δ) in a neighborhood of 0 given in the remark above can be improved for Banach spaces of dimension greater than or equal to two. We next calculate the moduli of a Hilbert space of (real) dimension greater than one. Example 4.3. Let H be a Hilbert space of dimension over R greater than or equal to two. Then: H (δ) = (a) ΦS (b) For δ ∈ (0, 1], ΦH (δ) = max 4 − 2δ . For δ ∈ (1, 2), ΦH (δ) = √ δ. 4 − 2δ for every δ ∈ (0, 2). (cid:110) δ, (cid:112) 2 − √ (cid:111) (cid:112) 2 − √ Proof. As we commented in the introduction, both ΦH and ΦS H only depend on the real structure of the space, so we may and do suppose that H is a real Hilbert space of dimension greater than or equal to 2. Let us also recall that H∗ identifies with H and that the action of a vector y ∈ H on a vector x ∈ H is nothing but their inner product denoted by (cid:104)x, y(cid:105). In particular, Π(H) =(cid:8)(z, z) ∈ SH × SH (cid:9). Therefore, for every δ ∈ (0, 2), ΦH (δ) (resp. ΦS H (δ)) is the infimum of those ε > 0 such that whenever x, y ∈ BH (resp. x, y ∈ SH ) satisfies (cid:104)x, y(cid:105) (cid:62) 1− δ, there is z ∈ SH such that (cid:107)x− z(cid:107) (cid:54) ε and (cid:107)y− z(cid:107) (cid:54) ε. We will use the following (easy) claim in both the proofs of (a) and (b). (cid:107)x − z(cid:107) = (cid:107)y − z(cid:107) = Indeed, we have (cid:107)x − z(cid:107)2 = 2 − 2(cid:104)x, z(cid:105) and 2(cid:104)x, z(cid:105) = (cid:113) 2 −(cid:112)2 + 2(cid:104)x, y(cid:105), being the other equality true by symmetry. 2(cid:104)x, x + y(cid:105) (cid:107)x + y(cid:107) = giving (cid:107)x − z(cid:107) = (cid:113) 2 −(cid:112)2 + 2(cid:104)x, y(cid:105). (cid:112)2 + 2(cid:104)x, y(cid:105) , 2 + 2(cid:104)x, y(cid:105) 12 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO Claim: Given x, y ∈ SH with x + y (cid:54)= 0, write z = x+y (cid:107)x+y(cid:107) to denote the normalized midpoint. Then H (δ) (cid:54)(cid:112) 2 − √ (cid:113) 2 −(cid:112)2 + 2(cid:104)x, y(cid:105) (cid:54) (cid:113) 2 − √ (a). Let first prove that ΦS consider z = x+y (cid:107)x+y(cid:107) ∈ SH and use the claim to get that 4 − 2δ. Take x, y ∈ SH with (cid:104)x, y(cid:105) (cid:62) 1 − δ (so x + y (cid:54)= 0), and observe that (cid:104)x, y(cid:105) = 1 − δ. Now, given z ∈ SH , we write z1 = (cid:104)z, e1(cid:105), z2 = (cid:104)z, e2(cid:105), and observe that (cid:107)x − z(cid:107) = (cid:107)y − z(cid:107) = 4 − 2δ. To get the other inequality, we fix an ortonormal basis {e1, e2, . . .} of H, consider x =(cid:112)1 − δ/2 e1 +(cid:112)δ/2e2 ∈ SH and y =(cid:112)1 − δ/2 e1 −(cid:112)δ/2e2 ∈ SH (cid:111) (cid:112)1 − δ/2 + max± z2 ±(cid:112)δ/22 + 1 − z2 1 + 1 − δ/2 − 2z1 (cid:110) (cid:112) 2 − √ (cid:110)z1 −(cid:112)1 − δ/22 + z2 ±(cid:112)δ/22 + 1 − z2 (cid:112)1 − δ/2 + 2z2(cid:112)δ/2 (cid:62) 2 − 2(cid:112)1 − δ/2. max{(cid:107)z − x(cid:107)2,(cid:107)z − y(cid:107)2} = max± = z2 = 2 − 2z1 4 − 2δ, as desired. (b). We first fix δ ∈ (0, 1) and write ε0 = max H (δ) (cid:62)(cid:112) It follows that ΦS 2 − √ 1 − z2 4 − 2δ (cid:111) δ, 2 . The inequality ΦH (δ) (cid:62) ε0 follows H (δ) and the result in item (a). To get the other inequality, we by Remark 4.2, the fact that ΦH (δ) (cid:62) ΦS first observe that 1 − z2 2 (5) ΦH (δ) (cid:54) ΦLin {x,y}(δ) ∀x, y ∈ BH with (cid:104)x, y(cid:105) = 1 − δ. , √ 2(cid:107)P(cid:107) 2(cid:107)P(cid:107) (cid:17) Consider M = , which is the normalized midpoint between A = (1, 0) and Figure 1 helps to the better understanding of the rest of the proof. 1 − δ, 1(cid:3), and q1 = 1−δ(cid:107)P(cid:107) ∈ (cid:2)1 − δ, This follows from the obvious fact that Φ·(δ) increases when we restrict to subspaces. So, we are done if we restrict to the two-dimensional case and consider two points P = ((cid:107)P(cid:107), 0), Q = (q1, q2) with q2 (cid:62) 0 and (cid:107)P(cid:107) (cid:62) (cid:107)Q(cid:107), satisfying (cid:104)P, Q(cid:105) (cid:62) 1− δ, and we find z ∈ SH such that (cid:107)P − z(cid:107) (cid:54) ε0 and (cid:107)Q− z(cid:107) (cid:54) ε0. Now, it is straightforward to check that we have (cid:107)P(cid:107) ∈ (cid:2)√ 1 − δ(cid:3). (cid:113)(cid:107)P(cid:107)−(1−δ) (cid:16)(cid:113) 1−δ+(cid:107)P(cid:107) (cid:113) 1 − ( 1−δ(cid:107)P(cid:107) )2(cid:1) and write ∆ to denote the arc of the unit sphere of H between A and M . We B =(cid:0) 1−δ(cid:107)P(cid:107) , claim that Q ∈(cid:83) z∈∆ B(z, ε0) and P ∈(cid:84) z∈∆ B(z, ε0). If q2 (cid:54)(cid:113)(cid:107)P(cid:107)−(1−δ) z∈∆ B(z, ε0). Observe that this gives that there is z ∈ ∆ ⊂ SH show that Q = (q1, q2) ∈(cid:83) whose distance to P and Q is less than or equal to ε0, finishing the proof. Let us prove the claim. First, we (cid:113)(cid:107)P(cid:107)−(1−δ) (cid:16) (cid:17) , the ball of radius ε0 centered in the point of ∆ with second coordinate equal to q2 contains the point Q since ε0 (cid:62) dist ((q1, 0), A) (cid:62) dist (Q, ∆). For greater values of q2, write first C = , which belongs to B(M, ε0) by the previous (cid:118)(cid:117)(cid:117)(cid:116)2 − argument. Also, as M is the normalized mid point between A and B, we have by the claim at the (cid:115) beginning of this proof that 2 −(cid:112)2 + 2(cid:104)A, B(cid:105) = (cid:113) 2 − √ (cid:107)M − B(cid:107) = 4 − 2δ (cid:54) ε0 (cid:113) 2(cid:107)P(cid:107) 2(cid:107)P(cid:107) 2 + 2 q1, 1 − δ (cid:107)P(cid:107) (cid:54) BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 13 B D C M 1 − δ 1−δ(cid:107)P (cid:107) q1 √ 1 − δ P A Figure 1. Calculating ΦH (δ) for δ ∈ (0, 1) so, also, (cid:107)M − D(cid:107) (cid:54) ε0. Therefore, both the points C and C belong to B(M, ε0), so also the whole segment [C, D] is contained there, and this proves the first part of the claim. To show the second part of the claim, that P ∈(cid:84) z∈∆ B(z, ε0), we consider the function f (p) := 1 + p2 −(cid:112)2p(p + 1 − δ) (cid:0)p ∈ [ 1 − δ, 1](cid:1) √ and observe that it is a convex function, so √ f (p) (cid:54) max{f (1), f ( (cid:113) 1 − δ} (cid:54) ε2 0. 1 + (cid:107)P(cid:107)2 −(cid:112)2(cid:107)P(cid:107)((cid:107)P(cid:107) + 1 − δ) (cid:54) ε0, It follows that (cid:107)P − M(cid:107) = z∈∆ B(z, ε0). B(P, ε0) or, equivalently, that P ∈(cid:84) hence M ∈ B(P, ε0). As also A ∈ B(P, ε0), it follows that the whole circular arc ∆ is contained in Let now fix δ ∈ (1, 2). Analogously to what we did before in equation (5), to show that ΦH (δ) (cid:54) √ δ, it is enough to consider the two-dimensional case and that, given p = ((cid:107)p(cid:107), 0) ∈ BH , q = (q1, q2) ∈ BH with q2 (cid:62) 0, to find z ∈ SH such that (cid:107)z − P(cid:107),(cid:107)z − Q(cid:107) (cid:54) √ 1 −(cid:0)(cid:107)p(cid:107) + q1 Q = −√ does the job. For the other inequality, we fix an ortonormal basis {e1, e2, . . .} of H, consider (cid:32)(cid:107)p(cid:107) + q1 δ. Routine computations show that (cid:33) (cid:1)2 δ − 1 e1 ∈ BH , δ − 1 e1 ∈ BH ∈ SH P = (cid:114) , z = √ 2 2 14 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO and observe that (cid:104)P, Q(cid:105) = 1 − δ. For any z ∈ SH , we write z1 = (cid:104)z, e1(cid:105) and we compute max{(cid:107)z − P(cid:107)2,(cid:107)z − Q(cid:107)2} = max (cid:110)z1 − √ = max± z1 ± √ √ δ − 1z1 (cid:62) δ. δ − 12 + 1 − z12 , z1 + δ − 12 + 1 − z12 = (z1 + δ − 12 + 1 − z12(cid:111) √ √ δ − 1)2 + 1 − z12 (cid:3) It follows that ΦH (δ) (cid:62) √ = δ + 2 δ, as desired. X (δ) (cid:54) ΦX (δ) (cid:54) √ Our next aim is to present a number of examples for which the values of the Bishop-Phelps-Bollob´as moduli are the maximum possible, namely ΦS 2δ for small δ's. As we always have ΦS 2δ for small δ's (actually, the two facts are equivalent, see Proposition 3.8), and this is what we will show. It happens that all of the examples have in common that they contains an isometric copy of the real space (cid:96)(2)∞ or (cid:96)(2) 1 . In the next section we will show that the latter is a necessary condition that it is not actually sufficient. 2δ, it is enough if we prove the formally stronger result that ΦS X (δ) = ΦX (δ) = X (δ) = √ √ The first result is about Banach spaces admitting an L-descomposition. As a consequence we will calculate the moduli of L1(µ) spaces. Proposition 4.4. Let X be a Banach space. Suppose that there are two (non-trivial) subspaces Y and Z such that X = Y ⊕1 Z. Then ΦX (δ) = ΦS Proof. Fix δ ∈ (0, 1/2] and consider (y0, y∗ 2δ for every δ ∈ (0, 1/2]. √ (cid:32)√ (cid:16) x0 = 2δ 2 y0 , 1 − X (δ) = (cid:33) 0) ∈ Π(Y ) and (z0, z∗ (cid:17) ∈ SX z0 x∗ 0 = √ 2δ 2 (cid:16)(cid:0)1 − 0 ) ∈ Π(Z) and write 0 , z∗ 2δ(cid:1)y∗ (cid:17) ∈ SX∗ . √ 0 It is clear that Re x∗ 0(x0) = 1 − δ. Now, suppose that we may choose (x, x∗) ∈ Π(X) such that √ (cid:107)x0 − x(cid:107) < 2δ and (cid:107)x∗ 0 − x∗(cid:107) < √ 2δ. Write x = (y, z) ∈ Y ⊕1 Z, x∗ = (y∗, z∗) ∈ Y ∗ ⊕∞ Z∗ and observe that 1 = Re x∗(x) = Re y∗(y) + Re z∗(z) (cid:54) (cid:107)y∗(cid:107)(cid:107)y(cid:107) + (cid:107)z∗(cid:107)(cid:107)z(cid:107) (cid:54) (cid:107)y(cid:107) + (cid:107)z(cid:107) = 1, therefore, we have (6) Now, we have √ (cid:12)(cid:12)(cid:12)(cid:0)1 − (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) + y0 Re y∗(y) = (cid:107)y∗(cid:107)(cid:107)y(cid:107). √ 2δ(cid:1) − (cid:107)y∗(cid:107)(cid:12)(cid:12)(cid:12) (cid:54)(cid:13)(cid:13)(cid:13)(cid:0)1 − 0 − y∗(cid:13)(cid:13)(cid:13) < 2δ(cid:1)y∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:16) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:62) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16) z0 − z 2δ 2 1 − √ + (cid:17) a contradiction. We have proved that ΦX (δ) (cid:62) √ (cid:107)x0 − x(cid:107) = 2δ 2 2δ 2 1 − √ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)√ from which follows that (cid:107)y∗(cid:107) < 1 and so, y = 0 by (6), giving (cid:107)z(cid:107) = (cid:107)x(cid:107) = 1. But then, √ 2δ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:17) − (cid:107)z(cid:107) √ 2δ 2 √ 2δ, 2δ, being the other inequality always true. (cid:3) The result above produces the following example. Example 4.5. Let (Ω, Σ, µ) be a measure space such that L1(µ) has dimension greater than one and let E be any non-zero Banach space. Then, ΦL1(µ,E)(δ) = ΦS Indeed, we may find two measurable sets A, B ⊂ Ω with empty intersection such that Ω = A ∪ B. Then Y = L1(µA, E) and Z = L1(µB, E) are non-null, L1(µ, E) = Y ⊕1 Z and so the results follows from Proposition 4.4. 2δ for every δ ∈ (0, 1/2]. L1(µ,E)(δ) = √ so z∗ = 0 and x∗ ∈ Y . We now define x∗ 0 = x∗(x0) = It is clear that Re x∗ 2δ 2 (cid:32)√ (cid:12)(cid:12)(cid:12)(cid:0)1 − y∗ 0 , √ 1 − (cid:33) (cid:17) (cid:16) (cid:12)(cid:12)(cid:12) (cid:54)(cid:0)1 − 2δ(cid:1)y∗(y0) + z∗(z0) ∈ SX∗ 2δ 2 z∗ 0 √ √ 2δ(cid:1)y0 + z0 ∈ X x0 =(cid:0)1 − 2δ(cid:1)(cid:107)y∗(cid:107) + (cid:107)z∗(cid:107) (cid:54) (cid:107)y∗(cid:107) + (cid:107)z∗(cid:107) = 1. √ 0(x0) = 1 − δ. Now, suppose that we may choose (x, x∗) ∈ Π(X) such that √ (cid:107)x0 − x(cid:107) < 2δ and (cid:107)x∗ 0 − x∗(cid:107) < √ 2δ. and first observe that (cid:107)x0(cid:107) (cid:54) 1; indeed, for every x∗ = y∗ + z∗ ∈ SX∗ one has BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 15 Particular case of the above example are (cid:96)1 and L1[0, 1]. It is immediate that with a dual argument than the one given in Proposition 4.4 it is possible to deduce the same for a Banach space which decomposes as an (cid:96)∞-sum. Actually, in this case we will get a better result using ideals instead of subspaces. Proposition 4.6. Let X be a Banach space. Suppose that X∗ = Y ⊕1 Z where Y and Z are (non- w∗ (cid:54)= X∗ (w∗ is the weak∗-topology σ(X∗, X)). Then trivial) subspaces of X∗ such that Y ΦX (δ) = ΦS Proof. We claim that there are y0, z0 ∈ SX and y∗ 2δ for every δ ∈ (0, 1/2]. w∗ (cid:54)= X∗ and Z 0 ∈ SZ such that X (δ) = √ 0 ∈ SY and z∗ y∗(z0) = 0 ∀y∗ ∈ Y, z∗(y0) = 0 ∀z∗ ∈ Z. 0 analogous. By assumption there is y0 ∈ SX such that Indeed, we define y0 and y∗ z∗(y0) = 0 for every z∗ ∈ Z and we may choose x∗ ∈ SX∗ such that Re x∗(y0) = 1 and we only have to prove that x∗ ∈ Y and then write y∗ 0 = x∗. But we have x∗ = y∗ + z∗ with y∗ ∈ Y , z∗ ∈ Z and Re y∗ 0(y0) = 1, Re z∗ 0 (z0) = 1, 0, being z0 and z∗ 1 = Re x∗(y0) = Re y∗(y0) (cid:54) (cid:107)y∗(cid:107) (cid:54) (cid:107)y∗(cid:107) + (cid:107)z∗(cid:107) = 1, We consider the semi-norm (cid:107)·(cid:107)Y defined on X by (cid:107)x(cid:107)Y := sup{y∗(x) : y∗ ∈ SY } which is smaller than or equal to the original norm, write x∗ = y∗ + z∗ with y∗ ∈ Y and z∗ ∈ Z, and observe that 1 = Re x∗(x) = Re y∗(x) + Re z∗(x) (cid:54) (cid:107)y∗(cid:107)(cid:107)x(cid:107)Y + (cid:107)z∗(cid:107)(cid:107)x(cid:107) (cid:54) (cid:107)y∗(cid:107) + (cid:107)z∗(cid:107) = 1. Therefore, we have, in particular, that (7) Now, we have(cid:12)(cid:12)(cid:12)(cid:0)1 − √ < from which follows that (cid:107)x(cid:107)Y < 1 and so, y∗ = 0 by (7) and (cid:107)z∗(cid:107) = (cid:107)x∗(cid:107) = 1. But then, 2δ (cid:12)(cid:12)(cid:12) = 2δ(cid:1) − (cid:107)x(cid:107)Y (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)√ 2δ 2 y∗ (cid:12)(cid:12)(cid:12)(cid:0)1 − (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16) Re y∗(x) = (cid:107)y∗(cid:107)(cid:107)x(cid:107)Y . √ 2δ(cid:1)(cid:107)y0(cid:107)Y − (cid:107)x(cid:107)Y (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:62) (cid:17) (cid:12)(cid:12)(cid:12) (cid:54)(cid:13)(cid:13)(cid:13)(cid:0)1 − (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:16) 2δ 2 √ √ + √ (cid:13)(cid:13)(cid:13)Y 2δ(cid:1)y0 − x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:17) − (cid:107)z∗(cid:107) 2δ 2 √ √ 1 − (cid:107)x∗ 0 − x∗(cid:107) = 0 − z∗ z∗ a contradiction. Again, we have proved that ΦX (δ) (cid:62) √ 2δ 2 1 − 0 √ 2δ, 2δ, being the other inequality always true. (cid:3) Of course, the first consequence of the above result is to Banach spaces which decompose as (cid:96)∞-sum of two subspaces. Indeed, if X = Y ⊕∞ Z for two (non-trivial) subspaces Y and Z, then X∗ = Y ⊥ ⊕1 Z⊥ and Y ⊥ and Z⊥ are w∗-closed, so far away of being dense. Therefore, Proposition 4.6 applies. We have proved the following result. Corollary 4.7. Let X be a Banach space. Suppose that there are two (non-trivial) subspaces Y and Z such that X = Y ⊕∞ Z. Then ΦX (δ) = ΦS 2δ for every δ ∈ (0, 1/2]. X (δ) = √ As a consequence, we obtain the following examples, analogous to the ones presented in Example 4.5. 16 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO Examples 4.8. (a) Let (Ω, Σ, µ) a measure space such that L∞(Ω) has dimension greater than one and let E be any non-zero Banach space. Then, √ ΦL∞(µ,E) = ΦS L∞(µ,E)(δ) = 2δ (cid:0)δ ∈ (0, 1/2](cid:1). √ (cid:0)δ ∈ (0, 1/2](cid:1). (b) Let Γ be a set with more than one point and let E be any non-zero Banach space. Then, √ Φc0(Γ,E) = ΦS c0(Γ,E)(δ) = 2δ and Φc(Γ,E) = ΦS c(Γ,E)(δ) = 2δ Our next aim is to deduce from Proposition 4.6 that also arbitrary C(K) spaces have the maximum moduli and for this we have to deal with the concept of M -ideal. Given a subspace J of a Banach space X, J is called M -ideal if J⊥ is a L-summand on X∗ (use [10] for background). In this case, X∗ = J⊥ ⊕1J (cid:93) where J (cid:93) = {x∗ ∈ X∗ : (cid:107)x∗(cid:107) = (cid:107)x∗J(cid:107)} ≡ J∗. Now, if X contain a non-trivial M -ideal J, one has X∗ = J⊥ ⊕1 J (cid:93) and to apply Proposition 4.6 we need that J (cid:93) to be not σ(X∗, X)-dense. Actually, J (cid:93) is not dense in X∗ if and only if there is x0 ∈ X \{0} such that (cid:107)x0 + y(cid:107) = max{(cid:107)x0(cid:107),(cid:107)y(cid:107)} for every y ∈ J (this is easy to verify and a proof can be found in [3]). Let us enunciate what we have shown. Corollary 4.9. Let X be a Banach space. Suppose that there is a non-trivial M -ideal J of X and a point x0 ∈ X \ {0} such that (cid:107)x0 + y(cid:107) = max{(cid:107)x0(cid:107),(cid:107)y(cid:107)} for every y ∈ J. Then, ΦX (δ) = ΦS 2δ for every δ ∈ (0, 1/2]. X (δ) = √ With the above corollary we are able to prove that the moduli of any non-trivial C0(L) space are maximum. Example 4.10. Let L be a locally compact Hausdorff topological space with at least two points and let E be any non-zero Banach space. Then ΦC0(L,E)(δ) = ΦS Indeed, we may find a non-empty non-dense open subset U of L and consider the subspace 2δ for every δ ∈ (0, 1/2]. C0(L,E)(δ) = √ J =(cid:8)f ∈ C0(L, E) : fU = 0}, which is an M -ideal of C0(L, E) by [10, Corollary VI.3.4] (use the simpler [10, Example I.1.4.a] for the scalar-valued case) and it is non-zero since L \ U has non-empty interior. As U is open and non- empty, we may find a non-null function x0 ∈ C0(L, E) whose support is contained in U . It follows that (cid:107)x0 + y(cid:107) = max{(cid:107)x0(cid:107),(cid:107)y(cid:107)} for every y ∈ J by disjointness of the supports. A sufficient condition to be in the hypotheses of Corollary 4.9 is that a Banach space X contains two non-trivial M -ideals J1 and J2 such that J1 ∩ J2 = {0} since, in this case, J1 and J2 are complementary M -summands in J1 + J2 [10, Proposition I.1.17]. Let us comment that this is actually what happens in C(K) when K has more than one point. Corollary 4.11. Let X be a Banach space. Suppose there are two non-trivial M -ideals J1 and J2 such that J1 ∩ J2 = {0}. Then ΦX (δ) = ΦS 2δ for every δ ∈ (0, 1/2]. X (δ) = √ of (cid:81) A sufficient condition for a Banach space to have two non-intersecting M -ideals is that its centralizer is non-trivial (i.e. has dimension at least two). We are not going into details, but roughly speaking, the centralizer Z(X) of a Banach space X is a closed subalgebra of L(X) isometrically isomorphic to C(KX ) where KX is a Hausdorff topological space, and it is possible to see X as a C(KX )-submodule Xk for suitable Xk's. We refer to [2, §3.B] and [10, §I.3] for details. It happens that every M -ideal of C(KX ) produces an M -ideal of X in a suitable way (see [2, §4.A]) and if Z(X) contains more than one point, then two non-intersecting M -ideals appear in X, so our corollary above applies. √ Corollary 4.12. Let X a Banach space. If Z(X) has dimension greater than one, then ΦX (δ) = ΦS 2δ for every δ ∈ (0, 2]. X (δ) = k∈KX To give some new examples coming from this corollary, we recall that the centralizer of a unital (complex) C∗-algebra identifies with its center (see [10, Theorem V.4.7] or [2, Example 3 in page 63]). Example 4.13. Let A be a unital C∗-algebra with non-trivial center. Then, ΦA(δ) = ΦS every δ ∈ (0, 1/2]. A(δ) = 2δ for √ BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 17 It would be interesting to see whether the algebra L(H) for a finite- or infinite-dimensional Hilbert space H has the maximum Bishop-Phelps-Bollob´as moduli. None of the results of this section applies to it since its center is trivial and, despite it contains K(H) as an M -ideal, there is no element x0 ∈ L(H) satisfying the requirements of Corollary 4.9 (see [3, page 538]). Let us also comment that the bidual of 2δ for every δ ∈ (0, 1/2]. L(H) is a C∗-algebra with non-trivial centralizer, so ΦL(H)∗∗ (δ) = ΦS If there is δ ∈ (0, 1/2] such that ΦL(H)(δ) < 2δ, then this would be an example when the inequality in Proposition 3.6 is strict. L(H)∗∗ (δ) = √ √ We finish this section with two pictures: one with the Bishop-Phelps-Bollob´as moduli of R, C and (cid:96)(2)∞ , and another one with the corresponding values of the spherical Bishop-Phelps-Bollob´as moduli. 2 1 2 1 0 1 2 0 1 2 Figure 2. The value of ΦX (δ) for R, C and (cid:96)(2)∞ Figure 3. The value of ΦS for R, C and (cid:96)(2)∞ X (δ) 5. Banach spaces with the greatest possible modulus Our goal in this section is to show that Banach spaces with the greatest possible moduli contain almost isometric copies of the real (cid:96)2∞. Let us first recall the following definition. Definition 5.1. Let X, E be Banach spaces. X is said to contain almost isometric copies of E if, for every ε > 0 there is a subspace Eε ⊂ X and there is a bijective linear operator T : E −→ Eε with (cid:107)T(cid:107) < 1 + ε and (cid:107)T −1(cid:107) < 1 + ε. The next result is well-known and has a straightforward proof. Lemma 5.2. A real Banach space E contains an isometric copy of (cid:96)(2)∞ if and only if there are elements u, v ∈ SE such that (cid:107)u− v(cid:107) = (cid:107)u + v(cid:107) = 2. E contains almost isometric copies of (cid:96)(2)∞ if and only if there are elements un, vn ∈ SE, n ∈ N such that (cid:107)un − vn(cid:107) −→ 2 and (cid:107)un + vn(cid:107) −→ 2 as n → ∞. The class of spaces X that do not contain almost isometric copies of (cid:96)(2)∞ was deeply studied by James [11] (see also the exposition in Van Dulst's book [9]), who gave to such spaces the name "uniformly non-square". He proved in particular, that every uniformly non-square space must be reflexive, that this property is stable under passing to subspaces, quotient spaces and duals. In fact, a general result is true [12]: for every 2-dimensional space E if a real Banach space X does not contain almost isometric copies of E then X is reflexive. 18 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO √ The aim of this section is to prove that if a real Banach space X satisfies that its Bishop-Phelps- 2δ in at least one point δ ∈ (0, 1/2), then X (and, equivalently, the dual space) 2δ if and only if Bollob´as modulus is √ contains almost isometric copies of (cid:96)(2)∞ . Actually, as shown in Remark 3.8, ΦX (δ) = ΦS X (δ) = We will use some lemmas and ideas of Bishop and Phelps [4], but for the reader's convenience we will 2δ. Therefore, we may use the formally stronger hypothesis of ΦS X (δ) = √ √ 2δ. refer to the corresponding lemmas in the already classical Diestel's book [8]. From now on, X will denote a real Banach space. For t > 1 and x∗ ∈ SX∗ , we denote K(t, x∗) := {x ∈ X : (cid:107)x(cid:107) (cid:54) t x∗(x)}. Observe that K(t, x∗) is a convex cone with non-empty interior. Lemma 5.3 (A particular case of [8, Chapter 1, Lemma 1]). For every z ∈ BX , every x∗ ∈ SX∗ and every t > 1, there is x0 ∈ SX such that x0 − z ∈ K(t, x∗) and [K(t, x∗) + x0] ∩ BX = {x0}. Lemma 5.4 ([8, Chapter 1, Lemma 2] with a little modification that follows from the proof there). Let x∗, y∗ ∈ SX∗ and suppose that x∗(cid:0)ker y∗ ∩ SX dist(cid:0)x∗, Lin y∗(cid:1) (cid:54) ε/2 a functional that separates x0 + K(t, x∗) from BX , so y∗(x0) = 1 and y∗(cid:0)K(t, x∗)(cid:1) ⊂ [0,∞). Then (cid:1) ⊂ (−∞, 1/t] and so, dist(cid:0)x∗, Lin y∗(cid:1) (cid:54) 1/t and min{(cid:107)x∗ − y∗(cid:107),(cid:107)x∗ + y∗(cid:107)} (cid:54) 2/t. x∗(cid:0)ker y∗ ∩ SX Lemma 5.5. Let z ∈ BX , x∗ ∈ SX∗ , t > 1, and let x0 ∈ SX be from Lemma 5.3. Denote y∗ ∈ SX∗ (cid:1) ⊂ (−∞, ε/2]. Then min{(cid:107)x∗ − y∗(cid:107),(cid:107)x∗ + y∗(cid:107)} (cid:54) ε. and Proof. This also can be extracted from [8, Chapter 1], but it is better to give a proof. For every w ∈ ker y∗ ∩ SX we have that w does not belong to the interior of K(t, x∗), so 1 = (cid:107)w(cid:107) (cid:62) t x∗(w), i.e. x∗(ker y∗ ∩ SX ) ⊂ (−∞, 1/t]. An application of Lemma 5.4 completes the proof. (cid:3) Now we are passing to our results. At first, for the sake of simplicity, we consider the easier finite- dimensional case. Lemma 5.6. Let X be a finite-dimensional real space. Fix ε ∈ (0, 1). Suppose that (x, x∗) ∈ SX × SX∗ satisfies that x∗(x) = 1 − ε2 2 and that max{(cid:107)y − x(cid:107),(cid:107)y∗ − x∗(cid:107)} (cid:62) ε for every pair (y, y∗) ∈ Π(X). Then for t = 2 ε , there exists y0 ∈ [x + K(t, x∗)]∩ SX such that x∗(y0) = 1. Proof. Consider a sequence tn > t, n ∈ N, with limn tn = t. Using Lemma 5.3, we get yn ∈ SX such that (8) Let y∗ [0,∞). Then, according to Lemma 5.5, (9) n ∈ X∗ be a functional that separates K(tn, x∗) + yn from BX , i.e. y∗ (K(tn, x∗) + yn) ∩ BX = {yn}. yn − x ∈ K(tn, x∗) n(yn) = 1 and y∗ n(K(tn, x∗)) ⊂ min{(cid:107)x∗ − y∗ n(cid:107)} (cid:54) 2/tn < ε. n(cid:107),(cid:107)x∗ + y∗ and But (cid:107)x∗ + y∗ n(cid:107) (cid:62) (x∗ + y∗ n)(yn) = 1 + x∗(yn) = 1 + x∗(x) + x∗(yn − x) = 2 − ε2 2 Since (yn − x) ∈ K(tn, x∗), we have x∗(yn − x) (cid:62) (cid:107)(yn − x)(cid:107)/tn (cid:62) 0 so + x∗(yn − x). (we have used here that 0 < ε < 1). Comparing with (9), we get (cid:107)x∗ − y∗ n(cid:107) < ε, so the condition of our lemma says that (cid:107)x − yn(cid:107) (cid:62) ε. Without loss of generality (passing to a subsequence if necessary) we can (cid:107)x∗ + y∗ n(cid:107) (cid:62) 2 − ε2 2 > ε BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 19 assume that yn tend to some y0. Then ε (cid:54) lim n (cid:54) 2 ε (cid:107)yn − x(cid:107) (cid:54) lim n ε2 (x∗(y0) − 1 + 2 tnx∗(yn − x) = t(x∗(y0) − x∗(x)) ) (cid:54) 2 ε (1 − 1 + ) = ε. ε2 2 This means that all the inequalities in the above chain are in fact equalities. In particular, x∗(y0) = 1 and (cid:107)yn − x(cid:107) = t(cid:0)x∗(y0) − x∗(x)(cid:1), (cid:107)y0 − x(cid:107) = lim n i.e. y0 ∈ [x + K(t, x∗)] ∩ SX . Lemma 5.7. Under the conditions of Lemma 5.6, there are y∗ ∈ SX∗ and α (cid:62) 1 − ε 2 with (cid:3) (10) and there is v ∈ SX such that (11) (cid:107)x∗ − αy∗(cid:107) (cid:54) ε 2 and (cid:107)x∗ − y∗(cid:107) (cid:62) ε, x∗(v) = y∗(v) = 1. Proof. Let y0 be from the previous lemma. Fix a strictly increasing sequence of tn > 1 with limn tn = t and let us consider two cases. Case 1 : Suppose there exists m0 ∈ N with int(cid:2)K(tm0, x∗) + x(cid:3) ∩ BX (cid:54)= ∅. Then, using the fact that for every closed convex set with non-empty interior, the closure of the interior is the whole set, we get int(cid:2)x + K(tn, x∗)(cid:3) ∩ BX . So, we can pick y0 ∈(cid:2)x + K(t, x∗)(cid:3) ∩ BX = int(cid:2)x + K(t, x∗)(cid:3) ∩ BX = zn ∈(cid:2)x + K(tn, x∗)(cid:3) ∩ BX (cid:2)K(tn, x∗) + vn vn − zn ∈ K(tn, x∗) (cid:91) n(cid:62)m0 (12) such that zn −→ y0. In particular, x∗(zn) −→ 1. Let us apply Lemma 5.3: there are vn ∈ SX such that (13) Then x∗(vn − zn) (cid:62) 0, i.e. 1 (cid:62) x∗(vn) (cid:62) x∗(zn) −→ 1, so x∗(vn) −→ 1. Condition (12) implies that zn − x ∈ K(tn, x∗) which, together with (13), mean that vn − x ∈ K(tn, x∗). Consequently, (cid:3) ∩ BX = {vn}. and If we denote y∗ Since we are working under the conditions of Lemma 5.6, it follows that n ∈ SX∗ to the functional that separates vn + K(tn, x∗) from BX , then (vn, y∗ n) ∈ Π(X). (cid:107)vn − x(cid:107) (cid:54) tnx∗(vn − x) (cid:54) tn ε2 2 < ε. Also, by Lemma 5.5, dist (x∗, Lin y∗ n) (cid:54) 1/tn, so there are αn ∈ R such that (cid:107)y∗ n − x∗(cid:107) (cid:62) ε. (cid:107)x∗ − αny∗ n(cid:107) (cid:54) 1/tn. Again, without loss of generality, we may assume that the sequences (αn), (vn) and (y∗ denote α := limn αn, y∗ := limn y∗ and y∗(v) = limn y∗ n) have limits. Let us n, and v := limn vn. Then (cid:107)v(cid:107) = 1, (cid:107)y∗(cid:107) = 1, x∗(v) = limn x∗(vn) = 1, n(vn) = 1. This proves (11). Also, Consequently, (14) so, α (cid:62) 1 − ε 2 . (cid:107)x∗ − αy∗(cid:107) = lim n (cid:107)x∗ − αny∗ n(cid:107) (cid:54) 1 t = ε 2 . (cid:62) (cid:107)x∗ − αy∗(cid:107) (cid:62) (x∗ − αy∗)(v) = 1 − α, ε 2 20 Case 2 : Assume that for every n ∈ N we have int(cid:2)K(tn, x∗) + x(cid:3) ∩ BX = ∅. Let us separate x + CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO int (K(tn, x∗)) from BX by a norm-one functional y∗ (cid:0)x + int(cid:2)K(tn, x∗)(cid:3)(cid:1) > 1, n, that is, y∗ n so, in particular, y∗ Again, passing to a subsequence, we can assume that there exists y∗ = limn y∗ n(x) (cid:62) 1. n which satisfies (cid:107)y∗(cid:107) = 1, n(x) (cid:62) 1. So, y∗(x) = 1, i.e. (x, y∗) ∈ Π(X). By the conditions of our lemma, this 1 (cid:62) y∗(x) (cid:62) limn y∗ implies that Since (cid:107)y∗ − x∗(cid:107) = max{(cid:107)x − x(cid:107),(cid:107)y∗ − x∗(cid:107)} (cid:62) ε. int(cid:2)x + K(tn, x∗)(cid:3), we can select zn ∈ int(cid:2)x + K(tn, x∗)(cid:3) in such a way that zn −→ y0. Then y0 ∈ x + K(t, x∗) = (cid:91) n∈N y∗(y0) = lim y∗ n(zn) (cid:62) 1, n hence, y∗(y0) = 1. This means that condition (11) works for v := y0. The remaining conditions can be (cid:3) deduced from Lemma 5.5 the same way as in the case 1. We are now able to state and prove the main result of the section in the finite-dimensional case. Theorem 5.8. Let X be a finite-dimensional real Banach space. Suppose that there is a δ ∈ (0, 1/2) 2δ). Then X∗ contains an isometric copy of (cid:96)(2)∞ X (δ) = such that ΦX (δ) = (hence, X also contains an isometric copy of (cid:96)(2)∞ ). 2δ (or, equivalently, ΦS √ √ 2δ ∈ (0, 1). There is a sequence of pairs (xn, x∗ n) ∈ SX × SX∗ such that x∗ n(xn) > √ Proof. Denote ε := 1 − δ = 1 − ε2 2 and max{(cid:107)y − xn(cid:107),(cid:107)y∗ − x∗ n(cid:107)} (cid:62) ε − 1 n for every pair (y, y∗) ∈ Π(X). Since the space is finite-dimensional, we can find a subsequence of (xn, x∗ n) that converges to a pair (x, x∗) ∈ SX × SX∗ . This pair satisfies that x∗(x) (cid:62) 1 − δ and for every (y, y∗) ∈ Π(X), max{(cid:107)y − x(cid:107),(cid:107)y∗ − x∗(cid:107)} (cid:62) max{(cid:107)y − xn(cid:107),(cid:107)y∗ − x∗ n(cid:107)} − max{(cid:107)x − xn(cid:107),(cid:107)x∗ − x∗ n(cid:107)} (cid:62) ε − 1 n − max{(cid:107)x − xn(cid:107),(cid:107)x∗ − x∗ n(cid:107)} −→ ε. Since by Theorem 2.1, x∗(x) cannot be strictly smaller than 1 − δ, we have x∗(x) = 1 − δ. Therefore, we may apply Lemma 5.7 to find y∗ ∈ SX∗ and α (cid:62) 1 − ε 2 for which conditions (10) and (11) are fulfilled. Now we claim that in fact there is only one number γ ∈ R for which (15) and this γ equals 1 − ε (cid:107)x∗ − γy∗(cid:107) (cid:54) ε 2 2 . So α = 1 − ε 2 and, we also claim that (16) Indeed, when we were proving equation (14), we proved that every γ ∈ R that satisfies (15) must satisfy γ (cid:62) 1 − ε 2 . On the other hand, the function γ (cid:55)−→ (cid:107)x∗ − γy∗(cid:107) is convex, so the set G of those γ ∈ R satisfying (15) also must be convex; but 1 /∈ G, so γ < 1. Finally, according to (10), and (cid:107)x∗ − αy∗(cid:107) = ε 2 (cid:107)x∗ − y∗(cid:107) = ε. ε 2 (cid:62) 1 − γ = (cid:107)y∗ − γy∗(cid:107) (cid:62) (cid:107)x∗ − y∗(cid:107) − (cid:107)x∗ − γy∗(cid:107) (cid:62) ε 2 . This means that all the inequalities above must be equalities, so γ (cid:54) 1 − ε The claim is proved. 2 , and also (16) must be true. BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 21 Now, let us define u∗ := x∗ − αy∗ (cid:107)x∗ − αy∗(cid:107) = 2 ε (x∗ − (1 − ε 2 )y∗), and let us show that functionals u∗ and y∗ span a subspace of X∗ isometric to (cid:96)(2)∞ . According to Lemma 5.2, it is sufficient to show that (cid:107)u∗ − y∗(cid:107) = (cid:107)u∗ + y∗(cid:107) = 2. Let us do this. At first, (cid:13)(cid:13)(cid:13)(cid:13) 2 ε )y∗) − y∗(cid:13)(cid:13)(cid:13)(cid:13) = (x∗ − (1 − ε 2 (cid:107)x∗ − y∗(cid:107) = 2. 2 ε (cid:107)u∗ − y∗(cid:107) = At second, 2 (cid:62) (cid:107)u∗ + y∗(cid:107) = (cid:107) 2 ε (x∗ − (1 − ε 2 )y∗) + y∗(cid:107) = 2 ε (cid:107)x∗ − y∗ + εy∗(cid:107) (cid:62) 2 ε (x∗ − y∗ + εy∗)(v) = 2. (cid:3) √ Let us comment that for complex Banach spaces, we cannot expect that Theorem 5.8 provides a complex copy of (cid:96)(2)∞ in the dual of the space. Namely, the two-dimensional complex space X = (cid:96)(2) 1 2δ for δ ∈ (0, 1/2) but it does not contain the complex space (cid:96)(2)∞ (of course, it satisfies ΦX (δ) = contains the real space (cid:96)(2)∞ as a subspace since (cid:96)(2) and (cid:96)(2)∞ are isometric in the real case). We do 1 2δ for some not know whether it is true a result saying that if a complex space X satisfies ΦX (δ) = δ ∈ (0, 1/2), then X contains a copy of the complex space (cid:96)(2) or a copy of the complex space (cid:96)(2)∞ . √ 1 Let us extend the result of Theorem 5.8 to the infinite-dimensional case. Roughly speaking, we proceed as in the proof of such theorem, but instead of selecting convergent subsequences, we select subsequences such that their numerical characteristics (like norms of elements, pairwise distances, or values of some important functionals) have limits. Theorem 5.9. Let X be an infinite-dimensional Banach space. Suppose that there is δ ∈ (0, 1/2) such 2δ). Then X∗ (and hence also X) contains almost that ΦX (δ) = isometric copies of (cid:96)(2)∞ . √ 2δ (or, equivalently, ΦS X (δ) = √ √ 2δ. There is a sequence of pairs (xn, x∗ n) ∈ SX × SX∗ such that x∗ n(xn) > 1 − δ = Proof. Denote ε := 1 − ε2 2 and max{(cid:107)y − xn(cid:107),(cid:107)y∗ − x∗ (17) for every pair (y, y∗) ∈ Π(X). Since we have x∗ limn x∗ that there is a sequence (yn) of elements in SX such that n(yn − xn) x∗ n(xn) = 1 − δ. Denote t = 2 (cid:107)yn − xn(cid:107) (cid:54) t lim n(cid:107)} (cid:62) ε − 1 n n(xn) (cid:54) 1 − (ε − 1 n )2/2 by Theorem 2.1, we deduce that ε . Now, we are going to proceed like in Lemma 5.6 in order to show n lim n (18) Pick a sequence (tn) with tn > t, n ∈ N and limn tn = t. Using Lemma 5.3, for every n ∈ N we get yn ∈ SX such that (19) For given n ∈ N, let u∗ u∗ n(yn) = 1 and u∗ yn − xn ∈ K(tn, x∗ n) n ∈ SX∗ be a functional that separates K(tn, x∗ n)) ⊂ [0,∞). Then, according to Lemma 5.5, we have n) + yn) ∩ BX = {yn}. n) + yn from BX , that is, satisfying n(K(tn, x∗ (K(tn, x∗ lim n and and x∗ n(yn) = 1. min{(cid:107)x∗ n − u∗ n(cid:107),(cid:107)x∗ n + u∗ n(cid:107)} (cid:54) 2/tn < ε. As we have n + u∗ (cid:107)x∗ n − u∗ n + u∗ n(cid:107) (cid:62) (x∗ n(yn) = 1 + x∗ n(cid:107) < ε, so (17) says that (cid:107)xn − yn(cid:107) (cid:62) ε − 1 n)(yn) = 1 + x∗ we get (cid:107)x∗ subsequence if necessary, we can assume that the following limits exist: limn (cid:107)xn − yn(cid:107), limn x∗ n . Without loss of generality, passing to a n(yn − xn) n(xn) + x∗ n(yn − xn) (cid:62) 2 − ε2 2 > ε, 22 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO and limn x∗ n(yn). Then ε (cid:54) lim n (cid:107)yn − xn(cid:107) (cid:54) lim n(yn) − 1 + x∗ n tnx∗ ε2 2 n(yn − xn) = t lim ) (cid:54) 2 ε (1 − 1 + ε2 2 n ) = ε. (cid:54) 2 ε (lim n n(yn − xn) x∗ This means that all the inequalities in the above chain are in fact equalities. In particular, limn x∗ and n(yn) = 1, (20) ε = lim n (cid:107)yn − xn(cid:107) = t lim n n(yn − xn), x∗ so the analogue of Lemma 5.6 is proved. Now, we proceed with analogue of Lemma 5.7: we need to show that there are y∗ n ∈ SX∗ and αn (cid:62) 0, αn −→ 1 − ε 2 with (cid:107)x∗ n − αny∗ (21) and there is a sequence of vn ∈ SX such that x∗ (22) n(vn) = lim n Case 1 : Assume that there exist r > 0 and n ∈ N such that, for all m > n, y∗ n(vn) = 1. n(cid:107) (cid:54) ε 2 and (cid:107)x∗ n(cid:107) (cid:62) ε, n − y∗ lim n (cid:0)(cid:2)K(t − r, x∗ m) + xm (cid:3) ∩ BX (cid:1) \ (xm + rBX ) (cid:54)= ∅. This means that for all m > n there is zm such that (cid:107)zm − xm(cid:107) > r, (cid:107)zm(cid:107) (cid:54) 1 and (cid:107)zm − xm(cid:107) (cid:54) (t − r)x∗ m(zm − xm). For λ ∈ (0, 1) denote ym,λ := λzm + (1 − λ)ym. Clearly, ym,λ ∈ BX . Denote also λm = inf{λ : ym,λ ∈ xm + K(t, x∗ m)}, and let us show that (23) lim m λm = 0. Observe first that λm is smaller than every value of λ for which On the one hand, if (cid:107)ym − xm(cid:107) − tx∗ job is done. On the other hand, if (cid:107)ym − xm(cid:107) − tx∗ (cid:107)ym,λ − xm(cid:107) (cid:54) tx∗(ym,λ − xm). m(ym − xm) (cid:54) 0, then λ = 0 belongs to the set in question, and the m(ym − xm) (cid:62) 0, then there is λ for which λ(cid:107)zm − xm(cid:107) + (1 − λ)(cid:107)ym − xm(cid:107) = tλx∗ m(zm − xm) + t(1 − λ)x∗ m(ym − xm) is positive and belongs to the set in question. This means that λm (cid:54) (cid:107)ym − xm(cid:107) − tx∗ (cid:107)ym − xm(cid:107) − tx∗ m(ym − xm) + tx∗ m(ym − xm) m(zm − xm) − (cid:107)zm − xm(cid:107) , but the limit of the right-hand side equals 0 thanks to (20). So condition (23) is proved. This means that ym,λm ∈ xm + K(t, x∗ m) and (cid:107)ym,λm − ym(cid:107) (cid:54) 2λm −→ 0. Let us pick a little bit bigger λm > λm in such a way that we still have (cid:107)ym,λm − ym(cid:107) −→ 0, but for some tn < t with tn −→ t, we have (cid:107)ym,λm ) = limn x∗ − xm(cid:107) (cid:54) tnx∗ n(yn) = 1. Let us apply Lemma 5.3. There are vn ∈ SX such − xm). m(ym,λm (24) Then, in particular, limn x∗ that n(yn,λn (cid:2)K(tn, x∗ n) + vn (cid:3) ∩ BX = {vn}. (25) vn − yn,λn ∈ K(tn, x∗ n) and BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 23 n(vn − yn,λn Then x∗ part of (22). Condition (24) imply that yn,λn vn − xn ∈ K(tn, x∗ ) (cid:62) 0, i.e. 1 (cid:62) x∗ n). Consequently, n(vn) (cid:62) x∗ n(yn,λn − xn ∈ K(tn, x∗ ) −→ 1, so x∗ n(vn) −→ 1. This proves the first n) which, together with (25), mean that (cid:107)vn − xn(cid:107) (cid:54) tnx∗ n(vn − xn) (cid:54) tn ε2 2 < ε. If we denote by y∗ (this proves the second part of (22) even in a stronger form) so, thanks to (17), n ∈ SX∗ the functional that separates vn + K(tn, x∗) from BX , then (vn, y∗ n) ∈ Π(X) Also, by Lemma 5.5, dist (x∗ n, Lin y∗ n) (cid:54) 1/tn, so there are αn ∈ R such that Again, without loss of generality, we may assume that the sequences (αn) and (cid:107)x∗ Then, n − αny∗ n(cid:107) converge. (cid:107)y∗ n − x∗ n(cid:107) (cid:62) ε − 1 n . (cid:107)x∗ − αny∗ n(cid:107) (cid:54) 1/tn. (cid:107)x∗ n − αny∗ n(cid:107) (cid:54) 1 t = ε 2 . lim n Consequently, (cid:62) lim (cid:107)x∗ n − αny∗ n(cid:107) (cid:62) lim n − αny∗ (x∗ n)(v) = 1 − lim αn , ε 2 n n n so limn αn (cid:62) 1 − ε 2 . Starting at this point, (21) can be deduced in the same way as it was done for (16). Case 2 : Assume that there is a sequence of rn > 0, rn −→ 0 and that there is a subsequence of (xm, x∗ m) (that we will again denote (xm, x∗ m)) such that (cid:0)(cid:2)K(t − rm, x∗ (cid:2)K(t − rm, x∗ m) + xm m) + xm Then also Let us separate 1 − rm from BX by a norm-one functional y∗ n, that is, y∗ (26) so, in particular, y∗ is a sequence (xn, y∗ m(xm) (cid:62) 1− rm and limm y∗ n) ∈ Π(X), such that (cid:0)K(t − rm, x∗ n m) + xm (for all m ∈ N). (for all m ∈ N). (cid:3) ∩ BX (cid:1) \ (xm + rmBX ) = ∅ (cid:3) ∩ (1 − rm)BX = ∅ (cid:2)K(t − rm, x∗ m) + xm 1 (cid:3) (cid:1) > 1 − rm m(xm) = 1. By the Bishop-Phelps-Bollob´as theorem, there max{(cid:107)xn − xn(cid:107),(cid:107)y∗ n − y∗ n(cid:107)} −→ 0 as n → ∞. Again, passing to a subsequence, we can assume that all the numerical characteristics that appear here have the corresponding limits. According to (17), for n big enough, we have n(cid:107)} (cid:62) ε − 1 n n(cid:107) = max{(cid:107)xn − xn(cid:107),(cid:107)y∗ n(cid:107) (cid:62) ε. We can select zn ∈ xn + K(t − rn, x∗ n − x∗ n) in such a way that (cid:107)zn − yn(cid:107) −→ 0. Then so limn(cid:107)y∗ n − x∗ n − x∗ (cid:107)y∗ , 1 (cid:62) lim n y∗ n(yn) = lim n y∗ n(zn) (cid:62) lim (1 − rn) = 1. n This means that condition (22) works for vn := yn. Now consider an arbitrary w ∈ ker y∗ n( w) = −rn. Then, by (26), w /∈ int(cid:0)K(t − rn, x∗ n ∩ SX . Taking a convex combination with an element h of the n(h) almost equals −1, we can construct an element w ∈ BX such that (cid:107) w− w(cid:107) (cid:54) 2rn n)(cid:1), so (cid:107) w(cid:107) (cid:62) (t − rn)x∗ unit sphere where y∗ and y∗ n( w). Consequently, x∗ n(w) (cid:54) x∗ n( w) + 2rn (cid:54) 1 t − rn + 2rn. 24 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO Observe that we have shown that the values of the functional x∗ Therefore, by Lemma 5.4, n on ker y∗ n∩ SX do not exceed 1 t−rn + 2rn. and so there are αn ∈ R such that dist (x∗ n, Lin y∗ + 2rn −→ 1 t n) (cid:54) 1 t − rn n − αny∗ (cid:107)x∗ n(cid:107) (cid:54) 1 t . lim n The remaining conditions in (21) and (22) can be deduced the same way as in the case 1. Finally, (21) and (22) imply that limn (cid:107)x∗ n(cid:107) = limn (cid:107)x∗ from the corresponding part of the Theorem 5.8 demonstration. n − y∗ n − y∗ n(cid:107) = 2: the proof does not differ much (cid:3) Corollary 5.10. Let X be a uniformly non-square Banach space. Then, ΦS 2δ for every δ ∈ (0, 1/2). Consequently, every superreflexive Banach space can be equivalently renormed in such a way that, in the new norm, ΦS 2δ for all δ ∈ (0, 1/2). X (δ) (cid:54) ΦX (δ) < √ X (δ) (cid:54) ΦX (δ) < √ It would be interesting to obtain a quantitative version of the above corollary. 6. A three dimensional space E containing (cid:96)(2)∞ with ΦE(δ) < √ 2δ In the last section we decided to decorate our paper with two diamonds: Figure 4. The unit ball of D 1 2 Figure 5. The unit ball of D∗ 1 2 √ The first of them represents the unit ball of the space Dε that we construct below, and on the second picture one can see the unit ball of D∗ ε . Like in the previous section, for every δ ∈ (0, 1/2) we denote ε = 2δ, so 0 < ε < 1. We denote ε ⊂ R3 the absolute convex hull of the following 11 points Ak, k = 1, . . . , 11 (or, what is the same, the B3 convex hull of 22 points ±Ak, k = 1, . . . , 11): A1 = (0, 0, ), 3 4 A2 = (1 − ε, 1, A6 = (1, 1 − ε, ε 2 ε 2 ), A3 = (1 − ε,−1, ), A7 = (−1, 1 − ε, ), A4 = (ε − 1, 1, ε 2 ), A8 = (1, ε − 1, ε 2 A10 = (1, 1, 0), A11 = (1,−1, 0). ε 2 ε 2 ), A5 = (ε − 1,−1, ), A9 = (−1, ε − 1, ε 2 ε 2 ), ), Denote Dε ("D" from "Diamond") the normed space(cid:0)R3,(cid:107) · (cid:107)(cid:1), for which B3 ε is its unit ball. Then D∗ ε on x ∈ Dε is just can be viewed as R3 with the polar of B3 the standard inner product in R3. Let us list, without proof, some properties of Dε whose verification is straightforward: ε as the unit ball, and the action of x∗ ∈ D∗ ε BISHOP-PHELPS-BOLLOB ´AS MODULI OF A BANACH SPACE 25 • The subspace of Dε formed by vectors of the form (x1, x2, 0) is canonically isometric to (cid:96)(2)∞ . • There are no other isometric copies of (cid:96)(2)∞ in Dε. • The subspace of D∗ ε formed by vectors of the form (x1, x2, 0) is canonically isometric to (cid:96)(2) 1 (and so, is isometric to (cid:96)(2)∞ ). • There are no other isometric copies of (cid:96)(2)∞ in D∗ ε . • The following operators act as isometries both on Dε and D∗ ε : (x1, x2, x3) (cid:55)−→ (x2, x1, x3), (x1, x2, x3) (cid:55)−→ (x1,−x2, x3). In other words, changing the sign of one coordinate or rearranging the first two coordinates do not change the norm of an element. √ The following theorem shows that the existence of an (cid:96)(2)∞ -subspace does not imply that ΦX (δ) = 2δ, even in dimension 3. Theorem 6.1. Let δ ∈ (0, 1/2), ε = √ 2δ, and X = Dε. Then ΦX (δ) < √ √ 2δ. Proof. Assume contrary that ΦX (δ) = of a pair (x, x∗) ∈ SX × SX∗ with the following properties: x∗(x) = 1 − δ and (27) Also, repeating the proof of Theorem 5.8 for this x∗ ∈ SX∗ , we can find u∗, y∗ ∈ SX∗ such that the pair (u∗, y∗) is 1-equivalent to the canonical basis of (cid:96)(2) max{(cid:107)z − x(cid:107),(cid:107)z∗ − x∗(cid:107)} (cid:62) ε for every pair (z, z∗) ∈ Π(X). 2δ. Like in the proof of Theorem 5.8, this implies the existence and 1 u∗ = 2 ε (x∗ − (1 − ε 2 )y∗). 2 u∗ + (1 − ε This means that x∗ = ε 2 )y∗. What can be this (u∗, y∗) if we take into account that there is in X∗? It can be either u∗ = (1, 0, 0), y∗ = (0, 1, 0), or a pair of vectors only one isometric copy of (cid:96)(2) 1 that can be obtained from this one by application of isometries, i.e. just 8 possibilities. Consequently, x∗ either equals to the vector (ε/2, 1 − ε/2, 0), or to a vector that can be obtained from this one by application of isometries, again just 8 possibilities. By duality argument, there are u, y ∈ SX such that the pair (u, y) is 1-equivalent to the canonical basis of (cid:96)(2) 1 and x = ε 2 u + (1 − ε 2 )y. Since the only (up to isometries) pair u, y ∈ SX of this kind is u = (1, 1, 0), y = (1,−1, 0), we get x = (1, 1 − ε, 0), or can be obtained from this one by application of isometries. So there are 8 × 8 = 64 possibilities for the pair (x, x∗). Taking into account that x∗(x) = 1 − δ we reduce this number to 8 possibilities: x = (1 − ε, 1, 0), x∗ = (ε/2, 1 − ε/2, 0) and images of this pair under remaining 7 reflections and rotations of the underlying R2. If we show that this choice of (x, x∗) do not satisfy condition (27) then, by symmetry, the remaining choices would not satisfy (27) neither, and this would give us the desired contradiction. Indeed, the pair (z, z∗) ∈ Π(X) that do not satisfy (27) for x = (1 − ε, 1, 0), x∗ = (ε/2, 1 − ε/2, 0) is the following one: z = (1 − ε, 1, ε/2), z∗ = (ε/2, 1 − ε/2, ε). Let us check the required properties. At first, z = A2 ∈ SX . Then, z∗(z) = 1. The last property means, that (cid:107)z∗(cid:107) (cid:62) 1, so in order to check that (cid:107)z∗(cid:107) = 1 it remains to show that z∗(Ak) (cid:54) 1 for all k. This is true for ε < 1. Finally, (cid:107)z − x(cid:107) = (cid:107)(0, 0, ε/2)(cid:107) = ε (cid:3) 3 ε < ε, and (cid:107)z∗ − x∗(cid:107) = (cid:107)(0, 0, ε)(cid:107) = (cid:104)(0, 0, ε), A1(cid:105) = 3 3 A1(cid:107) = 2 4 ε < ε. 2(cid:107) 4 References [1] Albiac F., Kalton N., Topics in Banach Space Theory, Graduate Texts in Mathematics 233, Springer, New York, 2006. [2] E. Behrends, M -structure and the Banach-Stone Theorem, Lecture Notes in Math. 736, Springer-Verlag, Berlin, 1979. 26 CHICA, KADETS, MART´IN, MORENO-PULIDO, AND RAMBLA-BARRENO [3] E. Behrends, M -complements of M -ideals, Rev. Roumaine. Math. Pures Appl. 29 (1984), 537 -- 541. [4] E. Bishop and R. R. Phelps, A proof that every Banach space is subreflexive, Bull. Amer. Math. Soc 67 (1961), 97 -- 98. [5] B. Bollob´as, An extension to the theorem of Bishop and Phelps, Bull. London Math. Soc. 2 (1970), 181 -- 182. [6] F. F. Bonsall and J. Duncan, Numerical Ranges II, London Math. Soc. Lecture Note Series 10, Cambridge 1973. [7] B. Cascales, V. Kadets, and A. J. Guirao, A Bishop-Phelps-Bollob´as type theorem for uniform algebras, preprint. [8] J. Diestel, Geometry of Banach spaces Lecture notes in Math. 485, Springer-Verlag, Berlin, 1975. [9] D. Van Dulst, Reflexive and superreflexive Banach spaces. Mathematical Centre Tracts. 102. Amsterdam: Mathe- matisch Centrum. V, 1978, 273 pp. [10] P. Harmand, D. Werner, and D. Werner, M -ideals in Banach spaces and Banach algebras, Lecture Notes in Math. 1547, 1993. [11] R.C. James, Uniformly non-square Banach spaces. Ann. Math. (2) 80, (1964) 542-550 [12] Kadets, V.M. On two-dimensional universal Banach spaces. (Russian) C. R. Acad. Bulg. Sci. 35(1982), 1331-1332. [13] R. R. Phelps, Support Cones in Banach Spaces and Their Applications, Adv. Math. 13 (1974), 1 -- 19. [14] R. R. Phelps, Convex functions, monotone operators and differentiability (second edition), Lecture Notes in Math. 1364, Springer-Verlag, Berlin, 1993. (Chica & Mart´ın) Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain E-mail address: [email protected] [email protected] (Kadets) Department of Mechanics and Mathematics, Kharkiv V.N. Karazin National University, 61022 Kharkiv, Ukraine E-mail address: [email protected] (Moreno-Pulido & Rambla-Barreno) Departamento de Matem´aticas, Universidad de C´adiz, Puerto Real (C´adiz), Spain E-mail address: [email protected] [email protected]
1002.3902
1
1002
2010-02-20T15:22:24
Approximation of operators in Banach spaces
[ "math.FA" ]
It is a translation of an old paper of mine. We describe the topology tau_p in the space Pi_p(Y,X), for which the closures of convex sets in tau_p and in *-weak topology of the space Pi_p(Y,X) are coincident. Thereafter, we investigate some properties of the space Pi_p, related to this new topology. 2010-remark: Occasionally, the topology is coincides with the lambda_p-topology from the paper "Compact operators which factor through subspaces of l_p", Math. Nachr. 281(2008), 412-423 by Deba Prasad Sinha and Anil Kumar Karn.
math.FA
math
APPROXIMATION OF OPERATORS IN BANACH SPACES O.I. Reinov 0 1 0 2 b e F 0 2 ] . A F h t a m [ 1 v 2 0 9 3 . 2 0 0 1 : v i X r a §0. Notations This work is the first part of some investigations which are concerned with an approx- imation of ones or others classes of operators in Banach spaces (the approximation can be understood in different sences). The main ojects of the study will be the spaces, pos- sessing (or not possessing) the so-called approximation perperties APp and the bounded approximation properties BAPp of order p, where p > 0 (see, e.g., [9], [10]). In the pro- pounded part, we consider only the properties APp where p > 1, and our considerations are bounded by some equivalent reformulations of the first-given definition from [9]. The main accent is put on the investigation of the conditions, under which the finite di- mensional operators are dense in norm in the space of quasi-p-nuclear operators, as well as in the space of all absolutely p-summing operators in the topology of "πp-compact convergence". We will keep the standard notations and the terminology. If A is a bounded subset of a Banach space X, then Γ(A) is the closed absolutely convex hull of A; XA is the Banach space with "the unit ball" Γ(A); ΦA : XA → X is the canonical embedding. For B ⊂ X, it is denoted by B the closures of the set B in the topology τ and in the norm k · k respectively. When it is necessary, we denote by k · kX the norm in X. Other notations: Πp, QNp, Np, Ip are the ideals of absolutely p-summing, quasi- and B k·k τ p-nuclear, p-nuclear , strictly p-integral operators, respectively (see [8], [10]); X ∗b⊗pY is the complete tensor product associated with Np(X, Y ); X ∗bb⊗pY = X ∗ ⊗ Y πp (i.e. the closure of the set of finite dimensional operators in Πp(X, Y )). Finally, if p ∈ [1, +∞] then p′ is the adjoint exponent. Everywhere bellow, we use to suppose that p ∈ [1, +∞]; however, the proofs are given for p ∈ (1, +∞) (with some non-essential changes, all proofs pass through the cases p = 1 and p = +∞). The domain of changes for p is noted specially only in the places where it is necessary, or where a vagueness can be arised. [R -- KGU] Reinov O.I., Approximation of operators in Banach spaces, in book Primenenie funkcional'nogo analiza v teorii priblizheni, Kalinin: KGU, 1985, 128-142. 1 §1. Approximation of absolutely p-summing operators In this paragraph we will describe the topology τp in the space Πp(Y, X), for which the closures of convex sets in τp and in ∗-weak topology of the space Πp(Y, X) are coincident (this is an answer to a corresponding question of P. Saphar in [12], p.385). Thereafter, we will investigate some properties of the space Πp, related to this new topology. Thus, consider in the space Πp(Y, X) the topology τp of πp-compact convergence, a local base (in zero) of which is defined by sets of type ωK,ε = {U ∈ Πp(Y, X) : πp(U ΦK) < ε} , where ε > 0, K = Γ(K) -- a compact subset of Y. 1.1. Proposition.. Let R be a linear subspace in Πp(Y, X), containing Y ∗ ⊗ X. Then (∗) 1 x′ functional on (R, τp). ϕ(U ) = trace U ◦ z, U ∈ R . (R, τp)′ is isomorphic to a factor space of the space X ∗b⊗p′ Y. More precisely, if ϕ ∈ (R, τp)′, then there exists an element z =P∞ On the other hand, for every z ∈ X ∗b⊗p′ Y the relation (∗) defines a linear continuous n ⊗ yn ∈ X ∗b⊗p′ Y such that The Proof of this proposition will be given in detail. However, we will omit details in analogues cases hereinafter. So, let ϕ be a linear continuous functional on (R, τp). Then one can find a neighborhood of zero ωK = ωK,ε, such that ϕ is bounded on it: ΦK−→ ∀ U ∈ ωK, ϕ(U ) 6 1. We may assume that ε = 1. Consider the operator U ΦK : YK U−→ X. Since the mapping ΦK is compact, U ΦK ∈ QNp(YK, X). Put ϕk(U ΦK ) = Y if V = U ΦK ∈ RK and the space QNp(YK, X), the linear functional ϕK is bounded: πp(V ) 6 1, then ϕK(V ) = ϕ(U ) 6 1. Therefore, ϕK can be extended to a linear ϕ(U ) for U ∈ R . On the linear subspace RK = (cid:8)V ∈ QNp(YK , X) : V = U ΦK(cid:9) of continuous functional eϕ on the whole QNp(YK, X); moreover, because of the injectivity that eϕ ∈ QNp(YK, C(K))∗. Let us mention that of the ideal QNp, considering X as a subspace of some space C(K), we may assume (1) eϕ(jU ΦK) = ϕK (U ΦK) = ϕ(U ) (here j is an isometric embedding of X into C(K)). Furthermore, since QNp(YK, C(K))∗ = Ip′ (C(K), (YK)∗∗), we can find an operator Ψ : C(K) → (YK)∗∗, for which Let An ∈ (YK)∗ ⊗ C(K), πp(An − jU ΦK) → 0. Then eϕ(A) = trace ΨA, A ∈ (YK )∗ ⊗ C(K). (2) Consider the operator Φ∗∗ compact, we have Φ∗∗ eϕ(jU ΦK ) = lim trace ΨAn. −→ (YK)∗∗ K Ψ : C(K) Ψ K Ψ ∈ Np′ (C(K), Y ) = C(K)∗b⊗p′ Y. LetP∞ 2 1 µn ⊗ yn ∈ C(K)∗b⊗p′ Y Φ∗∗ K−→ Y. Since Ψ ∈ Ip′ , and ΦK is be a representation of the operator Φ∗∗ generates an operator Φ∗∗ K Ψ. Put z = P j∗(µn) ⊗ yn. The element z K Ψj from X to Y. We will show now that trace U ◦z = eϕ(jU ΦK) (note that U ◦ z is an element of the space X ∗b⊗X, so the trace is well defined). We have: trace U ◦ z = trace(cid:16)X j∗(µn) ⊗ U yn(cid:17) =Xhj∗(µn), U yni = (3) K Ψ = trace (jU ΦK)∗∗Ψ, =Xhµn, jU yni = trace jU Φ∗∗ Ψ −→ (YK )∗∗ Φ∗∗ K−→ Y U−→ X j −→ C(K). Since πp(An − 1 wm ⊗ fm ∈ where (jU ΦK )∗∗Ψ : C(K) jU ΦK) → 0, then πp (A∗∗ (YK)∗ ⊗ C(K), then n − (jU ΦK)∗∗) → 0. Moreover, if A := An =PN n Ψ =Xm hΨ∗wm, fmi =Xm hwm, Ψfmi = trace ΨA. trace A∗∗ Hence, trace (jU ΦK )∗∗Ψ = lim trace A∗∗ n Ψ = lim trace ΨAn. Now, it follows from n x′ nkq c−1 ΦK (z) = z. (cid:4) n < +∞. Consider neighborhood ωK,ε of zero in τp. For this, we need can find a compact K ⊂ Y and an operator A2 : lq → YK , for which A1 = ΦKA2. n ⊗ yn, {cn} ∈ c0 andP kx′ (3) and (2) that eϕ(jU ΦK) = trace U ◦ z. Finally, we get from (1): ϕ(U ) = trace U ◦ z. Thus, the functional ϕ is defined by an element of X ∗b⊗p′ Y. Inversely, if z ∈ X ∗b⊗p′ Y, put ϕ(U ) = trace U ◦ z for U ∈ R (the trace is defined since U ◦ z ∈ X ∗b⊗X). We have to show that the linear functional ϕ is bounded on a 1.2. Lemma. If z ∈ X ∗b⊗qY, then z ∈ X ∗b⊗qYK , where K = Γ(K) is a compact in Y. Proof of the lemma. Let z =P x′ the operator A1 : lq → Y, A1{an} = P ancnyn. Since this operator is compact, one Put z0 =P c−1 n ⊗ en (en are orths in lq). Then z := (1 ⊗ A2)(z0) ∈ X ∗b⊗qYK , and z ∈ X ∗b⊗p′ YK . If U ∈ ωK,1, then πp(U ΦK ) < 1 and trace U ΦK ◦ z 6 kzkX ∗ b⊗p′ YK 1.3. Corollary. (R, τp)′ = (R, σ)′, where σ = σ(R, X ∗b⊗p′ Y ). Thus, the closures of Denote by X ∗e⊗p′ Y the closure of the set X ∗ ⊗ Y in the space Ip′ (X, Y ∗∗) (dual to Y ∗bb⊗pX). 1.4. Proposition. Let A be the intersection of the unit ball of the (dual to X ∗e⊗p′ Y ) space G = G(Y, X ∗∗) with the subspace Y ∗ ⊗ X. ∗-weak closure of the set A in G ∩ Πp(Y, X) coincides with the closure of A in (Πp(Y, X), τp). Let us continue the proof of the theorem. Let K be a compact subset of Y, for which · convex subsets of the space Πp(Y, X) in τp and in σ are the same. (cid:4) πp(U ΦK) 6 C. (cid:4) Proof. Let us consider the canonical mappings X ∗b⊗p′ Y Πp(Y, X ∗∗) −−−−→ X ∗e⊗p′ Y, ←−−−− G(Y, X ∗∗). j ∗ j 3 Since j∗ is one-to-one, then the closures of the bounded sets in G(Y, X ∗∗), in topologies σ(G, X ∗e⊗p′ Y ) and σ(G, X ∗b⊗p′ Y ) are the same. Therefore, if B is a convex bounded set in G(Y, X) ⊂ Πp(Y, X), then the closure of the set B in(cid:0)Πp(Y, X) ∩ G, σ(G, X ∗e⊗p′ Y )(cid:1) coincides with the closure of the set B in the space (cid:0)Πp(Y, X), σ(Πp(Y, X), X ∗b⊗p′ Y )(cid:1) and, therefore, by Corollary 1,3, -- with the closure of B in (Πp(Y, X), τp) . (cid:4) 1.5. Corollary. With notations of the proposition 1.4, the closure of the set A in τp coincides with the closure of A in the space L(Y, X) in the topology of compact convergence. For the proof, it is enough to use the previous assertion, considering the canonical mapping from X ∗b⊗1Y into X ∗e⊗p′ Y instead of the map j from the proof of the propo- sition 1.4 (and to apply either Proposition 1.1 for p = +∞, or results on duality from [4]). (cid:4) 1.6. Corollary. Let C > 0 and T ∈ Πp(Y, X). The following assertions are equivalent: 1) there ia a net {Tα} , Tα ∈ Y ∗ ⊗ X, converging to T in the topology τp such that πp(Tα) 6 C; 2) there is a net {Tα} , Tα ∈ Y ∗ ⊗ X, converging to T in the topology of compact convergence, such that πp(Tα) 6 C. (cid:4) 1.7. Proposition. For an operator T ∈ Πp(Y, X), T (Y ) = X, the following are the same: 1) T ∈ Y ∗ ⊗ X 2) there is a net of operators Rα ∈ Y ∗ ⊗ Y such that T Rα → T in the topology τp. τp ; Proof. Assuming that 2) is not valid, we (by Corollary 1.3) get: (4) 6 ∃ Rα ∈ Y ∗ ⊗ Y : T Rα → T in (cid:0)Πp(Y, X), σ(Πp(Y, X), X ∗b⊗p′ Y )(cid:1) . Consider the associated with T mappings: X ∗b⊗p′ Y Πp(Y, X ∗∗) eT−−−−→ Y ∗b⊗1Y, ←−−−− L(Y, Y ∗∗). eT ∗ ∗ (the closure is taken in the space Πp(Y, X) in ∗-weak topology). It follows from (4) that T is not zero on where eT (z) = z ◦ T for z ∈ X ∗b⊗p′ Y. Let Z = eT ∗(Y ∗ ⊗ Y ) the subspace Z ⊥ ⊂ X ∗b⊗p′ Y, i.e. there exists an element A ∈ Z ⊥ such that hT, Ai = trace AT = 1. But AT = eT (A), and if R ∈ Y ∗ ⊗ Y, then heT (A), Ri = hA, (eT ) ∗(R)i = 0. Hence, the element eT (A) of the projective tensor product Y ∗b⊗Y is not zero (since trace eT (A) = 1), but generates a null-operator in Y. For any y′ ∈ Y ∗ and T y ∈ T (Y ) tensor element A ∈ X ∗b⊗p′ Y generates a zero-operator. Again, by using the equality trace AT = 1, we conclude that T can not be approximated in ∗-weak topology by finite rank operators. Now, it follows from Corollary 1.3 that 1) is not fulfilled. (cid:4) we have: hA, T y ⊗ y′i = hAT y, y′i = 0. Since T (Y ) = X, we obtain that a non-zero Next two statements give us sufficient (but not necessary, as we will see below) con- ditions for the density of the set of all finite rank operators in the space of operators Πp(Y, X) in the topology τp of πp-compact convergence. 4 1.8. Proposition1. If QNp(Y, X) = Y ∗ ⊗ X each Y Πp(Y, X) = Y ∗ ⊗ X τp . πp for every Banach space Y, then for successfully instead of V an operator eV ΦK, where eV ∈ Y ∗ ⊗ X. Let V =PN Proof. Let U ∈ Πp(Y, X), ε > 0, K = Γ(K) be a compact in Y. By the assumptions, there is an operator V ∈ (YK)∗ ⊗ X, such that πp(V − U ΦK ) < ε. We need to set n=1 zn ⊗ xn. Note that we can consider only the case where Φ∗∗ K is one-to-one (else, with the help of the construction of [2] we change YK by a space YK0 , for which the operator ΦK0 is compact In this case Y ∗ is norm dense in (YK)∗ and, and the operator Φ∗∗ therefore, for every positive number sequence {εn} there exist the elements y′ n ∈ Y ∗, n ⊗ xn ∈ Y ∗ ⊗ X. Let {an} be a sequence of for which ky′ K0 is one-to-one). 1 y′ K k the elements of the space YK, such that sup(cid:8)P han, a′ip : ka′kY ∗ mXi=1 n − zn, aii xnkp 6 n − znkY ∗ < εn. Put eV =PN mXi=1 k(eV ΦK − V )aikp = NXn=1 mXi=1 kxnkp′!p/p′ 6 NXn=1 NXn=1 n − zn, aiip NXn=1 NXn=1 kxnkp′!p/p′ 6 NXn=1 Y ∗ K n − znkp hy′ ky′ hy′ 6 6 K 6 1(cid:9) 6 1. We have: kxnkp′!p/p′ εp n. NXn=1 If we take εn small enough then the last number is less then ε, and, from the inequality πp(V − U ΦK ) 6 ε, we get that πp(eV ΦK − U ΦK ) 6 ε + πp(V − eV ΦK) 6 2ε. Hence, eV − U ∈ ωK,2ε. Thus, we have shown that for every neighborhood ωK,ε there exists an operator eV ∈ Y ∗ ⊗ X, for which eV − U ∈ ωK,ε. Therefore U ∈ Y ∗ ⊗ X 1.9. Proposition. If the canonical mapping j : X ∗b⊗p′ Y → Np′ (X, Y ) is one-to-one dual to X ∗b⊗p′ Y, coincides with Πp(Y, X ∗∗). On the other hand, in any case j−1(0)⊥ = then Πp(Y, X) = Y ∗ ⊗ X Proof. If the map j is one-to-one then the annihilator j−1(0)⊥ of its kernel in the space, (the closure in ∗-weak topology of the space Πp(Y, X ∗∗)); by Corollary 1.3, τp. (cid:4) Y ∗ ⊗ X τp . ∗ ∗ Πp(Y, X) ∩ Y ∗ ⊗ X = Y ∗ ⊗ X τp . Therefore, Πp(Y, X) = Y ∗ ⊗ X τp. (cid:4) it follows from 1.3 and 1.9 For a reflexive space X, the dual space to X ∗b⊗p′ Y is equal to Πp(Y, X). Consequently, 1.10. Corollary. For a reflexive space X the canonical mapping j : X ∗b⊗p′Y → Np′ (X, Y ) is one-to-one iff the set of finite rank operators is dense in the space Πp(Y, X) in the topology τp of πp-compact convergence. (cid:4) To conclude this part of our considerations, let us give an assertion which shows the following. It follows from the existing of an absolutely p-summing operator, which is non-approximated in the topology τp, that there esits a non-approximated (in the same topology) quasi-p-nuclear operator (with values in the same space). 1In the original paper [R -- KGU] this Proposition was sounded as follows: "If QNp(Y, X) = Y ∗ ⊗ X πp , then Πp(Y, X) = Y ∗ ⊗ X τp ." 5 τp, then there 1.11. Proposition. If there exists an operator T ∈ Πp(Y, X) \ Y ∗ ⊗ X exist a reflexive space Z and an operator U ∈ QNp(Z, X), which is not in the closure Y ∗ ⊗ X τp. For the proof it is enough to remember the definition of the topology τp and to use the following two facts: a) if V is a compact operator then it can be represented as a composition of two compact operators; b) the product AB of a compact operator B and absolutely p-summing operator A is a quasi-p-nuclear map (see [5], [8]). (cid:4) 1.12. Remark. 2 §2. Approximation properties of Banach spaces In this paragraph, we fix the space of images of operators (or the spaces where oper- ators are defined), and investigate the conditions under which it is possible to approx- imate (by finite rank operators) all absolutely p-summing mappings with values in a given space (or acting from a given space).3 2.1. Proposition. For a Banach space X the following are equivalent: 1) for every Banach space Y the equality QNp(Y, X) = Y ∗ ⊗ X 2) for every Banach space Y the equality Πp(Y, X) = Y ∗ ⊗ X 3) for every Banach space Y one has QNp(Y, X) ⊂ Y ∗ ⊗ X 4) for every reflexive Banach space Y the equality QNp(Y, X) = Y ∗ ⊗ X τp holds; πp holds; τp;4 τp holds. Proof. Implications 2) =⇒ 3) =⇒ 4) are evident; 1) =⇒ 2) by Proposition 1.8. For the proof of the implication 4) =⇒ 1), consider an operator V ∈ QNp(Y, X). By using the results of [2], we can factorize the operator V by the following way: V = U A, where A ∈ Lc(Y, Z), U ∈ QNp(Z, X), and, moreover, the space Z is reflexive. Put K = A(BallY ). From 4), it follows that ∀ ε > 0 ∃eV ∈ Z ∗ ⊗ X : eV − U ∈ ωK,ε, i.e. πp(eV ΦK − U ΦK ) < ε. Hence, πp(V −eV A) 6 CA πp(U ΦK −eV ΦK ) 6 CAε. (cid:4) 2.2. Corollary. If for every reflexive space Y the canonical mapping X ∗b⊗p′ Y → Np′ (X, Y ) is one-to-one then for each Banach space Y the equality QNp(Y, X) = Y ∗ ⊗ X holds. (cid:4) πp For the proof, Proposition 1.9 may be applied, and then use the implication 2) =⇒ 1) of the previous fact. (cid:4) It follows from Corollaries 1.10 and 2.2 2In the original paper [R -- KGU] this non-essential remark is sounded exactly as "At the moment, I do not know whether the inverse of Proposition 1.8 is true in the case where the space X is reflexive. 3In the original paper [R -- KGU], there was here the phrase "The next statement, among other things, gives us a partial inversion of Proposition 1.8." 4In the original paper [R -- KGU], there was here an evident misprint: "for every Banach space Y the equality QNp(Y, X) = Y ∗ ⊗ X τp holds;" 6 2.3. Corollary. For a reflexive Banach space X the following are equivalent: 1) for each space Y the canonical mapping X ∗b⊗p′ Y → Np′ (X, Y ) is one-to-one; 2) for each space Y the set of finite rank operators is dense in the Banach space QNp(Y, X). (cid:4) The next statement, among other things, gives us a partial inversion of Proposition 1.9. 2.4. Corollary. For every Banach space X the following are equivalent: 1) for each space Y the canonical mapping Y ∗b⊗pX → Np(Y, X) is one-to-one; 2) for each reflexive Banach space Y the canonical mapping Y ∗b⊗pX → Np(Y, X) is 3) for each (reflexive) Banach space Y X ∗ ⊗p′ Y τp′ = Πp′ (X, Y ). one-to-one; Proof. Concerning the equivalence 1) ⇐⇒ 2), see [1]; the implications 1) =⇒ 3) and 3) =⇒ 2) follow from 1.9 and 1.10, respectively. (cid:4) The previous statement yields the following well known result: 2.5. Corollary. If a Banach space X has the approximation property then for any p > 1 and any space Y the canonical mapping Y ∗b⊗pX → Np(Y, X) is one-to-one. (cid:4) The above results lead us to the following definition which is equivalent to corre- sponding definition in [9], [10], [11]. 2.6. Definition. Let p > 1. A Banach space X has the property APp (the approxima- tion property of order p), if every absolutely p′-summing operator, acting from the space X, can be approximated in the topology of π′ p-compact convergence τp′ by operators of finite rank. It follows from Proposition 1.8 and Corollary 2.4 that 2.7. Corollary. If for each Banach space Y the equality QNp′ (X, Y ) = X ∗ ⊗ Y holds then the space X has the property APp . (cid:4) πp′ Recall for the sake of completeness the following assertion on a characterization of the spaces with the property APp (a proof can be found in [1]). 2.8. Proposition. A Banach space X has the property APp iff for every Banach space Y, for every operator T ∈ Πp′ (X, Y ), for each weakly p′-summable sequence {xk} of elements of the space X and for every ε > 0 there is a finite rank operator R : X → Y, such thatP kU xk − Rxkkp′ < ε. As the property AP1 (the usual approximation property og Grothendieck), the prop- erties APp are very useful in investigations of the questions of different kinds in the geometrical theory of operatos (for instance, when describing the dual spaces of some spaces of operators; so, if X possesses the property APp, then Np(Y, X)∗ = Πp′(X, Y ∗∗) for every space Y ). However, we would like (concluding this paragrap) to adduce a sim- ple fact, which is valid without any assumptions on approximation (see [3] for p = 1 and [12] for p > 1, where an analogues assertion was proved with a supposition of approxi- mation property; see also [7], Theorem 4.6, where it is obtained a little bit less general result). 7 2.9. Proposition. Let p ∈ [1, +∞). The following are equivalent 1) Banach spaces X and Y are reflexive; 2) the space QNp(X, Y ) is reflexive; 3) the space Πp(X, Y ) is reflexive. Proof. Since, for reflexive spaces X and Y, the equality QNp(X, Y ) = Πp(X, Y ) holds, it is sufficient to prove only the implication 1) =⇒ 2). For this, imbedd the space Y into some space C(K) isometrically, and let us consider the space QNp(X, Y ) as a subspace of QNp(X, C(K)). Then any continuous functional Φ on QNp(X, Y ) has an for T ∈ QNp(X, Y ) (where j is the imbedding of Y into C(K)). Since the space Y is reflexive, one has U j ∈ Np′ (Y, X) [7]. Therefore, as a functional, Φ is generated by an extension with the same norm to a linear functional eΦ on QNp(X, C(K)). In turn, the functional eΦ is generated by an operator U ∈ Ip′ (C(K), X ∗∗), so hΦ, T i = trace U jT element of the space Y ∗b⊗p′ X. Thus, the natural mapping Y ∗b⊗p′ X → QNp(X, Y )∗ is an "onto" map. Since (cid:0)Y ∗b⊗p′ X(cid:1)∗ = QNp(X, Y ), we get that the space QNp(X, Y ) is reflexive. (cid:4) §3. Counterexamples A lot of counterexamples concerning AP's can be found in [9], [10], [11]. We will use them partially. Recall that for any p > 1, p 6= 2, there exists a separable reflexive Banach space without the property APp . Moreover, for every p > 1, p 6= 2, there exist a separable reflexive E, an operator R ∈ Πp′ (E, E) and a tensor element t ∈ E∗b⊗pE, so that trace t ◦ R = 1 and trace t ◦ A = 0 for each finite rank operator A ∈ E∗ ⊗ E (for details, we refer the reader to the papers [9], [10], [11]). It is often very useful to apply the following fact when constructing some counterex- amples (see [6]): 3.1. Lemma. For every separable Banach space E there exist a separable conjugate Banach space H = Y ∗ with a basis and operators Q : H → E and U : H ∗ → E∗ so that Q(H) = E, U (H ∗) = E∗, kU k 6 1, kQk 6 1, U Q∗ = idE∗ and the space (idE∗ −Q∗U ) (H ∗) is isomorphic to the space Y. Now we are ready for constructions of our counterexamples. Firstly, we will show that the inversion of Proposition 1.9 and Corollary 2.2 are not valid. 3.2. Proposition. For every p ∈ [1, +∞], p 6= 2, there exist a separable reflexive space E, a separable conjugate space H with a basis such that the canonical mapping QNp(E, H) = E∗ ⊗ H πp and Πp(E, H) = E∗ ⊗ H τp. element and operator so that trace t ◦ R = 1 and t = 0 as an operator. Set g = t ◦ Q ∈ j : H ∗b⊗p′ E → Np′ (H, E) is not one-to-one. On the other hand, since H has the AP, Proof. Let E, t ∈ E∗b⊗p′ E and R ∈ Πp(E, E) be the mentioned above spaces, tensor H ∗b⊗p′ E and V = U ∗R ∈ QNp(E, H ∗∗). Then trace V ◦ g = 1 (consequently, the tensor element z ∈ Z ∗b⊗p′ Z and an operator Ψ ∈ Πp(Z, Z ∗∗) such that trace Ψ ◦ z = 1, but 3.3. Corollary. For every p ∈ [1, +∞], p 6= 2, there exist a Banach space Z, an element g is not equal to zero) and g = 0 as an operator. (cid:4) trace Φ ◦ z = 0 for all Φ ∈ Πp(Z, Z). 8 every operator A ∈ Πp(E, H) (because of the space H has approximation property). n, 0) ⊗ (0, yn) and define the operator Ψ ∈ QNp(Z, Z ∗∗) by Proof. Let us use notation introduced in the proof of Proposition 3.2. Let P y′ be any representation of a tensor element g ∈ H ∗b⊗p′ E. Note that trace A ◦ g = 0 for Put Z = H ⊕ E, z =P(y′ trace Ψ ◦ z =Xh(y′ n, 0), (V yn, 0)i =Xhy′ Ψ(h, y) = (V y, 0). We have: n ⊗ yn n, V yni = trace V ◦ g = 1. On the other hand, denoting by PH and PE the natural projectors from Z onto H and E respectively, we have, for arbitrary operator Φ ∈ Πp(Z, Z), Φ(h, y) = ΦH(h) + ΦE(y) = (PHΦH (h) + P EΦH(h)) + (PH ΦE(y) + P EΦE(y)) , whence, trace Φ ◦ z =Xh(y′ n, 0), [PHΦ(0, yn)]i =Xh(y′ n, 0), PHΦE(yn)i. Denoting by A the operator PH ΦE, we get: A ∈ Πp(E, H); trace Φ ◦ z = trace A ◦ g = 0. (cid:4) 3.4. Remark. For p = +∞ we get nonzero tensor element z ∈ Z ∗b⊗1Z, vanishing on the subspace L(Z, Z) of the space L(Z, Z ∗∗). This is an answer to a question of Swedish mathematician Sten Kaijser, who was one who is directed my attention for a possibility of the existemce of such an element z. Now we will show that the inversion of Proposition 1.85 and Corollary 2.7 are invalid too. 3.5. Proposition. For every p ∈ [1, +∞], p 6= 2, there exist a separable reflexive space E, a separable conjugate space H with a basis (so, with the property APp′ ) such that τp (see Corollary 2.4). QNp(H, E) 6= H ∗ ⊗ E πp; on the other hand, Πp(H, E) = H ∗ ⊗ E Proof. With the notation of the proof of Proposition 3.2, set L = RQ. Since trace (t ◦ RQ∗∗Y ∗) = 1 and t = 0 as an operator, the map RQ∗∗ can not be approximated by finite rank operators in the space QNp(H ∗∗, E). Moreover, L 6⊂ H ∗ ⊗ E πp. (cid:4) In conclusion, let us bring the following, at first sight somewhat surprising, statement which shows, roughly speaking, that there are spaces X and Y, for which the closures in (Πp(X, Y ∗∗), w∗) of the set of all finite dimensional operators is minimal: coincides with X ∗bb⊗pY. 5In [R -- KGU] I meant that, for fixed X and Y, in Proposition 1.8 Πp(Y, X) = Y ∗ ⊗ X τp ; QNp(Y, X) = Y ∗ ⊗ X πp 9 3.6. Proposition. For every p ∈ [1, +∞) there exist (separable and reflexive) Banach spaces X and Y such that Proof. Let X and Y be th separable and reflexive spaces such that the natural mapping X ∗ ⊗ Y τp = X ∗bb⊗pY. can be identified with a subspace X ∗ ⊗ Y τp of the spaces Πp(X, Y ) (see also the proof Y ∗b⊗p′ X → Np′ (Y, X) is not one-to-one. By Proposition 1.1, the dual space to Np′ (Y, X) of Proposition 1.9). Since this subspace is reflexive (Proposition 2.9) and (X ∗bb⊗pY )∗ = Np′ (Y, X), we have: X ∗bb⊗pY = X ∗ ⊗ Y 3.7. Remark. Seemingly, it is unknown whether Proposition 3.6 is true in the case where p = +∞. τp . (cid:4) References 1. Bourgain J., Reinov O.I., On the approximation properties for the space H∞, Math. Nachr. 122 (1985), 19-27. 2. Davis W.J, Figiel T., Johnson W.B., Pelczynski A., Factoring weakly compact operators, J. Func- tional Analysis 17 (1974), 311-327, M R50#8010.. 3. Y.Gordon, D.R.Lewis, H.R.Retherford, Banach ideals of operators with applications, J. Funct. Anal. 14 (1973), no. 1, 85-129. 4. Grothendieck A., Produits tensoriels topologiques et espases nucl´eaires, Mem. Amer. Math. Soc. 16 (1955), pp. 196 + 140. 5. Johnson W.B., Factoring compact operators, Israel J. Math. 9 (1971), 337-345. 6. Lindenstrauss J., On James' paper "Separable Conjugate Spaces", Israel J. Math. 9 (1971), 279-284. 7. Makarov B.M., Samarskij V.G., Weak sequencial completeness and close properties of he spaces of operators, Theory of operators and theory of functions, 1, LGU, Leningrad, 1983, 122-144 8. Pietsch A., Operator ideals, North-Holland, Deutscher Verlag der Wiss., Berlin, 1978. 451 p.. 9. Reinov O.I., Approximation properties of order p and the existence of non-p-nuclear operators with p-nuclear second adjoints, Doklady AN SSSR 256 (1981), no. 1, 43-47. 10. Reinov O.I., Approximation properties of order p and the existence of non-p-nuclear operators with p-nuclear second adjoints, Math. Nachr. 109 (1982), 125-134. 11. Reinov O.I., Disappering tensor elements in the scale of p-nuclear operators, Theory of operators and theory of functions (LGU) (1983), 145-165. 12. Saphar P., Produits tensoriels d'espaces de Banach et classes d'applications lineaires, Studia Math. 38 (1970), 71-100 M R43#878.. 198904, Saint Petersburg, St. Petersburg State University, Dept. Math., E-mail address: [email protected] 10
0711.3919
2
0711
2010-03-05T20:27:00
Small Subspaces of L_p
[ "math.FA" ]
We prove that if $X$ is a subspace of $L_p$ $(2<p<\infty)$, then either $X$ embeds isomorphically into $\ell_p \oplus \ell_2$ or $X$ contains a subspace $Y,$ which is isomorphic to $\ell_p(\ell_2)$. We also give an intrinsic characterization of when $X$ embeds into $\ell_p \oplus \ell_2$ in terms of weakly null trees in $X$ or, equivalently, in terms of the "infinite asymptotic game" played in $X$. This solves problems concerning small subspaces of $L_p$ originating in the 1970's. The techniques used were developed over several decades, the most recent being that of weakly null trees developed in the 2000's.
math.FA
math
SMALL SUBSPACES OF Lp R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Abstract. We prove that if X is a subspace of Lp (2 < p < ∞), then either X embeds isomorphically into ℓp ⊕ ℓ2 or X contains a subspace Y, which is isomorphic to ℓp(ℓ2). We also give an intrinsic characterization of when X embeds into ℓp ⊕ℓ2 in terms of weakly null trees in X or, equivalently, in terms of the "infinite asymptotic game" played in X. This solves problems concerning small subspaces of Lp originating in the 1970's. The techniques used were developed over several decades, the most recent being that of weakly null trees developed in the 2000's. 0 1 0 2 r a M 5 ] . A F h t a m [ 2 v 9 1 9 3 . 1 1 7 0 : v i X r a 1. Introduction The study of "small subspaces" of Lp (2 < p < ∞) was initiated by Kadets and Pe lczy´nski [KP] who proved that if X is an infinite dimensional subspace of Lp, then either X is isomorphic to ℓ2 and the L2-norm is equivalent to the Lp-norm on X, or for all ε > 0 X contains a subspace Y which is 1 + ε-isomorphic to ℓp. In [JO1] it was shown that if X does not contain an isomorph of ℓ2 then X embeds isomorphically into ℓp ([KW] showed that, moreover, for all ε > 0, X 1 + ε-embeds into ℓp). W.B. Johnson [J] solved the analogous problem for X ⊆ Lp (for all 1 < p < 2) by proving that X embeds into ℓp if for some K < ∞ every weakly null sequence in SX, the unit sphere of X, admits a subsequence K-equivalent to the unit vector basis of ℓp. Using the machinery of [OS1] (see also [OS2]) and the special nature of Lp, these results were unified in [AO] as: X ⊆ Lp (1 < p < ∞) embeds into ℓp if (and only if) every weakly null tree in SX admits a branch equivalent to the unit vector basis of ℓp. After ℓp and ℓ2 the next smallest natural subspace of Lp (2 < p < ∞) is ℓp ⊕ ℓ2. Indeed if X ⊆ Lp does not embed into either ℓp or ℓ2, it contains an isomorph of ℓp ⊕ ℓ2. The next small natural subspace after ℓp ⊕ ℓ2 is ℓp (ℓ2) or, as it is sometimes denoted, (P ℓ2)p. In [JO2] it was shown that if X ⊆ Lp (2 < p < ∞) and X is a quotient of a subspace of ℓp ⊕ ℓ2 then X embeds into ℓp ⊕ ℓ2. The motivating problem for this paper (and our main result) dates back to the 1970's. We prove that if X ⊆ Lp (2 < p < ∞) and X does not embed into ℓp ⊕ ℓ2 then X contains an isomorph of ℓp (ℓ2). To solve this we first give an intrinsic characterization of when X embeds into ℓp ⊕ ℓ2. The terminology is explained in Section 3. We assume that our space K∼ B Lp is defined over an atomless and separable probability space (Ω, Σ, P). We write A if A ≤ KB and B ≤ KA. X will always denote an infinite dimensional Banach space. Theorem A. Let X be a subspace of Lp (2 < p < ∞). Then the following are equivalent. a) X embeds into ℓp ⊕ ℓ2. 2000 Mathematics Subject Classification. 46B20, 46B25, Research of the last two authors was partially supported by the National Science Foundation. 1 2 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT b) Every weakly null tree in SX admits a branch (xi) satisfying for some K and all scalars (ai), (1.1) (cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13) (k · k2 denotes the L2-norm) [0, 1], and all scalars (ai) K∼ (cid:16)Xaip(cid:17)1/p ∨(cid:13)(cid:13)(cid:13)X aixi (cid:13)(cid:13)(cid:13)2 . c) Every weakly null tree in SX admits a branch (xi) satisfying, for some K, (wi) ⊆ (1.2) then K∼ (cid:16)Xaip(cid:17)1/p ∨(cid:16)Xai2w2 i(cid:17)1/2 . (cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13) p(cid:17)1/p kxk K∼ (cid:16)Xkxnkp ∨ kxk2 . 2(cid:17)1/2 p(cid:17)1/p =(cid:16)Xkxnkp Under any of these conditions the embedding of X into ℓp ⊕ ℓ2 is given by: producing a blocking (Hn) of the Haar basis for Lp and 1 ≤ K < ∞, so that, if X ∋ x =P xn, xn ∈ Hn, ∨(cid:16)Xkxnk2 Since (P Hn)p is isomorphic to ℓp this suffices. The next task is to show that if X violates these conditions then X contains a comple- mented subspace isomorphic to ℓp (ℓ2). We will present two proofs of this. The first proof will roughly show that X must contain "skinny" uniform copies of ℓ2 and hence contain uniform ℓ2's, (Xn)n∈N for which if xn ∈ SXn then the xn's are almost disjointly supported and hence behave like the unit vector basis of ℓp. Then an argument due to Schechtman will prove that a subspace of X which is isomorphic to ℓp(ℓ2) contains an isomorphic copy of ℓp(ℓ2) which is complemented in Lp. The second proof will lead to more precise result using the random measure machinery of D. Aldous [Ald] and the stability theory of Lp [KM]. For easier reading we will, however, recall all relevant definitions and results concerning random measures and stability theory. We will show that the complemented copy of ℓp(ℓ2) is witnessed by stabilized ℓ2 sequences living on almost disjoint supports, meaning that the joint support of the elements of the Xn's is almost disjoint, not only the support of the elements of a given sequence (xn) with xn ∈ Xn, for n ∈ N. This yields the following: If X is a subspace of Lp, and X is not contained in ℓ2⊕ ℓp, then X must contain a complemented copy of ℓp(ℓ2). Moreover, it admits a projection onto a subspace isomorphic to ℓp(ℓ2), whose norm is arbitrarily close to that of the minimal norm projection of Lp onto any subspace isomorphic to ℓ2. Theorem B. Let X ⊆ Lp (2 < p < ∞). If X does not embed into ℓp⊕ ℓ2 then for all ε > 0, X contains a subspace Y , which is 1 + ε-isomorphic to ℓp (ℓ2), and Y is complemented in Lp by a projection of norm not exceeding (1 + ε)γp where γp = kxkp, x being a symmetric L2 normalized Gaussian random variable. Moreover, we can write Y as the complemented sum of Yn's where Yn is (1+ε)-isomorphic to ℓ2 and Y is (1 + ε)-isomorphic to the ℓp-sum of the Yn's, and there exists a sequence (An) of disjoint measurable sets so that kyAnkp ≥ (1 − ε2−n)kyk for all y ∈ Yn and n ∈ N. The original proof of the [JO2] result about quotients of subspaces of ℓp ⊕ ℓ2, is quite complicated, and a byproduct of our results will be to give a much easier proof (see Sec- tion 7). In addition, we can characterize when X ⊆ Lp (2 < p < ∞) embeds into ℓp ⊕ ℓ2 in terms of its asymptotic structure [MMT]. From the [KP] and [JO1] results we first note that X ⊆ Lp (2 < p < ∞) embeds into ℓp if and only if it is asymptotic ℓp, and X embeds into ℓ2 if and only if it is asymptotic ℓ2. SMALL SUBSPACES OF Lp 3 structure of X, there exists (wi)n Let us say X is asymptotic ℓp⊕ ℓ2 if for some K and all (ei)n 1 ⊆ [0, 1] so that for all (ai)n K∼ (cid:16) n X1 1 ∈ {X}n, the nth asymptotic 1 ⊆ R, ai2wi2(cid:17)1/2 aip(cid:17)1/p ∨(cid:16) n X1 X1 (1.3) n . We note that the space ℓp⊕ℓ2 is itself asymptotic ℓp⊕ℓ2. Indeed, denote by (fi) and (gi) the unit vector bases of ℓp and ℓ2, respectively, viewed as elements of ℓp⊕ ℓ2. For (x, y) ∈ ℓp⊕ ℓ2 we put k(x, y)k = kxkp∨kyk2. Since (fi) and (gi) are 1-subsymmetric and ℓp⊕ ℓ2 is reflexive, the elements of the nth asymptotic structure of ℓp ⊕ ℓ2 are exactly the sequences (zi)n i=1 in ℓp ⊕ ℓ2, for which there are 0 = k0 < k1 < k2 < . . . kn in N, and (aj), (bj) in R with aiei(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) zi = ki Xj=ki−1+1 (ajfj + bjgj), ki vi =(cid:16) so that kzik = vi ∨ wi = 1, where Xj=ki−1 ajp(cid:17)1/p i=1 ⊂ [−1, 1] we therefore compute ∨(cid:16) n Xi=1 Xi=1 For (ξi)n n ξizi(cid:13)(cid:13)(cid:13) =(cid:16) n Xi=1 ξipvp (cid:13)(cid:13)(cid:13) Assuming now that (otherwise (1.3) follows immediately) , and wi =(cid:16) ki Xj=ki−1 bj2(cid:17)1/2 . (cid:13)(cid:13)(cid:13) It follows therefore that (zi) satisfies (1.3) with K = 2 and we deduce that ℓp ⊕ ℓ2 is asymptotic ℓp ⊕ ℓ2. For n ∈ N let (e(n) i,j : i, j ≤ n) be the unit vector basis of ℓn Xi,j=1 ai,j2(cid:17)p/2(cid:17)1/p , for all (ai,j) ⊂ R. ai,je(n) 2 ), i.e. p (ℓn n Xi=1(cid:16) n i,j(cid:13)(cid:13)(cid:13) =(cid:16) n Xj=1 Note that (e(n) i,j ) is, ordered lexicographically, isometrically in the (n2)th asymptotic struc- ture of ℓp(ℓ2), for all n ∈ N, but it is not hard to deduce from the aforementioned description of the asymptotic structure of ℓp ⊕ ℓ2, that (e(n) i,j ) is not (uniformly in n ∈ N) in the (n2)th asymptotic structure of ℓp ⊕ ℓ2. Theorem B yields therefore the following Corollary C. X ⊆ Lp (2 < p < ∞) embeds into ℓp ⊕ ℓ2 if and only if X is asymptotic ℓp ⊕ ℓ2. i(cid:17)1/p (cid:16) n Xi=1 i +(cid:16) n Xi=1 ξipvp ξipvp we deduce that n Xi=1 (cid:13)(cid:13)(cid:13) ξizi(cid:13)(cid:13)(cid:13) p ≥ 1 2h n Xi=1 ≤(cid:16) n Xi=1 ξip(cid:17)1/p ∨(cid:16) n Xi=1 ξi2w2 i(cid:17)1/2 . ξi2w2 i(cid:17)1/2 ≥(cid:16) n Xi=1 i(cid:17)1/p ξi2w2 i(cid:17)1/2 , ξi2w2 i(cid:17)p/2i ≥ 1 2 n Xi=1 ξip(vp i ∨ wp i ) = 1 2 n Xi=1 ξip. 4 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Indeed, if X does not embed into ℓp ⊕ ℓ2, thenby Theorem B it contains an isomorph of ℓp (ℓ2), which is not asymptotic ℓp ⊕ ℓ2. Using Theorem A and Theorem B we will be able to deduce the following additional surprising characterization of subspaces of Lp which embed into ℓp ⊕ ℓ2. It is analogous to the characterization of subspaces of Lp which embed in ℓp via normalized weakly null sequences (see the aforementioned result from [J]) and we thank W. B. Johnson for having pointed it out to us. Corollary D. X ⊆ Lp (2 < p < ∞) embeds into ℓp ⊕ ℓ2 if and only if there exists a K ≥ 1 so that every normalized weakly null sequence in SX admits a subsequence (xi) satisfying for all scalars (ai), (1.4) (cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13) K∼ (cid:16)Xaip(cid:17)1/p ∨(cid:16)X a2 i kxik2 . 2(cid:17)1/p A proof of Corollary D will be given at the end of Section 5. It is worth noting that (1.4) is a reformulation of (1.1) in (b) of Theorem A. The difference here is that the constant K is uniform and not dependent on the particular sequence. Without the uniformity assump- tion, the Corollary would be false (see Theorem 2.4 below). In Section 2 we recall some inequalities for unconditional basic sequences and martingales in Lp. Section 3 contains the proof of Theorem A, along with the necessary preliminaries on weakly null trees, and the "infinite asymptotic game." In Section 4 we introduce a dichotomy of Kadets -- Pe lczynski type and apply the results of Section 2 to embed a class of subspaces of Lp into ℓp ⊕ ℓ2. Section 5 considers the subspaces of Lp which do not embed in ℓp ⊕ ℓ2; we show that such subspaces contain "thinly supported ℓ2's". More precisely, for some K < ∞, we find sub- spaces Yn, n ∈ N, which are K-isomorphic to ℓ2, but for which the natural equivalence of k · kp and k · k2 on Yn is bad. By this we mean that kykp ≥ Mnkyk2, for all y ∈ Yn, for some sequence (Mn) ⊂ R, with Mn ր ∞, as n ր ∞. This will enable us to argue that we can choose the Yn's so that vectors yn ∈ SYn, n ∈ N, are almost disjointly supported and hence the closed linear span of the Yn's is isomorphic to ℓp(ℓ2). Section 6 refines the result of Section 5, obtaining alomst disjointly supported ℓ2's, by applying techniques from Aldous's paper [Ald] on random measures. As well as the new proof of the result from [JO2] mentioned above, Section 7 includes a construction of subspaces of Lp, isomorphic i=1 ℓ2(cid:1)p. In to ell2, which embed only with bad constants in spaces of the form ℓp ⊕(cid:0)Lm Section 8 we recall what is known and not known about small Lp-spaces and raise a problem about when X ⊂ Lp embeds into ℓp(ℓ2). In light of the deep work of [BRS] in constructing uncountably many separable Lp spaces, it is likely that further study of their ordinal index will be needed to make progress on classifying the next group of smaller Lp-spaces. We are especially grateful to the referee for two incredibly detailed reports which greatly improved our exposition. 2. Some inequalities in Lp We first recall the well known fact that an unconditional basic sequence in Lp is trapped between ℓp and ℓ2. Proposition 2.1. (see e.g. [AO]) Let (xi) be a normalized λ-unconditional basic sequence in Lp (2 < p < ∞). Then for all (ai) ⊆ R λ−1(cid:16)Xaip(cid:17)1/p ≤(cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13)p ≤ λ Bp(cid:16)Xai2(cid:17)1/2 . SMALL SUBSPACES OF Lp 5 is the Rademacher sequence. H. Rosenthal proved that if the xi's are independent and mean zero random variables in In Proposition 2.1, Bp is the Khintchin constant, kP airik ≤ Bp(P ai2)1/2, where (ri) Lp then they span a subspace of ℓp ⊕ ℓ2. Theorem 2.2. [R] Let 2 < p < ∞. There exists Kp < ∞ so that if (xi) is a normalized mean zero sequence of independent random variables in Lp, then for all (ai) ⊆ R (cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13)p Kp∼ (cid:16)Xaip(cid:17)1/p ∨(cid:16)Xai2kxik2 2(cid:17)1/2 . D. Burkholder extended this result to martingale difference sequences as follows. Theorem 2.3. ([B], [BDG], [H]) Let 2 < p < ∞. There exists Cp < ∞ so that if (zi) is a martingale difference sequence in Lp, with respect to the sequence (Fn) of σ-algebras, then (cid:13)(cid:13)(cid:13)X zi(cid:13)(cid:13)(cid:13)p Cp∼ (cid:16)Xkzikp p(cid:17)1/p ∨(cid:13)(cid:13)(cid:13)(cid:16)X E[z2 i Fi−1](cid:17)1/2(cid:13)(cid:13)(cid:13)p , where E(xF) denotes the conditional expectation of an integrable random variable x with respect to a sub-σ-algebra F. From [KP], it follows that every normalized weakly null sequence in Lp admits a subse- quence (xi), which is either equivalent to the unit vector basis of ℓp or equivalent to the unit vector basis of ℓ2. The latter occurs if ε = limi kxik2 > 0 and the lower ℓ2 estimate is (essentially) Using Theorem 2.3, W.B. Johnson, B. Maurey, G. Schechtman, and L. Tzafriri obtained a quantitative improvement. ε(cid:16)Xai2(cid:17)1/2 ≤(cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13)p . Theorem 2.4. [JMST, Theorem 1.14] Let 2 < p < ∞. There exists Dp < ∞ with the following property. Every normalized weakly null sequence in Lp admits a subsequence (xi) satisfying for some w ∈ [0, 1], for all (ai) ⊆ R, Dp∼ (cid:16)Xaip(cid:17)1/p ∨ w(cid:16)Xai2(cid:17)1/2 Thus in particular [(xi)], the closed linear subspace generated by (xi) uniformly embeds (cid:13)(cid:13)(cid:13)X aixi(cid:13)(cid:13)(cid:13)p . into ℓp ⊕ ℓ2. 3. A criterion for embeddability in ℓp ⊕ ℓ2 In this section we prove Theorem A, and thus provide an intrinsic characterization of subspaces of Lp which isomorphically embed into ℓp ⊕ ℓ2. This characterization is based on methods developed in [OS1] and [OS2]. We will need the following notation. Let Z be a Banach space with a finite dimensional decomposition (FDD) E = (En). For n (z) = zn, n ∈ N, we denote the n-th coordinate projection by P E for z =P zi ∈ Z, with zi ∈ Ei, for all i ∈ N. For a finite A ⊂ N we put P E A =Pn∈A P E n : Z → En with P E n , i.e. P E n . 6 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT c00 denotes the vector space of sequences in R which are eventually 0 with unit vector basis (ei). More generally, if (Ei) is a sequence of finite dimensional Banach spaces, we define the vector space ∞ c00(⊕∞ i=1Ei) =n(zi) : zi ∈ Ei, for i ∈ N, and {i ∈ N : zi 6= 0} is finite(cid:9). The linear space c00(⊕∞ i=1Ei) is dense in each Banach space for which (En) is an FDD. If A ⊂ N is finite we denote by ⊕i∈AEi the linear subspace of c00(⊕Ei) generated by the elements of (Ei)i∈A. A blocking of (Ei) is a sequence (Fi) of finite dimensional spaces for which there is an increasing sequence (Ni) in N so that (N0 = 0) Fi = ⊕Ni j=Ni−1+1Ej, for any i ∈ N. sequence of finite dimensional spaces. Then we define for x = (xi) ∈ c00(⊕∞ Let V be a Banach space with a normalized 1-unconditional basis (vi) and E = (Ei) a i=1Ei) . Xi=1 kxik · vi(cid:13)(cid:13)(cid:13)V kxk(E,V ) =(cid:13)(cid:13)(cid:13) k · k(E,V ) is a norm on c00(⊕∞ i=1Ei), and we denote the completion of c00(⊕∞ i=1 Ei(cid:1)V . respect to k · k(E,V ), by (cid:0) ⊕∞ For z ∈ c00(⊕Ei) we define the E-support of z by suppE(z) = {i ∈ N : P E i (z) 6= 0}. A non-zero sequence (zj ) ⊂ c00(⊕Ei) is called a block sequence of (Ei) if max suppE(zn) < min suppE(zn+1), for all n ∈ N, and it is called a skipped block sequence of (Ei) if 1 < min suppE(z1) and max suppE(zn) < min suppE(zn+1) − 1, for all n ∈ N. Let δ = (δn) ⊂ (0, 1]. If Z is a space with an FDD (Ei), we call a sequence (zj) ⊂ SZ = {z ∈ Z : kzk = 1} a δ-skipped block sequence of (En), if there are 1 ≤ k1 < ℓ1 < k2 < ℓ2 < ··· in N so that kzn − P E (kn,ℓn](zn)k < δn, for all n∈ N. Of course one could generalize the notion of δ-skipped block sequences to more general sequences, but we prefer to introduce this notion only for normalized sequences. It is important to note that, in the definition of δ-skipped block sequences, k1≥ 1, and, thus, that the E1-coordinate of z1 is small (depending on δ1). Let i=1Ei), with T∞ = [ℓ∈N(cid:8)(n1, n2, . . . , nℓ) : n1 < n2 < ··· nℓ are in N(cid:9) . T∞ is naturally partially ordered by extension, i.e., (m1, m2, . . . mk) (cid:22) (n1, n2, . . . nℓ) if k ≤ ℓ and ni = mi, for i ≤ k. We call ℓ the length of α = (n1, n2, . . . nℓ) and denote it by α, with ∅ = 0 In this paper trees in a Banach space X are families in X indexed by T∞. For a tree (xα)α∈T∞ in X, and α = (n1, n2, . . . , nℓ)∈ T∞∪{∅}, we call the sequences of the form (x(α,n))n>nℓ nodes of (xα)α∈T∞ . The sequences (yn), with yi = x(n1,n2,...,ni), for i ∈ N, for some strictly increasing sequence (ni) ⊂ N, are called branches of (xα)α∈T∞ . Thus, branches of a tree (xα)α∈T∞ are sequences of the form (xαn) where (αn) is a maximal linearly ordered (with respect to extension) subset of T∞. If (xα)α∈T∞ is a tree in X and if T ′ ⊂ T∞ is closed under taking initial segments (if (n1, n2, . . . , nℓ) ∈ T ′ and m < ℓ then (n1, n2, . . . , nm) ∈ T ′) and has the property that for each α∈ T ′ ∪ {∅} infinitely many direct successors of α are also in T ′ then we call (xα)α∈T ′ a full subtree of (xα)α∈T∞ . Note that (xα)α∈T ′ could then be relabeled to a family indexed by T∞ and note that the branches of (xα)α∈T ′ are branches of (xα)α∈T∞ and that the nodes of (xα)α∈T ′ are subsequences of certain nodes of (xα)α∈T∞ . We call a tree (xα)α∈T∞ in X normalized if kxαk = 1, for all α ∈ T∞ and weakly null if every node is a weakly null sequence. If X has an F DD (Ei) we call (xα)α∈T∞ a block tree SMALL SUBSPACES OF Lp 7 with respect to (Ei) if every node and every branch (yn) is a block sequence with respect to (Ei). Note that, if (Ei) is an FDD for X and if (εα)α∈T∞ ⊂ (0, 1), every normalized weakly null tree (xα)α∈T∞ ⊂ X has a full subtree (zα)α∈T∞ which is an (εα)-perturbation of a block tree (yα) with respect to (Ei), i.e. kzα− yαk ≤ εα, for any α ∈ T∞. Let us also mention that the proof of the fact, that normalized weakly null sequences have basic subsequences whose basis constants are arbitrarily close to 1, generalizes to trees. This means that for a given ε > 0, and for any Banach space X, every normalized weakly null tree in X has a full subtree, all of whose nodes and all of whose branches are basic, and their basis constant does not exceed 1+ε. We now can state the main results of this section. Theorem 3.1. Let X be a subspace of Lp, 2 < p < ∞, and assume that there is a C > 1 so that every normalized weakly null tree in X admits a branch (yi) for which ∞ Xi=1 (cid:13)(cid:13)(cid:13) aiyi(cid:13)(cid:13)(cid:13)p ∞ Xi=1 C∼ max (cid:16) ∞ aip(cid:17)1/p Xi=1 n (x))n∈N, x(cid:1) ∈(cid:0) ⊕∞ T (x) =(cid:0)(P H aiyi(cid:13)(cid:13)(cid:13)2! for all (ai) ∈ c00. n=1 Hn(cid:1)ℓp ⊕ L2 ֒→ ℓp ⊕ L2, ,(cid:13)(cid:13)(cid:13) Then there is a blocking H = (Hn) of the Haar basis (hn) so that T : X → ℓp ⊕ L2, is an isomorphic embedding. Theorem 3.1 is a special case of the following result. By a 1-subsymmetric basis we mean one that is 1-unconditional and 1-spreading. Theorem 3.2. Let X and Y be separable Banach spaces, with X reflexive. Let V be a Banach space with a 1-subsymmetric and normalized basis (vi), and let T : X → Y be linear and bounded. Assume that for some C ≥ 1 every normalized weakly null tree of X admits a branch (xn) so that ∞ ∞ (3.1) Xi=1 anvn(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T(cid:16) ∞ Xi=1 Then there is a sequence of finite dimensional spaces (Gi), so that X is isomorphic to a C∼ (cid:13)(cid:13)(cid:13) anxn(cid:13)(cid:13)(cid:13)X (cid:13)(cid:13)(cid:13) Xi=1 i=1 Gi(cid:1)V ⊕ Y. subspace of (cid:0) ⊕∞ More precisely, under the above assumptions, if Z is any reflexive space with an FDD (Ei), and if S : X → Z is an isomorphic embedding, then there is a blocking (Gi) of (Ei) so that S is a bounded linear operator from X to (cid:0) ⊕∞ anxn(cid:17)(cid:13)(cid:13)(cid:13)Y for all (ai) ∈ c00. i=1 Gi(cid:1)V and the operator x 7→(cid:0)S(x), T (x)(cid:1), is an isomorphic embedding. (S, T ) : X →(cid:0) ⊕∞ i=1 Gi(cid:1)V ⊕ Y, Remark. Theorem 3.1 can be obtained from Theorem 3.2 by letting V = ℓp, Y = L2, Z = Lp, with the FDD (Ei) given by the Haar basis, S is the inclusion map from X into Lp and T is the formal identity map from Lp to L2 restricted to X. As noted in [OS2, Corollary 2, Section 2] (see also [OS1] for similar versions) the tree condition in Theorem 3.2 can be interpreted as follows in terms of the "infinite asymptotic game", (IAG) as it has been called by Rosendal [Ro]. 8 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Let C ≥ 1 and let A(C) be the set of all sequences (xn) in SX which are C-basic and satisfy condition (3.1). The (IAG) is played by two players: Player I chooses a subspace X1 of X having finite co-dimension, and Player II chooses x1 ∈ SX1, then, again Player I chooses a subspace X2 of X of finite codimension , and Player II chooses an x2 ∈ SX2. These moves are repeated infinitely many times, and Player I is declared the winner of the game if the resulting sequence (xn) is in A(C). A(C) is closed with respect to the infinite product of (SX , d), where d denotes the discrete topology on SX. This implies that this game is determined [Ma], i.e., either Player I or Player II has a winning strategy and as noticed in [OS2, Corollary 2, Section 2] for all ε > 0 Player I has a winning strategy for A(C+ε) if and only if for all ε > 0, every weakly null tree in SX has a branch, which lies in A(C+ε). Proof of Theorem A using Theorem 3.1. The interpretation of our tree condition in terms of the infinite asymptotic game, easily implies that the existence of a uniform C ≥ 1, so that all weakly null trees (xα) ⊂ SX admit a branch in A(C), is equivalent to the condition, that every weakly null tree (xα) ⊂ SX admits a branch in A(C), for some C ≥ 1. Indeed, if such a uniform C does not exist, Player II could choose a sequence (Cn) in R+ which increases to ∞ and could play the following strategy: first he follows his winning strategy for achieving a sequence (xn) outside of A(C1) and after finitely many steps, s1, he must have chosen a sequence x1, x2, . . . , xs2, which is either not C1-basic or does not satisfy (3.1) for some a = (ai)s1 i=1 ∈ Rs1. Then Player II follows his strategy for getting a sequence outside of A(C2), and continues that way using C3, C4 etc. It follows that the infinite sequence (xn), which is obtained by Player II cannot be in any A(C). Therefore Player II has a winning strategy for choosing a sequence outside ofSC≥1 A(C) which means that there is a weakly null tree, (zα), none of whose branches are in SC≥1 A(C) . Using Theorem 3.1, we deduce therefore (b)⇒(a) of Theorem A. The implication (a)⇒(c) in Theorem A is easy, using arguments like those above establishing that ℓp⊕ℓ2 is asymptotic ℓp ⊕ ℓ2. In order to show (c)⇒ (b) let (xα) be a normalized weakly null tree in Lp. After passing to a full subtree, and perturbing, we can assume that (xα) is a block tree with respect to the Haar basis. By (c) there is branch (zn), a sequence (wi) ⊂ [0, 1] and C ≥ 1 so that (cid:13)(cid:13)(cid:13)X aizi(cid:13)(cid:13)(cid:13)p C∼ (cid:16)Xaip(cid:17)1/p ∨(cid:16)X w2 i a2 i(cid:17)1/2 for all (ai) ∈ c00. Since (zi) is an unconditional sequence and since k · k2 ≤ k · kp on Lp it follows from Proposition 2.1 that for some constant cp (3.2) (3.3) (cid:13)(cid:13)(cid:13)X aizi(cid:13)(cid:13)(cid:13)p ≥ cp(cid:16)Xaip(cid:17)1/p . ∨(cid:13)(cid:13)(cid:13)X aizi(cid:13)(cid:13)(cid:13)2 We claim that our branch (zn) satisfies (1.1) for some K < ∞. Assuming this were not true, then we could use (3.2), and choose a normalized block sequence (yn) of (zn), say yn = kn Xi=kn−1+1 aizi, with ai ∈ R, for i ∈ N and 0 = k0 < k1 < . . ., SMALL SUBSPACES OF Lp 9 so that for all n ∈ N (3.4) (3.5) For any (bi) ∈ c00 it follows therefore from (3.2) that kn i = 1, and kn w2 i a2 Xi=kn−1+1 (cid:16) aip(cid:17)1/p Xi=kn−1+1 (cid:13)(cid:13)(cid:13)X bnyn(cid:13)(cid:13)(cid:13)p ∨ kynk2 < 2−n. C∼ (cid:16)Xbn2(cid:17)1/2 , thus (yn) is C-equivalent to the unit vector basis of ℓ2. The result by Kadets and Pe lczy´nski [KP] yields that k · kp and k · k2 must be equivalent on Y . But limn→∞ kynk2 = 0 by (3.5), so we have a contradiction. (cid:3) For the proof of Theorem 3.2 we need to recall some results from [OS1] and [OS2]. The following result restates Corollary 2.9 of [OS2], versions of which where already shown in [OS1]. Theorem 3.3. [OS2, Corollary 2.9 (c) ⇐⇒ (d), and "Moreover"-part] Let X be a subspace of a reflexive space Z with an F DD (Ei) and let Then the following are equivalent. A ⊂ {(xn) : xn ∈ SX for n ∈ N }. a) For any ε = (εn) ⊂ (0, 1) every weakly null tree in SX admits a branch in Aε, where Aε =(cid:8)(xn) ⊂ SX : ∃(zn)∈A kzn − xnk ≤ εn for n ∈ N(cid:9), and where Aε denotes the closure in the product of the discrete topology on SX. b) For any ε = (εn) ⊂ (0, 1) there is a blocking (Fi) of (Ei) so that every cε-skipped block sequence (xn) ⊂ SX of (Fi) lies in Aε. Here c ∈ (0, 1) is a constant which only depends on the projection constant of (Ei) in Z. We also need a blocking lemma which appears in various forms in [KOS], [OS1], [OS2] [OSZ] and ultimately results from a blocking trick of W. B. Johnson [J]. In the statement of Lemma 3.4 (and elsewhere) reference is made to the weak∗-topology of Z, a space with a boundedly complete FDD (Ei). By this we mean the weak∗-topology on Z obtained by regarding it as the dual space of the norm closure of the span of (E∗ i ) in Z ∗. This is then just the topology of coordinatewise convergence in Z with respect to the coordinates of (Ei). Lemma 3.4. [OS2, Lemma 3, Section 3] Let X be a subspace of a space Z having a boundedly complete FDD E = (Ei) with projection constant K with BX being a w∗-closed subset of Z. Let δi ↓ 0. Then there exist 0 = N0 < N1 < ··· in N with the following properties. For all x ∈ SX there exists (xi)∞ i=1 ⊆ X, and for all i ∈ N, there exists ti ∈ (Ni−1, Ni) satisfying (t0 = 0 and t1 > 1) j=1 xj, a) x =P∞ b) kxik < δi or kP E c) kP E (ti−1,ti)x − xik < δi, d) kxik < K + 1, e) kP E ti xk < δi. (ti−1,ti)xi − xik < δikxik, 10 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Proof of Theorem 3.2. Assume X embeds in a reflexive space Z with an FDD E = (Ei). By Zippin's theorem [Z] such a space Z always exists. After renorming we can assume that the projection constant K = supm≤n kP E [m,n]k = 1 and that X is (isometrically) a subspace of Z. We also assume without loss of generality that kTk = 1. For a sequence x = (xi) ∈ SX and a =P aiei ∈ c00 we define =(cid:13)(cid:13)(cid:13)X aivi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T(cid:16)X aixi(cid:17)(cid:13)(cid:13)(cid:13)Y Then · x is a norm on c00 and we denote the completion of c00 with respect to · x by Wx. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X aiei(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x . Define A =(cid:26)x = (xn) ⊂ SX : x is 3 2 -basic and 3 2 C-equivalent to (ei) in Wx (cid:27) . Observe that condition a) of Theorem 3.3 is satisfied for this set A. Indeed, given any weakly null tree in SX we may assume, as noted before the statement of Theorem 3.1 that, by passing to a full subtree, the branches are basic with a constant close to 1, and, thus the first requirement of the definition of A can be satisfied. The hypothesis from Theorem 3.2 then guarantees that Aε contains the required branch. We first choose a null sequence ε = (εi) ⊂ (0, 1), which decreases fast enough to 0 to ensure that every sequence x = (xn) in Aε is 2-basic and 2C equivalent to (ei) in Wx. By Theorem 3.3 applied to ε we can find a blocking F = (Fi) of (Ei) and a sequence, so that every cε-skipped block sequence (xi) ⊂ SX of (Fi) (c is the constant in Theorem 3.3 (b)) is 2-basic and 2C-equivalent to (ei) in Wx. We put δ = (δi) = cε. Then we apply Lemma 3.4 to get a further blocking (Gi), Gi = ⊕Ni j=Ni−1+1Fj, for i ∈ N and some sequence 0 = N0 < N1 < N2 . . ., so that for every x ∈ SX there is a sequence (ti) ⊂ N , with ti ∈ (Ni−1, Ni) for i ∈ N, and t0 = 0, and a sequence (xi) satisfying (a)-(e). We also may assume that P∞ 36C∼ (cid:16)(cid:13)(cid:13)(cid:13) i (x)kvi(cid:13)(cid:13)(cid:13)V(cid:17) ∨ kT (x)kY . i=1 δi < 1/36C and will show that for every x ∈ X Xi=1 kxkX This implies that the map X → (⊕Gi)V ⊕ Y, embedding. Let x ∈ SX and choose (ti) ⊂ N and (xi) ⊂ X as prescribed in Lemma 3.4. Letting B =(cid:8)i ≥ 2 : kP F (ti−1,ti)(xi) − xik ≤ δikxik(cid:9) it follows that (xi/kxik)i∈B is a δ-skipped block i (x)), T (x)), is an isomorphic sequence of (Fi) and therefore x 7→ ((P G kP G (3.6) ∞ (3.7) xi(cid:13)(cid:13)(cid:13)X (cid:13)(cid:13)(cid:13)Xi∈B We want to estimate(cid:13)(cid:13)P∞ i=1 kxikvi(cid:13)(cid:13)V ∨ kT (x)k. Since 1 6∈ B (no matter how large kx1k is) we will distinguish between the case that kx1k is essential and the case that kx1k is small enough to be discarded. kxikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T(cid:16)Xi∈B 2C∼ (cid:13)(cid:13)(cid:13)Xi∈B xi(cid:17)(cid:13)(cid:13)(cid:13). SMALL SUBSPACES OF Lp 11 If kx1k ≥ 1/8C then we deduce that (3.8) ∞ 1 ∞ ∞ kxikvi(cid:13)(cid:13)(cid:13)V ∨ kT (x)kY + kx1k +Xi6∈B 8C ≤ kx1k ≤(cid:13)(cid:13)(cid:13) Xi=1 kxikvi(cid:13)(cid:13)(cid:13)V ≤(cid:16)(cid:13)(cid:13)(cid:13) Xi∈B xi(cid:13)(cid:13)(cid:13) + 2 +X δi ≤ 2C(cid:13)(cid:13)(cid:13) Xi∈B ≤ 2Ckxk + 2C(cid:13)(cid:13)(cid:13) ≤ 2Ckxk + 2Ckx1k + 2CX δi + 2 +X δi ≤ 9C. xi(cid:13)(cid:13)(cid:13) + 2 +X δi Xi6∈B ∞ δi(cid:17) ∨ kT (x)kY [by (3.7), (d) of Lemma 3.4] and since kTk = 1] [By (3.7)] ∞ Xi=1 kxikvi(cid:13)(cid:13)(cid:13)V ∨ kT (x)kY(cid:17) + 3 4 If kx1k < 1/8C then 1 = kxk ≤(cid:13)(cid:13)(cid:13)Xi∈B ≤ 2C(cid:16)(cid:13)(cid:13)(cid:13)Xi∈B ≤ 2C(cid:16)(cid:13)(cid:13)(cid:13) Xi=1 ∞ and, thus, (3.9) 1 1 ∞ + 1 2 1 4C 1 4C xi(cid:17)(cid:13)(cid:13)(cid:13)Y(cid:17) + xi(cid:13)(cid:13)(cid:13) + kxikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T(cid:16)Xi∈B kxikvi(cid:13)(cid:13)(cid:13)V ∨ kT (x)kY(cid:17) + 4C ≤ 2C(cid:16)(cid:13)(cid:13)(cid:13) kxikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T (x)(cid:13)(cid:13)(cid:13)Y 8C ≤(cid:13)(cid:13)(cid:13) Xi=1 kxikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)(cid:13)T(cid:16)Xi∈B ≤(cid:16)(cid:13)(cid:13)(cid:13)Xi∈B xi(cid:13)(cid:13)(cid:13) + ≤ 2C(cid:13)(cid:13)(cid:13)Xi∈B ≤ 2Ckxk + 2Ckx1k + 2CX δi + kxikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)T (x)(cid:13)(cid:13). 9C∼ (cid:13)(cid:13)(cid:13) Xi=1 [By (3.7)] 1 4C ∞ 1 1 4C xi(cid:17)(cid:13)(cid:13)(cid:13)Y(cid:17) + 1 4C ≤ 8C. (3.8) and (3.9) imply that (3.10) (3.11) For n ∈ N define yn = P F kP F (tn−1,tn)(x) − xnk + kP F (3.10) that (tn−1,tn](x). From Lemma 3.4 (c) and (e) it follows that kyn− xnk ≤ tn (x)k ≤ 2δn and thus Pkyn − xnk ≤ 1/18C which implies by 1 18C∼ (cid:13)(cid:13)(cid:13) kyikvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)T (x)(cid:13)(cid:13). Xi=1 ∞ 12 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Since for n ∈ N we have (Nn−1, Nn] ⊂ (tn−1, tn+1) and and (tn−1, tn] ⊂ (Nn−2, Nn) (put N−1 = N0 = 0 and P G 0 = 0) it follows from the assumed 1-subsymmetry of (vn) and the assumed bimonotonicity of (Ei) in Z that 1 1 2(cid:13)(cid:13)(cid:13)Xn∈N kynkvn(cid:13)(cid:13)(cid:13)V ≤ kP G n−1(x)k + kP G n (x)k)vn(cid:13)(cid:13)(cid:13)V n (x)kvn(cid:13)(cid:13)(cid:13)V (tn−1,tn+1)(x)(cid:13)(cid:13)vn(cid:13)(cid:13)(cid:13)V 2(cid:13)(cid:13)(cid:13)Xn∈N(cid:0)kP G ≤(cid:13)(cid:13)(cid:13)Xn∈N ≤(cid:13)(cid:13)(cid:13)Xn∈N(cid:13)(cid:13)P F ≤(cid:13)(cid:13)(cid:13)Xn∈N(cid:0)kynk + kyn+1k(cid:1)vn(cid:13)(cid:13)(cid:13)V ≤ 2(cid:13)(cid:13)(cid:13)Xn∈N i (x)kvi(cid:13)(cid:13)(cid:13)V ∨(cid:13)(cid:13)T (x)(cid:13)(cid:13). 36C∼ (cid:13)(cid:13)(cid:13) ∞ Xi=1 kP G 1 which implies with (3.11) that and finishes the proof of our claim. , kynkvn(cid:13)(cid:13)(cid:13)V (cid:3) 4. Embedding small subspaces in ℓp ⊕ ℓ2 For a subspace X of Lp (where p > 2, as everywhere in this paper) we shall say that a function v in Lp/2 is a limiting conditional variance associated with X if there is a weakly null sequence (xn) in X such that x2 n converges to v in the weak topology of Lp/2. It is equivalent to say that, for all E ∈ Σ (recall that Lp was defined over the atomless and separable probability space (Ω, Σ, P)) E[1Ex2 n] → E[1Ev] as n → ∞. The set of all such v will be denoted V (X). Note that, because p > 2, every weakly null sequence (xn) in X does of course have a subsequence (xnk ) such that x2 nk converges (to some v ∈ V (X)) for the weak topology of the reflexive space Lp/2. Limiting conditional variances occur naturally in the context of the martingale inequal- ities to be used in this section, and are closely related to the random measures of Section 6. It is therefore natural to express the basic dichotomy underlying our main Theorem B in terms of V (X). Proposition 4.1. Let X be a subspace of Lp, where p > 2. One of the following is true: (A) there is a constant M > 0 such that kvkp/2 ≤ Mkvk1 for all v ∈ V (X); (B) no such constant M exists, in which case there exist disjoint sets Ai ∈ Σ and elements vi ∈ V (X) (i ∈ N), such that k1Ai vikp/2 → 1 and k1Ω\Ai vikp/2 → 0 as i → ∞. Proof. This is a consequence of the Kadets -- Pe lczynski dichotomy. Either there exists an ε > 0 so that V (X) ⊂(cid:8)u ∈ Lp/2 : P[u ≥ εkukp/2] ≥ ε(cid:9) then kuk1 ≥ E(cid:2)εkukp/21[u≥εkukp/2](cid:3) ≥ ε2kukp/2, for all u ∈ V (X), and (A) holds for M = ε−2. Otherwise, by the construction in Theorem 2 of [KP], we obtain (B). (cid:3) SMALL SUBSPACES OF Lp 13 The rest of this section will be devoted to showing that if (A) holds then X embeds in ℓp ⊕ ℓ2. By Theorem 3.1, it will be enough to prove the following proposition. Proposition 4.2. Let X be a subspace of Lp, where p > 2, and assume that (A) holds in Proposition 4.1. Then there is a constant K such that every weakly null tree in SX has a branch (xi) satisfying K −1(cid:13)(cid:13)(cid:13)X cixi(cid:13)(cid:13)(cid:13)p ≤ maxn(cid:16)Xcip(cid:17)1/p ,(cid:13)(cid:13)(cid:13)X cixi(cid:13)(cid:13)(cid:13)2o ≤ K(cid:13)(cid:13)(cid:13)X cixi(cid:13)(cid:13)(cid:13)p for all ci ∈ R. Proof. Our proof, using Burkholder's martingale version of Rosenthal's Inequality (Theorem 2.3), is closely modeled on Theorem 1.14 of [JMST]. Let (xα)α∈T∞ be a weakly null tree in SX. Taking small perturbations, we may suppose that we are dealing with a block tree of the Haar basis. So for each α ∈ T∞, xα is a finite linear combination of Haar functions, say xα ∈ [hn]n≤n(α), and for each successor (α, k) of α in T∞, x(α,k) ∈ [hn]n(α)<n≤n(α,k). We may then proceed to choose a full subtree T ′ of T∞ having the properties (1) and (2), below, as we now describe. , First, we consider the first level of the tree, that is to say the sequence of elements x(n) (n) converges weakly in Lp/2 to some (n)]1/2 − with n ∈ N. We may extract a subsequence for which x2 v0 ∈ V (X) and then, by leaving out a finite number of terms, ensure that E[x2 E[v0]1/2 < 1 2 . the following hold (for n ∈ N, Hn denotes the σ-algebra generated by (hi : i≤ n)) : We now continue by taking subsequences of the successors of each α in such a way that (α,n) (with (α, n) ∈ T ′) of Lp/2 converge weakly to some vα ∈ V (X); (1) the elements x2 (2) for all (α, k) ∈ T ′ we have kE[x2 (α,k) Hn(α)]1/2 − E[vα Hn(α)]1/2k∞ < 2−α−1. To achieve the above, we use our earlier remark based on relexivity of Lp/2, and the fact that weak convergence implies norm convergence in the finite dimensional space [hn]n≤n(α). We now take any branch (xi) of the resulting subtree (xα)α∈T ′. So xi = xαi where αi is the initial segment (n1, n2, . . . , ni) of some branch (n1, n2, . . . ) of T ′. We consider the σ-algebras Fi where F0 = {∅, Ω} and Fi = Hn(αi) for i ≥ 1 and write Ei for the conditional expectation relative to Fi. Since we are dealing with a block tree the sequence (xi) is a block basis of the Haar basis, and hence a martingale-difference sequence with respect to (Fi). We may therefore apply Theorem 2.3 to conclude that the Lp-norm of a linear combination P cixj is Cp-equivalent to 1/2 We shall show that, provided we modify the constant of equivalence, we may replace the second term in this expression by Ei−1[x2 i ](cid:13)(cid:13)(cid:13) p/2o. maxn(cid:16)Xcip(cid:17)1/p i ,(cid:13)(cid:13)(cid:13)X c2 i ](cid:13)(cid:13)(cid:13) Ei−1[x2 1/2 , 1 i (cid:13)(cid:13)(cid:13)X c2 Now, by construction, the conditional expectations Ei−1[x2 which equals kP cixik2. i ] are close to Ei−1[vi−1], where, for j ≥ 1, vj denotes vαj . More precisely, we may use (2) above and the triangle inequality Using our assumption about V (X), the fact that all the vi are non-negative and the in- equalities (4.1) and (4.2) we obtain i 2−2i(cid:1)1/2(cid:13)(cid:13)(cid:13)p≤ maxci. i 2−2i)1/2(cid:13)(cid:13)(cid:13)2 ≤ maxci. + maxci + maxci 1/2 + max ci +(cid:0)1 + √M(cid:1) max ci 14 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT in Lp(ℓ2) to obtain We similarly get (4.1) (4.2) i i (cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)X c2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)X c2 (cid:13)(cid:13)(cid:13)X c2 i 1/2 Ei−1[x2 Ei−1[x2 1/2 Ei−1[x2 i 1/2 1/2 1/2 1 i i 1/2 p/2(cid:12)(cid:12)(cid:12)(cid:12)≤(cid:13)(cid:13)(cid:13)(cid:0)X c2 (cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:13)(cid:13)(cid:13)(X c2 p/2−(cid:13)(cid:13)(cid:13)X c2 i ](cid:13)(cid:13)(cid:13) 1 −(cid:13)(cid:13)(cid:13)X c2 i ](cid:13)(cid:13)(cid:13) p/2 ≤(cid:13)(cid:13)(cid:13)X c2 i ](cid:13)(cid:13)(cid:13) ≤(cid:16)X c2 ≤(cid:16)X c2 √M(cid:16)X c2 ≤ = √M(cid:13)(cid:13)(cid:13)X c2 √M(cid:13)(cid:13)(cid:13)X c2 Ei−1[vi−1](cid:13)(cid:13)(cid:13) Ei−1[vi−1](cid:13)(cid:13)(cid:13) Ei−1[vi−1](cid:13)(cid:13)(cid:13) i kEi−1[vi−1]]kp/2(cid:17)1/2 i(cid:13)(cid:13)vi−1(cid:13)(cid:13)p/2(cid:17)1/2 i(cid:13)(cid:13)vi−1(cid:13)(cid:13)1(cid:17)1/2 Ei−1[vi−1](cid:13)(cid:13)(cid:13) i ](cid:13)(cid:13)(cid:13) Ei−1[x2 + maxci + maxci 1 1/2 ≤ p/2 i i 1 which yields the left most inequality in Proposition 4.2. The right hand inequality is easy by Proposition 2.1 since k · kp ≥ k · k2 and (xi) is unconditional, being a block basis of the Haar basis. (cid:3) Corollary 4.3. Let X be a subspace of Lp, where p > 2, and assume that (A) holds in Proposition 4.1. Then X embeds isomorphically into ℓp ⊕ ℓ2. 5. Embedding ℓp(ℓ2) in X Theorem 5.1. Let X be a subspace of Lp (p > 2) and suppose that (B) of Proposition 4.1 holds. Then X contains a subspace isomorphic to ℓp(ℓ2). The first step in the proof is to find ℓ2-subspaces of X which have "thin support". The precise formulation of this notion that we shall use in the present section is given in the following lemma. Lemma 5.2. Suppose that (B) of Proposition 4.1 holds. Then, for every M > 0 there is an infinite-dimensional subspace Y of X, on which the Lp and L2 norms are equivalent, but in such a way that kykp ≥ Mkyk2 for all y ∈ Y . Proof. By hypothesis, for every M ′ > 0 there exists v ∈ V (X) such that kvk1 = 1 and kvkp/2 > M ′2. There is a weakly null sequence (xn) in X such that x2 n converges weakly to v in Lp/2. By taking small perturbations of the xn's (with respect to the Lp-norm) and by noting that the Cauchy-Schwarz inequality yields kx2 − y2kp/2 ≤ kx − ykp · kx + ykp, for x and y ∈ Lp, we may suppose that (xn) is a block basis of the Haar basis. Since the sequence x2 n is positive and weakly convergent, kx2 nk1 = E[x2 n] → E[v] = kvk1 = 1. SMALL SUBSPACES OF Lp 15 We can thus assume that kxnk2 = 1 for all n. We may choose a natural number K such that kE[v HK]kp/2 > M ′2 and by discarding the first few elements of (xn) we have that xn ∈ [hk]k>K, for all n. The xn are martingale differences with respect to a subsequence Fn = Hk(n) of the Haar filtration (with k(0) = K). Taking a further subsequence, we may suppose that Because (xn) is a martingale difference sequence, we can apply Theorem 2.3 to conclude that If we use (5.1) and apply the triangle inequality in Lp(ℓ2) we obtain (5.1) (cid:13)(cid:13)(cid:13)X cnxn(cid:13)(cid:13)(cid:13)p ≥ C −1 (cid:13)(cid:13)(cid:13)X cnxn(cid:13)(cid:13)(cid:13)p ≥ C −1 ≥ C −1 = C −1 n E[x2 = C −1 n Fn−1]1/2(cid:13)(cid:13)∞ < 2−n, for all n. p (cid:13)(cid:13)k(cnE1/2[x2 (cid:13)(cid:13)E[v Fn−1]1/2 − E[x2 p (cid:13)(cid:13)(cid:13)(cid:16)X c2 p (cid:13)(cid:13)k(cnE1/2[x2 p (cid:16)(cid:13)(cid:13)k(cnE1/2[vFn−1] : n ∈ N)kℓ2(cid:13)(cid:13)p − k(cn2−n : n ∈ N)kℓ2(cid:17) p (cid:16)(cid:13)(cid:13)(cid:13)(cid:0)X c2 nFn−1](cid:17)1/2(cid:13)(cid:13)(cid:13)p nFn−1] : n ∈ N)kℓ2(cid:13)(cid:13)p E[vFn−1](cid:1)1/2(cid:13)(cid:13)(cid:13)p−(cid:0)X c2 n2−2n(cid:1)1/2(cid:17) ≥ (cid:13)(cid:13)(cid:13)X cnxn(cid:13)(cid:13)(cid:13)2 n(cid:17)1/2 =(cid:16)X c2 n . nFn−1] : n∈ N)kℓ2(cid:13)(cid:13)p. M ′ − 1 Cp (cid:16)X c2 n(cid:17)1/2 . On the other hand, in L2, the xn are orthogonal, whence Provided M ′ is chosen large enough, we have kykp ≥ Mkyk2 for all y ∈ [xn] as required. (cid:3) The next step is to show that we can choose our ℓ2-subspaces to have p-uniformly inte- grable unit balls. Recall that a subset A of Lp is said to be p-uniformly integrable if, for every ε > 0 there exists K > 0 such that kx1[x>K]kp < ε for all x ∈ A. We shall need the following standard martingale lemma. Lemma 5.3. Let (xn) be a martingale difference sequence that is p-uniformly integrable. Then the set of linear combinations of the xn's with ℓ2-normalized coefficients is also p- uniformly integrable. nkxnk2 n = 1, noting that kyk2 Proof. We assume that kxnk2 ≤ 1 for all n and consider a vector y of the form Pn cnxn withPn c2 2 ≤ 1. Given ε > 0, we choose K > ε−1 such that kxj1Ek2 < ε for all j whenever P(E) < K −1. We consider the martingale (yn) where yn =Pj≤n cjxj (thus y = y∞) and introduce the stopping time By Doob's inequality P[τ < ∞] ≤ K −1kyk1 ≤ K −1. We note that if τ < ∞, then yτ ≤ K + cτ xτ so that 2 =P c2 τ = inf{n ∈ N : yn > K}. y ≤ K + y − yτ + cτ xτ 1[τ <∞]. We shall estimate the Lp-norms of the second two terms. For the first of these, we note that (yk − yk∧τ ) is a martingale, so that (C only depends on p) 16 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT 1/2 p/2 c2 nx2 n1[τ <n](cid:13)(cid:13)(cid:13) ky − yτkp ≤ C(cid:13)(cid:13)(cid:13)Xn n1[τ <n]kp/2(cid:17)1/2 ≤ C(cid:16)X c2 nkx2 (cid:2)since X c2 n ≤ 1(cid:3) ≤ C sup [because P[τ < ∞] ≤ K −1]. ≤ Cε n kxn1[τ <∞]kp [by the square function inequality] [by the triangle inequality in Lp/2] For the second term we use the fact that the sets [τ = n] are disjoint, so that kcτ xτ 1[τ <∞]kp =(cid:13)(cid:13)(cid:13)Xn as before. Thus, cnxn1[τ =n](cid:13)(cid:13)(cid:13)p =(cid:16)Xn cnpkxn1[τ =n]kp p(cid:17)1/p ≤ sup n kxn1[τ <∞]kp ≤ ε ky1[y>2K]kp ≤ K P1/p(cid:2)y − yτ+cτ xτ 1[τ <∞] > K(cid:3) + (C + 1)ε ≤ 2(1 + C)ε, which implies our claim. (cid:3) Lemma 5.4. Let Y be a subspace of Lp (p > 2), which is isomorphic to ℓ2. There is an infinite dimensional subspace Z of Y such that the unit ball BZ is p-uniformly integrable. Proof. Let (yn) be a normalized sequence in Y equivalent to the unit vector basis of ℓ2. By the Subsequence Splitting Lemma (see, for instance Theorem IV.2.8 of [G-D]), we can write yn = xn + zn, where the sequence (xn) is p-uniformly integrable, and the zn are disjointly supported. So (xn) and (zn) are weakly null. Taking a subsequence, we may suppose that the (xn) is a martingale difference sequence, so that the set of all ℓ2-normalized linear We now consider ℓ2-normalized blocks of the form combinations P cnxn is also p-uniformly integrable. k = (Nk − Nk−1)−1/2 XNk−1<n≤Nk y′ yn = x′ k + z′ k, where, x′ xn and z′ Because the zn are disjointly supported in Lp we have kz′ k = (Nk − Nk−1)−1/2 XNk−1<n≤Nk we can choose the Nk such that kz′ linear combinations of the xn, are p-uniformly integrable. Hence the y′ perturbations of the x′ yields the result. kkp ≤ (Nk − Nk−1)1/p−1/2, so k), being ℓ2 normalized k, which are small k, are also p-uniformly integrable. Another application of Lemma 5.3 (cid:3) k = (Nk − Nk−1)−1/2 XNk−1<n≤Nk kkp < 2−k. The sequence (x′ zn. We are now ready for the proof of Theorem 5.1. Proof of Theorem 5.1. By Lemmas 5.2 and 5.4 there exists, for each M > 0, a subspace ZM of X, isomorphic to ℓ2 with p-uniformly integrable unit ball, such that kykp ≥ Mkyk2 SMALL SUBSPACES OF Lp 17 for all y ∈ ZM . For a specified ε > 0, we shall choose inductively M1 < M2 < ··· and define Yn = ZMn, such that kym ∧ ynkp ≤ ε/n2n, To achieve this, we start by taking an arbitrary value for M1, say M=1. Recursively, if (5.2) whenever ym ∈ BYm, yn ∈ BYn and m < n. M1, . . . , Mn have been chosen, we use the p-uniform integrability of Sm≤n BYm to find Kn such that(cid:13)(cid:13)y − y ∧ Kn(cid:13)(cid:13)p < ε/(n + 1)2n+2 whenever y ∈ BYm and m ≤ n. We now choose We need to check that (5.2) is satisfied, so let yn+1 ∈ BYn+1 and let ym ∈ BYm with Mn+1 such that M 2 (n + 1)p2p(n+2)ε−p. n+1 > K p−2 n m ≤ n. We have and have chosen Kn in such a way as to ensure that ym ∧ yn+1 ≤ Kn ∧ yn+1 + (ym − ym ∧ Kn) For the first term, we note that (cid:13)(cid:13)ym − ym ∧ Kn(cid:13)(cid:13)p < ε/(n + 1)2n+2. E[(Kn ∧ yn+1)p] ≤ E[K p−2 n yn+12] = K p−2 n kyn+1k2 2 ≤ K p−2 n M −2 n+1, which is smaller than εp(n + 1)−p2−p(n+2), by our choice of Mn+1. Now let yn ∈ SYn for all n ∈ N. We shall show that the yn's are small perturbations of elements that are disjoint in Lp. Indeed, let us set n are disjointly supported and from (5.2) Then the y′ kyn − y′ Standard manipulation of inequalities now shows us that the closure of the sum Pn Yn in Lp is almost an ℓp-sum. Indeed, ε/m2m < ε/2n. nkp =(cid:13)(cid:13)(cid:13)yn ∧ _m6=n (1 − 2ε)(cid:16)Xcnp(cid:17)1/p y′ ym(cid:1). n = sign (yn)(cid:0)yn − yn ∧ _m6=n ym(cid:13)(cid:13)(cid:13)p ≤ Xm6=n(cid:13)(cid:13)yn ∧ ym(cid:13)(cid:13)p ≤ (n − 1)ε/n2n + Xm>n p(cid:17)1/p ≤(cid:16)Xcnpky′ nkp =(cid:13)(cid:13)(cid:13)X cny′ n(cid:13)(cid:13)(cid:13)p − ε(cid:16)Xcnp(cid:17)1/p ≤(cid:13)(cid:13)(cid:13)X cnyn(cid:13)(cid:13)(cid:13)p ≤(cid:13)(cid:13)(cid:13)X cny′ n(cid:13)(cid:13)(cid:13)p + ε(cid:16)Xcnp(cid:17)1/p − ε(cid:16)Xcnp(cid:17)1/p ≤ (1 + ε)(cid:16)Xcnp(cid:17)1/p . At this point in the proof, we have obtained subspaces Yn of X, each isomorphic to ℓ2 such that the closed linear span Pn Yn is almost isometric to (L Yn)p. By stability ([KM] or [AO]) we can take, for each n, a subspace Xn of Yn which is (1 + ε)-isomorphic to ℓ2. In this way we obtain a subspace of X which is almost isometric to ℓp(ℓ2). (cid:3) The last part of the claim of Theorem B, namely that we can pass to a further subspace of X which is still (1+θ)-isomorphic to ℓp(ℓ2) and, moreover, complemented in Lp follows from our results in the next section. G. Schechtman [S2] showed us that if one is not concerned 18 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT with minimizing the norm of the projection, then there is a short argument that gives a complemented copy of ℓp(ℓ2). We thank him for allowing us to present it here. Proposition 5.5. Let X ⊂ Lp be isomorphically equivalent to ℓp(ℓ2). Then there is a subspace Y of X which is isomorphic to ℓp(ℓ2) and complemented in Lp. Proof. Let {x(m, n) : m, n ∈ N} ⊂ X be a normalized basis of X equivalent to the usual unconditional basis of ℓp(ℓ2), i.e. there is a constant C ≥ 1 so that (cid:13)(cid:13)(cid:13) Xm,n∈N a(m, n)x(m, n)(cid:13)(cid:13)(cid:13) C∼ (cid:16)Xm∈N(cid:16)Xn∈N a(m, n)2(cid:17)p/2(cid:17)1/p In [PR] it was shown that for any C > 1 there is a gp(C) < ∞ so that every subspace E of Lp, which is C isomorphic to ℓ2, is gp(C) complemented in Lp. For m ∈ N let Pm : Lp → [(x(m, n) : n ∈ N] be a projection of norm at most gp(C). We can write for all (a(m, n)) ∈ c00(N2). Pm(x) =Xn∈N x∗(m, n)(x)x(m, n) for x ∈ Lp, where (x∗(m, n) : n ∈ N) is a weakly null sequence in Lq, 1 q = 1, and biorthogonal to x(m, n) : n ∈ N). By passing to subsequences, using a diagonal argument, and perturbing we may assume that there is a blocking (H(m, n) : m, n ∈ N) of the Haar basis of Lp, in some order, so that x(m, n) ∈ H(m, n) and x∗(m, n) ∈ H ∗(m, n), for m, n ∈ N, where (H ∗(m, n)) denotes the blocking of the Haar basis in Lq which corresponds to (H(m, n)) p + 1 We will show that the operator P : Lp → Lp, x 7→ Xm,n∈N x∗(m, n)(x)x(m, n), is bounded and, thus, it is a bounded projection onto [x(m, n) : m, n ∈ N]. For y =Pm,n∈N y(m, n), with y(m, n) ∈ H(m, n), if m, n ∈ N, we deduce that kP (y)k =(cid:13)(cid:13)(cid:13)Xm∈NXn∈N ≤ C(cid:16)Xm∈N(cid:16)Xn∈N ≤ C 2(cid:16)Xm∈N x∗(m, n)(y(m, n))x(m, n)(cid:13)(cid:13)(cid:13) (x∗(m, n)(y(m, n)))2(cid:17)p/2(cid:17)1/p kPm(ym)kp(cid:17)1/p ≤ C 2gp(C)(cid:0)Xm∈N kymkp(cid:17)1/p The Haar basis is unconditional in Lp, and if we denote the unconditional constant in Lp where ym =Pn∈N y(m, n) for m ∈ N. by Up we deduce from Proposition 2.1 that kyk ≥ U −1 claim. p (Pm∈N kymkp)1/p, which implies our Remark. G. Schechtman [S2] has also proved, by a more complicated argument, that if X ⊂ Lp, 1 < p < 2, is an isomorph of ℓp(ℓ2) then X contains a copy of ℓp(ℓ2) which is complemented in Lp. (cid:3) Let us now deduce the statement of Corollary D. Proof of Corollary D. First assume that X embeds into ℓp⊕ℓ2. Note that every weakly null sequence (xn) can be turned into a weakly null tree (xα), whose branches are exactly the SMALL SUBSPACES OF Lp 19 subsequences of (xn) (put x(n1,n2,...nℓ) = xnℓ for (n1, n2, . . . nℓ) ∈ T∞). This fact, together with the remarks at the beginning of the proof of Theorem A (about the existence of K), show that condition (b) of Theorem A for a subspace X of Lp implies, that there exists a K ≥ 1, so that every weakly null sequence in SX admits a subsequence (xi) satisfying for all scalars (ai) condition (1.1) in (b) of Theorem A. Conversely, assume that X does not embed into ℓp ⊕ ℓ2. Then Propositions 4.1 and 4.2 together with Theorem A imply that condition (B) of Proposition 4.1 is satisfied. Using now Lemma 5.2, we can find for every M < ∞ a subspace Y of X which is isomorphic to ℓ2, so that k · kp ≥ Mk · k2 on Y . This implies that there cannot be a K ≥ 1, so that every weakly null sequence in SX admits a subsequence (xi) satisfying (1.4). (cid:3) 6. Improving the embedding via random measures We shall give a quick review of what we need from the theory of stable spaces and random measures. We shall then obtain the optimally complemented embeddings of ℓp(ℓ2). We start this section by recalling some facts about random measures and their relation to types on Lp. The introductory part is valid for 1 < p < ∞. Later we will restrict ourselves again to the case p > 2. As far as possible, we shall follow the notation and terminology of [Ald]; for the theory of types and stability we refer the reader to [KM] (or [AO]). The lecture notes of Garling [G] is one of the few works where the connection between random measures and types on function spaces is explicitly considered. We shall denote by P the set of probability measures on R which is a Polish space for its usual topology. This topology, often called the "narrow topology", can be thought of as the topology induced by the weak* topology σ(Cb(R)∗,Cb(R)). A random measure on (Ω, Σ, P) is a mapping ξ : ω 7→ ξω; Ω → P which is measurable from Σ to the Borel σ-algebra of P. The set of all such random measures is denoted by M and is a Polish space when equipped with what Aldous calls the wm-topology. Sequential convergence for this topology can be characterized by saying that ξ(n) wm−→ ξ if and only if E(cid:20)1FZR f (t)dξ(n)(t)(cid:21) → E(cid:20)1FZR f (t)dξ(t)(cid:21) , for all F ∈ Σ and all f ∈ Cb(R). In interpreting the expectation operator in the above formula (and in similar expressions involving "implicit" ω's) the reader should bear in mind that ξ is random. If we translate the expectation into integral notation, E(cid:20)1FZR f (t)dξ(t)(cid:21) becomes ZFZR f (t) dξω(t) dP(ω). It is sometimes useful to use the notation ξF , when F is a non-null set in Σ for the probability measure given by ZR f (t) dξF (t) = P(F )−1E[1FZR f (t) dξ(t)] (f ∈ C0(R)). The usual convolution operation on P may be extended to an operation on M by defining ξ ∗ η to be the random measure with (ξ ∗ η)ω = ξω ∗ ηω. Garling (Proposition 8 of [G]) observes that this operation is separately continuous for the wm topology. This result is also implicit in Lemma 3.14 of [Ald]. We may also introduce a "scalar multiplication": when ξ ∈ M and α is a random variable, we define the random measure α.ξ by setting Z f (t) d(α.ξ)(t) =ZR f (αt) dξ(t) (f ∈ Cb(R)). 20 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Every random variable x on (Ω, Σ, P) defines a random (Dirac) measure ω 7→ δx(ω). Aldous [Ald, after Lemma 2.14] has remarked that (provided the probability space (Ω, Σ, P) is atomless) these δx form a wm-dense subset of M. While we do not need this fact here, it may be helpful to note that the definition given above of α.ξ is so chosen that wm−→ ξ. The Lp-norms extend to wm-lower semicontinuous δαxn [0,∞]-valued functions · p on M, defined by ξp = EhZR tp dξ(t)i1/p wm−→ α.ξ whenever δxn We shall write Mp for the set of all ξ for which ξp is finite. As a special case of the characterization of wm-compactness by the condition of "tight- ness" we note that a subset of Mp which is bounded for · p is wm-relatively compact. In particular, if (xn) is a sequence that is bounded in Lp then there is a subsequence (xnk ) wm−→ ξ for some ξ ∈ Mp. If (xn) is, moreover, p-uniformly integrable, an such that δxnk easy truncation argument shows that . n→∞kxnkp = lim lim n→∞ E(cid:16)Z tpdδxn(t)(cid:17) = E(cid:16)Z tpdξ(t)(cid:17). For a subspace X of Lp we write Mp(X) for the set of all ξ that arise as wm-limits of sequences (δxn) with (xn) an Lp-bounded sequence in X. It is an easy consequence of separate continuity that Mp(X) is closed under the convolution operation ∗ (c.f. the proof of [Ald, Proposition 3.9]). We recall that a function τ : X → R on a (separable) Banach space X is called a type if there is a sequence (xn) in X such that, for all y ∈ X, kxn + yk → τ (y) as n → ∞. The set of all types on X is denoted TX and is a locally compact Polish space for the weak w−→ τ if τn(y) → τ (y) for topology; this topology may be characterized by saying that τn all y ∈ X. If we introduce, for each x ∈ X, the degenerate type τx defined by τx(y) = kx + yk, then TX is the w-closure of the set of all τx. We introduce a "scalar multiplication" of types, defining α.τ , for α ∈ R and τ ∈ TX by setting α.τ = w-lim ταxn when τ = w-lim τxn. A Banach space X is stable if, for xm and yn in X, we have lim m→∞ n→∞kxm + ynk = lim lim n→∞ n→∞kxm + ynk, lim Stability of a Banach space X permits the introduction of a (commutative) binary oper- whenever the relevant limits exist. All Lp-spaces (1 ≤ p < ∞) are stable [KM]. ation ∗ on TX, defined by τ ∗ υ(z) = lim m→∞ n→∞kxm + yn + zk lim when τ = w-lim τxm and υ = w-lim τyn. A type τ ∈ TX is said to be an ℓq-type if (α.τ ) ∗ (β.τ ) = (αq + βq)1/q.τ for all real α, β. The big theorem of [KM] shows first that on every stable space there are ℓq-types for some value(s) of q, and secondly that the existence of an ℓq type implies that SMALL SUBSPACES OF Lp 21 the space has subspaces almost isometric to ℓq. In fact the proof of Th´eor`eme III.1 in [KM] proves something slightly more than the existence of such a subspace. We now record the statement we shall need. Proposition 6.1. Let X be a stable Banach space, let 1 ≤ q < ∞ and let (xn) be a sequence in X such that τxn converges to an ℓq-type τ on X. Then there is a subsequence (xnk ) such that τzn converges to τ for every ℓq-normalized block subsequence (zn) of (xnk ). The results of [KM] extended, and gave an alternative approach to the theorem of [Ald], which obtained ℓq's in subspaces of L1 using random measures. We shall need elements from both approaches. The link is provided by the following lemma, for which we refer the reader to the final paragraphs of [G]. We shall write Tp for TLp and, when X is a subspace of Lp, we shall write Tp(X) for the weak closure in Tp of the set of all τx with x ∈ X. Lemma 6.2. Let (xn) be a bounded sequence in Lp and suppose that δxn Suppose further that kxnkp → α as n → ∞. Then, for all y ∈ Lp p → E(cid:20)ZR y + tp dξ(t)(cid:21) + βp, wm−→ ξ in M. kxn + ykp where the non-negative constant β is given by αp = kξkp p + βp. The sequence (xn) is p-uniformly integrable if and only if β = 0. We thus have the following formula showing how the type τ = lim τxn ∈ Tp is related to the random measure ξ = wm- lim δxn ∈ Mp and the index of p-uniform integrability β. (6.1) τ (y)p = E(cid:20)ZR y + tpdξ(t)(cid:21) + βp. If q < p then a sequence (xn) as above in Lp can be thought of as a sequence in Lq. If we wish to distinguish the type determined on Lq from the type on Lp, we use superscripts. Of course, with no "β" term, because an Lp-bounded sequence is q-uniformly integrable. τ (q)(y)q = E(cid:20)ZR y + tqdξ(t)(cid:21) , The * operations on Tp and on Mp are related by the following lemma, also to be found in [G]. Lemma 6.3. Let τ1 and τ2 be types on Lp represented as τ1(y)p = E(cid:20)ZR y + tpdξ1(t)(cid:21) + βp 1 and τ2(y)p = E(cid:20)ZR y + tpdξ2(t)(cid:21) + βp 2 . Then (τ1 ∗ τ2)(y)p = E(cid:20)ZR y + tpd(ξ1 ∗ ξ2)(t)(cid:21) + βp 1 + βp 2 . It has been noted already in the literature (e.g. [G]) that the representation given in (6.1) is not in general unique. However, for most values of p, it is, as we now show. 22 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Proposition 6.4. Let 1 ≤ p < ∞ and assume that p is not an even integer. In the representation of a type τ on Lp by the formula (6.1) the random measure ξ and the constant w−→ τ we have β are uniquely determined by τ . If (xn) is any sequence in Lp with τxn δxn Proof. Suppose that ξ, β and ξ′, β′ yield the same type τ . For any non-null E ∈ Σ and any real number u, we consider τ (y) where y = u1E ∈ Lp to obtain wm−→ ξ and inf M limn→∞ kxn1[xn≥M ]kp = β. E(cid:20)ZR t + u1Epdξ(t)(cid:21) + βp = E(cid:20)ZR t + u1Epdξ′(t)(cid:21) + β′p, or, equivalently, where ZR t + updξE(t) =ZR t + updξ′ E(t) + αp, P(E)αp = β′p − βp + E(cid:20)1Ω\EZ tpdξ′(t) − 1Ω\EZ tpdξ(t)(cid:21) . wm−→ ξ. (cid:3) Now let (xn) be any sequence with τxn By the Equimeasurability Theorem (cf.[KK, page 903]), α = 0 and the measures ξE and ξ′ E are equal. Since this is true for all E, ξ = ξ′. w−→ τ . By the uniqueness that we have just proved, the only cluster point of the sequence δxn in M is ξ. Since (by L1-boundedness) {δxn : n ∈ N} is relatively wm-compact in M, it must be that δxn We have already noted that Mp(X) is closed under ∗ when X is a subspace of Lp. The next proposition, which is closely related to that of [Ald, Proposition 3.9], shows that under appropriate conditions Mp(X) is wm-closed. Proposition 6.5. Let 1 ≤ p < ∞ and let X be a subspace of Lp with no subspace isomorphic to ℓp. Then Mp(X) is wm-closed in M. Proof. The hypothesis implies that the Lp-norm is equivalent to the L1-norm on X, so that we may regard X as a (reflexive) subspace of L1. Aldous [Ald, Lemma 3.12] shows (by a straightforward uniform integrability argument) that ξ 7→ ξ1 is wm-continuous and finite on D, where D is the wm-closure of {δx : x ∈ X}. Thus every ξ in D is in the wm-closure of an L1-bounded subset of X, and hence, by equivalence of norms, in Mp(X). (cid:3) To finish this round-up of types and random measures, we need to mention the connection between ℓ2-types and the normal distribution (a special case of the connection between ℓq- types and symmetric stable laws). We write γ for the probability measure (or law) of a standard N (0, 1) random variable. If σ is a non-negative random variable then σ.γ is a random measure (a normal distribution with random variance). Provided σ ∈ Lp this random measure defines a type on Lp by τ (y)p = E(cid:20)ZR y + tp d(σ.γ)(t)(cid:21) = E(cid:20)ZR y + σtp dγ(t)(cid:21) . Now it is a property of the normal distribution that (α.γ) ∗ (β.γ) = (α2 + β2)1/2.γ for real α, β. By Lemma 6.3, this allows us to see that τ is an ℓ2-type on Lp. We are finally ready to return to the main subject matter of this paper. Lemma 6.6. Let X be a subspace of Lp, with p > 2, and let v be a non-zero element of Lp/2. The following are equivalent: SMALL SUBSPACES OF Lp 23 (1) v ∈ V (X); (2) there exists ξ ∈ Mp(X) such that RR t dξ = 0 and RR t2 dξ = v almost surely; (3) √v.γ ∈ Mp(X). Proof. We start by assuming (1). Let (xn) be a weakly null sequence in X such that (x2 n) converges weakly to v in Lp/2. Replacing (xn) with a subsequence, we may suppose that wm−→ ξ for some ξ ∈ Mp(X). Since the sequence (xn) is Lp-bounded, it is 2-uniformly δxn integrable and so and (6.3) (6.2) tdξ(t)(cid:21) = lim E [1Exn] = 0 t2dξ(t)(cid:21) = lim E(cid:2)1Ex2 E(cid:20)1EZR E(cid:20)1EZR for all E ∈ Σ. This yields (2). convergent to ξ. SinceRR dξ(t) = 0 a.s. it follows that (xn) is weakly null and since ξ 6= δ0, kxnk2 does not tend to zero. By [KP], it follows that X0, the closed linear span of a subsequence of (xi), is isomorphic to ℓ2. The assumption about ξ is that, for almost all ω, the probability measure ξω is the law of a random variable with mean 0 and variance v(ω). We now assume (2). Let (xn) be an Lp-bounded sequence in X such that δxn is wm- n(cid:3) = E [1Ev] , By the Central Limit Theorem n−1/2. (ξω ∗ ξω ∗ ··· ∗ ξω) } n terms {z tends to pv(ω).γ for all such ω. So in M we have Since Mp(X0) is closed under convolution and is closed in the wm-topology (by Proposi- tion 6.5), we see that √v.γ ∈ Mp(X0) ⊆ Mp(X). Finally, if we assume (3) we may take (xn) to be an Lp-bounded sequence in X such n−1/2.(ξ ∗ ξ ∗ ··· ∗ ξ) wm−→ √v.γ. that δxn 2-uniform integrability, show that (xn) is weakly null and that x2 wm−→ √v.γ. Calculations like those used in the proof of (1) =⇒ (2), justified by n tends weakly to v. (cid:3) We shall say that a sequence (yn) in Lp is a stabilized ℓ2 sequence with limiting conditional variance v if, for every ℓ2 normalized block subsequence (zn) of (yn), the following are true: (6.4) δzn wm−→ √v.γ as n → ∞; kznkp → γpk√vkp as n → ∞. (6.5) (Recall that γp = kxkp, where x is a symmetric L2 normalized Gaussian random variable). For p not an even integer, it is not hard to establish the existence of such sequences using Proposition 6.1 and Proposition 6.4. The proof of the next proposition avoids the irritating problem posed by non-unique representations, by switching briefly to the L1-norm. Proposition 6.7. Let X be a closed subspace of Lp (p > 2) and let v be a non-zero element of V (X). Then there exists a stabilized ℓ2 sequence in X with limiting conditional variance v. Proof. By Lemma 6.6 the random measure √v.γ is in Mp(X). Let (xn) be a bounded wm−→ √v.γ. For the moment, think of the xn as elements of L1 and sequence in X with δxn 24 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT consider the types τ (1) integrable, so the sequence (τ (1) xn defined on L1. By Lp-boundedness, the sequence (xn) is uniformly xn ) converges weakly to the ℓ2-type τ (1), where τ (1)(y) = E(cid:20)Z y + √vtdγ(t)(cid:21) . By Proposition 6.1 we may replace (xn) by a subsequence in such a way that τ (1) zn for every ℓ2-normalized block subsequence (zn). By Proposition 6.4 we have δzn for all such (zn). w−→ τ (1) wm−→ √v.γ We now return to the Lp-norm, for which we can assume, after passing to a subsequence, if necessary, that (xn) is equivalent to the unit vector basis of ℓ2. By stability of Lp there w−→ τ (p) for some ℓ2-type τ (p) on is an ℓ2-normalized block subsequence (yn) such that τ (p) yn w−→ τ (p) for every further such Lp. Moreover, by Proposition 6.1 we can arrange that τ (p) zn ℓ2-normalized block subsequence (zn). By (6.1) we have τ (p)(y)p = E(cid:20)ZR y + √vtpdγ(t)(cid:21) + βp, for some non-negative constant β. Now τ (p) is an ℓ2-type, so τ (p) ∗ τ (p) = √2.τ (p). That is to say (τ (p) ∗ τ (p))(y)p = E(cid:20)ZR y + √2vtpdγ(t)(cid:21) + (√2β)p. On the other hand, by Lemma 6.3, (τ (p) ∗ τ (p))(y)p = E(cid:20)ZR y + √vtpd(γ ∗ γ)(t)(cid:21) + 2βp = E(cid:20)ZR y + √2vtpdγ(t)(cid:21) + 2βp. Since p 6= 2, we are forced to conclude that β = 0. δzn wm−→√vγ, since the zn are normalized blocks of (xn). But also To sum up, for every ℓ2-normalized block subsequence (zn) of (yn) we have, first of all, kznkp → τ (p)(0) = E(cid:20)ZR √vtpdγ(cid:21)1/p = γpk√vkp. (cid:3) Theorem 6.8. Let X be a subspace of Lp (p > 2) and assume that (B) of Proposition 4.1 holds. Then, for every θ > 0, there is a subspace Y of X which is (1 + θ)-isomorphic to ℓp(ℓ2) and a projection P from Lp onto Y with kPk ≤ (1 + θ)γp. Remark. The fact that Theorem 6.8 is the optimal result concerning the norm of a pro- jection onto a copy of ℓp(ℓ2) follows from [GLR, Theorem 5.12], where it was show that Lp contains subspaces isometric to ℓ2 which are γp complemented. Proof. Let ε ∈ (0, 1) be fixed and, for m ∈ N, let vm ∈ V (X), together with disjoint sets Am ∈ Σ, Am ⊂ supp(vm), be chosen so that kv1/2 m 1Amkp = 1 and kv1/2 m kp p < 1 + εp2−(m+2)p. Using Proposition 6.7 choose for each m a stabilized ℓ2-sequence (x(m) n )n∈N in X with limiting conditional variance vm. By (6.4) we have lim inf n→∞ E[ynp1Am] ≥ γp p and lim inf n→∞ E[y2 nv p 2 −1 m 1Am] ≥ 1 SMALL SUBSPACES OF Lp 25 and, by (6.5), pk√v1/2 m kp lim n→∞ E[ynp] = γp p < γp for all ℓ2-normalized block subsequences (yn) of (x(m) n ), starting at a suitably large value of n, we may suppose that the following hold for all ℓ2-normalized linear combinations y of the x(m) n : (6.6) p(1 + εp2−(m+2)p), n ). By relabeling the sequence (x(m) (6.7) p 2 −1 p ≥ (1 − ε2−(m+2)p)γp ky1Amkp m 1Ami ≥ 1 − ε2−m−1 p Ehy2v (6.8) Of course, (6.6) and (6.8) imply that the closed linear span Ym = [x(m) isometric to ℓ2; indeed, by homogeneity, they yield p ≤ (1 + ε2−(m+2)p)γp p . kykp n ]n∈N is almost n)1/2, (1 − ε2−(m+2)p)1/pγp(X c2 n)1/2 ≤ kykp ≤ (1 + ε2−(m+2)p)1/pγp(X c2 n ∈ Ym. Moreover, from the same inequalities we obtain ky − y1Amkp ≤ ε2−mkykp for all y ∈ Ym. when y =P cnx(m) (6.9) If, for each m ∈ N, ym is an element of SYm then y′ m = ym1Am are disjointly supported and are small perturbations of the ym. As in the proof of Theorem 5.1, we see that, by an appropriate choice of ε, we can arrange for the closure of Pm Ym in X to be (1 + θ)-isomorphic to ℓp(ℓ2). We are now ready to show that the subspace Y = Pm Ym is m 1Am; thus kφmk1 = 1. Let Φm : Lp → Lp(φm) be defined by complemented in Lp. We shall do this by combining the disjoint perturbation procedure used above with a standard "change-of-density" argument. For each m let φm = vp/2 Φm(f ) = 1Amφ−1/p m f, which is well defined since Am ⊂ supp(vm), and observe that kΦm(f )kLp(φm) = kf 1Amkp. Let Jm : Lp(φm) → L2(φm) be the standard inclusion and let Im : Ym → Lp be the natural embedding. We note that for y ∈ Ym L2(φm) = E[y2φ−2/p m φm1Am] = E[y2v m 1Am ] ≥ (1 − ε2−m)2γ−2 kJmΦmImyk2 p kyk2 p, p 2 −1 by (6.7), (6.8) and homogeneity. So if Wm is the image Wm = JmΦmIm[Ym] then Wm is closed in L2(φm) and the inverse mapping satisfies kRmk ≤ (1 − ε2−m)−1γp. We now introduce the orthogonal projections Rm = (JmΦmIm)−1 : Wm → Ym Pm : L2(φm) → Wm and consider Qm : Lp → Ym defined to be Qm = RmPmJmΦm. For f ∈ Lp we have XkQmfkp p ≤ (1 − ε)−pγp p ≤XkRmkp · kΦmfkp pXkf 1Amkp Lp(φm) ≤ (1 − ε)−pγp pkfkp p, 26 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT the last inequality following by disjointness of the sets Am. Since we already know that Y =P Ym is naturally isomorphic to (L Ym)p, we see that the series P Qmf converges to an element Qf of Y . Moreover, the operator Q thus defined satisfies kQk ≤ γp/(1 − ε). To finish, we investigate kQ(y) − ykp, when y = P yk with yk ∈ Yk. If, as before, we write y′ k) = 0 for m 6= k. Thus k = yk1Ak we may note that Qk(yk) = Qk(y′ k) and Qm(y′ [since Qkyk = yk] kQ(y) − ykp =(cid:13)(cid:13)(cid:13)Xk (cid:16)Xm =(cid:13)(cid:13)(cid:13)Xk Xm6=k =(cid:13)(cid:13)(cid:13)Xk Xm =(cid:13)(cid:13)(cid:13)Q(cid:16)Xk ≤ kQkXk Qmyk − yk(cid:17)(cid:13)(cid:13)(cid:13)p Qmyk(cid:13)(cid:13)(cid:13)p k)(cid:13)(cid:13)(cid:13)p k(cid:17)(cid:13)(cid:13)(cid:13)p kkp ≤ γp(1 − ε)−1X 2−kεkykkp, Qm(yk − y′ yk − y′ kyk − y′ using our estimate for kQk and (6.9) at the last stage. We can now see that for suitable chosen ε, Q may be modified to give a projection Q : Lp → Y with k Qk ≤ (1 + θ)γp. (cid:3) 7. Quotients and embeddings 7.1. Subspaces of Lp that are quotients of ℓp ⊕ ℓ2. It was shown in [JO2] that a subspace of Lp (p > 2) that is isomorphic to a quotient of a subspace of ℓp ⊕ ℓ2 is in fact isomorphic to a subspace of ℓp ⊕ ℓ2. We can give an alternative proof of this result by applying the main theorem of this paper. Clearly all that is needed is to show that ℓp(ℓ2) is not a quotient of a subspace of ℓp ⊕ ℓ2. We shall prove something more general, namely that ℓp(ℓq) is not a quotient of a subspace of ℓp ⊕ ℓq when p, q > 1 and p 6= q. By duality it will be enough to consider the case p > q. For elements w = (w1, w2) of ℓp ⊕ ℓq we shall write kwkp = kw1kp, kwkq = kw2kq and kwk = kwkp ∨ kwkq. Lemma 7.1. Let 1 < q < p < ∞ and let W be a subspace of ℓp ⊕ ℓq. Let X = ℓq, let Q : W → X be a quotient mapping and let λ be a constant with 0 < λ < kQk−1. For every M > 0 there is a finite-codimensional subspace Y of X such that, for w ∈ W we have kwk ≤ M, Q(w) ∈ Y, kQ(w)k = 1 =⇒ kwkq > λ. Proof. Suppose otherwise. We can find a normalized block basis (xn) in X and elements wn of W with kwnk ≤ M , Q(wn) = xn and kwnkq ≤ λ. Taking a subsequence and perturbing slightly, we may suppose that wn = w + w′ n) is a block basis in ℓp⊕ ℓq, satisfying kw′ n) = xn. We may now estimate as follows using the fact that the w′ nk ≤ M , kw′ Since Q(w) = w-lim Q(wn) = 0, we see that Q(w′ n, where (w′ nkq ≤ λ. n are disjointly supported: N Xn=1 w′ n(cid:13)(cid:13)(cid:13) =(cid:16) N Xn=1 (cid:13)(cid:13)(cid:13) kw′ p(cid:17)1/p nkp ∨(cid:16) N Xn=1 kw′ q(cid:17)1/q nkq ≤ N 1/pM ∨ N 1/qλ. Since the xn are normalized blocks in X = ℓq we have SMALL SUBSPACES OF Lp 27 N 1/q =(cid:13)(cid:13)(cid:13) N Xn=1 xn(cid:13)(cid:13)(cid:13) ≤ kQk(cid:13)(cid:13)(cid:13) N Xn=1 w′ n(cid:13)(cid:13)(cid:13) ≤ MkQkN 1/p ∨ λkQkN 1/q. Since λkQk < 1, this is impossible once N is large enough. Proposition 7.2. If 1 < q < p < ∞ then ℓp(ℓq) is not a quotient of a subspace of ℓp ⊕ ℓq. Proof. Suppose, if possible that there exists a quotient operator (cid:3) ℓp ⊕ ℓq ⊇ Z Q −→ X =(cid:16)Mn∈N Xn(cid:17)p where Xn = ℓq for all n. Let K be a constant such that Q[KBZ] ⊇ BX, let λ be fixed with 0 < λ < kQk−1, choose a natural number m with m1/q−1/p > Kλ−1, and set M = 2Km1/p. Applying the lemma, we find, for each n, a finite-codimensional subspace Yn of Xn such that (7.1) z ∈ M BZ, Q(z) ∈ Yn, kQ(z)k = 1 =⇒ kzkq > λ. ) be a sequence in Yn, 1-equivalent to the unit vector basis of ℓq. For For each n, let (e(n) each m-tuple i = (i1, i2, . . . , im) ∈ Nm, let z(i) ∈ Z be chosen with i Q(z(i) = e(1) i1 + e(2) i2 + ··· + e(m) im , and kz(,, , i)k ≤ Km1/p. exist in Z Taking subsequences in each co-ordinate, we may suppose that the following weak limits z(i1, i2, . . . , im−1) = w-limim→∞ z(i1, i2, . . . , im) ... z(i1, i2, . . . , ij ) = w-limij+1→∞ z(i1, i2, . . . , ij+1) ... z(i1) = w-limi2→∞z(i1, i2). Notice that, for all j and all i1, i2, . . . , ij, the following hold: Q(z(i1, . . . , ij) = e(1) i1 kz(i1, . . . , ij )k ≤ Km1/p + ··· + e(j) ij kz(i1, . . . , ij ) − z(i1, . . . , ij−1)k ≤ 2Km1/p = M. Since Q(z(i1, . . . , ij) − z(i1, . . . , ij−1)) = e(j) ij ∈ SYj it must be that (7.2) kz(i1, . . . , ij ) − z(i1, . . . , ij−1)kq > λ, [by (7.1)]. We shall now choose recursively some special ij in such a way that kz(i1, . . . , ij)kq > λj1/q for all j. Start with i1 = 1; since kz(i1)k ≤ M and Q(z(i1)) = e(1) we certainly have kz(i1)kq > λ by 7.1. Since z(i1, k) − z(i1) → 0 weakly we can choose i2 such that i1 28 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT z(i1, i2) − z(i1) is essentially disjoint from z(i1). More precisely, because of 7.2, we can ensure that kz(i1, i2)kq = kz(i1) + (z(i1, i2) − z(i1))kq > (λq + λq)1/q = λ21/q. Continuing in this way, we can indeed choose i3, . . . , im in such a way that kz(i1, . . . , ij )kq ≥ λj1/q. However, for j = m this yields λm1/q ≤ Km1/p, contradicting our initial choice of m. Remark. The proof we have just given actually establishes the following quantitative result: if Y is a quotient of a subspace of ℓp⊕ ℓq then the Banach-Mazur distance d(cid:0)Y,(cid:0)Lm j=1 ℓq(cid:1)p(cid:1) is at least m1/q−1/p. (cid:3) j=1 ℓ2(cid:17)p 7.2. Uniform bounds for isomorphic embeddings. As we remarked in the introduc- tion, the Kalton -- Werner refinement [KW] of the result of [JO1] gives an almost isometric embedding of X into ℓp when X is a subspace of Lp (p > 2), not containing ℓ2. By contrast, the main result of the present paper does not have an almost isometric version, and indeed it is easy to see that there is no constant K (let alone K = 1 + ε) such that every subspace of Lp not containing ℓp(ℓ2) K-embeds in ℓp ⊕ ℓ2. It is enough to consider spaces X of the . A straightforward argument, or an application of the more general form X =(cid:16)Lm result mentioned in the remark above, shows that the Banach -- Mazur distance from X to a subspace of ℓp ⊕ ℓ2 is at least m1/2−1/p. If we are looking for a "uniform" version of our Main Theorem, it is perhaps not un- reasonable to conjecture the existence of a constant K such that every subspace of Lp not containing ℓp(ℓ2) K-embeds in some space of the form ℓp ⊕p(cid:0)Lm constant M exists, as is shown by the following proposition. The structure of the space X considered below suggests that if there is some uniform version of our main result then it will involve independent sums (see [Als]), rather than, or as well as, ℓp sums. The proof of the next result follows a construction due to Alspach and could be compiled from arguments in [Als, Chapter 2]. The following is a self contained proof. j=1 ℓ2(cid:1)p. However, no such Proposition 7.3. Let p > 2. For every K > 0 there is a subspace X of Lp, isomorphic to ℓ2, such that for all m ∈ N, X is not K-isomorphic to a subspace of ℓp ⊕p(cid:0)Lm Proof. Fix a constant M > 1. Let {vi, zj,k : i, j, k ∈ N} be a family of independent random variables in Lp[0, 1] with distributions defined as follows: for i, j ∈ N, zi,j is N (0, 1), while vi is {0, M}-valued with P[vi = M ] = 1 − P[vi = 0] = M −p/2. We set xi,j = zi,j√vi, noting that l=1 ℓ2(cid:1)p. kxi,jkp p = E[vp/2 i zi,jp] = E[vp/2 i ]E[zi,jp] = γp p . We now define Xi = [xi,j]j∈N and X = [xi,j]i,j∈N. We start by calculating the norm of a general element of X. Let x = Pi,j ci,jxi,j. By independence, and properties of the normal distribution, the distribution of x, conditional on v1, v2, v3, . . . is N (0, w), where w =Pi,j c2 i,jvi. So (7.3) p = E(cid:2)E[xp v1, v2, . . . ](cid:3) = γp kxkp p E(cid:2)(Xi (cid:0)Xj c2 i,j(cid:1)vi)p/2(cid:3) = γp p/2 p/2, p(cid:13)(cid:13)X aivi(cid:13)(cid:13) SMALL SUBSPACES OF Lp 29 the unit vector basis of ℓ2. Indeed, Jensen's inequality yields i,j, for i ∈ N. Let us first note that (7.3) implies that (xi,j) is equivalent to p/2 ≥ Ep/2(cid:2)X aivi] =(cid:0)X aiM 1−p/2(cid:1)p/2 =(cid:0)M 1/2−p/4(cid:0)Xi,j i,j(cid:1)1/2(cid:1)p. c2 p/2 where ai =Pj c2 (cid:13)(cid:13)X aivi(cid:13)(cid:13) On the other hands, letting vi = vi − E(vi) = vi − M 1−p/2, the triangle inequality in Lp/2 and the fact that for some C < ∞ (depending on M and p) the sequence (vi), as sequence in Lp/2, is C-equivalent to the unit vector basis in ℓ2, implies and, thus, (cid:13)(cid:13)X aivi(cid:13)(cid:13)p/2 ≤ M 1−p/2X ai +(cid:13)(cid:13)X aivi(cid:13)(cid:13)p/2 ≤ M 1−p/2X ai + C(cid:0)X a2 p/2 ≤(cid:0)(M 1−p/2 + C)1/2(cid:0)Xi,j (cid:13)(cid:13)X aivi(cid:13)(cid:13) p/2 i(cid:1)1/2 ≤ (M 1−p/2 + C)X ai c2 i,j)1/2(cid:1)p, which finishes the proof of our claim that (xi,j) is equivalent to the unit basis of ℓ2. We note two special cases of (7.3). First, if x = xi ∈ Xi for some i (thus ci′,j = 0 for all i′ 6= i and all j), we have n n i=1 xi, n c2 c2 i,j)1/2. n Xi=1 i,j)1/2 = γ−1 kxikp = γp(Xj In particular, kxikp = 1 if and only if (Pj c2 where the xi are normalized elements of Xi, vi)p/2(cid:3)1/p = kn−1 Summarizing, we can say that if xi are Lp-normalized elements of Xi then p . Secondly, if x = n−1/2Pn Xi=1 Xi=1 (Xj i,j)vi)p/2(cid:3)1/p = n−1/2E(cid:2)( kxkp = n−1/2γpE(cid:2)( vik1/2 p/2. Now, by the weak law of large numbers, n−1Pn i=1 vi converges in probability to the constant E[v1] = M 1−p/2. Because these averages are uniformly bounded (by M ), the convergence holds also for the Lp/2-norm. So as n → ∞ we have vi(cid:13)(cid:13)(cid:13)p/2 → M 1−p/2. (cid:13)(cid:13)(cid:13)n−1 Xi=1 vi(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)n−1 (cid:13)(cid:13)(cid:13)n−1/2 Xi=1 ℓ=1 ℓ2(cid:1)p, with T0 : X → ℓp and Ti : X → ℓ2, for ℓ=0 : X → Y = ℓp ⊕p(cid:0)Lm ℓ = 1, 2 . . . , m, be an isomorphic embedding. We assume that kT (x)k ≥ kxk for all x and shall show that kTk ≥ M (p−2)/4. We note that, for each i, the sequence (cid:0)T0(xi,j))(cid:1)∞ j=1 is a weakly null sequence in ℓp. So by taking vectors of the form Xr=1 p/2 → M (2−p)/4 as n → ∞. x′ i,k = γ−1 p k−1/2 with jk−1(k − 1) < j1(k) < j2(k) < ··· < jk(k), we construct an Lp-normalized, weakly null sequence (x′ xi(cid:13)(cid:13)(cid:13)p Let T = (Tℓ)m n Xi=1 i,k)∞ k=1 in Xi with kT0(x′ i,k)kp → 0 as k → ∞. n 1/2 k xi,jr(k), (7.4) 30 R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT Passing to a subsequence, we may assume that for all i ∈ N and all ℓ = 1, 2 . . . m the i,k))kp → 0, ℓ=1. Passing to a subsequence in i, we may i,k)) tends to a limit µi,ℓ as k → ∞. Since kT (x′ i,k)k ≥ 1 and kT0(x′ . . . lim lim k1→∞ lim k2→∞ sequence Tℓ(x′ it must be that kµikp ≥ 1, where µi = (µi,ℓ)m assume that µi converges to some µ ∈ Rm, as i → ∞, with kµkp ≥ 1. For ℓ = 1, 2 . . . m and n ∈ N we observe that Xi=1 i,ki(cid:1)k2 kn→∞kn−1/2Tℓ(cid:0) n−1/2(cid:16)(cid:13)(cid:13)Tℓ(cid:0) i,ℓ(cid:17)1/2 Since µn → µ, as n → ∞, where µn = (µn,ℓ)m = . . . = n−1/2(cid:16) n Xi=1 i,ki(cid:1)(cid:13)(cid:13) ℓ=1, we deduce = lim k1→∞ ≡ µn,ℓ. n−1 Xi=1 lim k2→∞ kn−1→∞ . . . lim µ2 n x′ x′ 2 + µ2 i,ℓ(cid:17)1/2 (7.5) lim n→∞ lim k1→∞ lim k2→∞ On the other hand, as we have already noted above (7.4), . . . lim kn→∞(cid:13)(cid:13)n−1/2T(cid:0) Xi=1 vi(cid:13)(cid:13) i,ki(cid:13)(cid:13) =(cid:13)(cid:13)n−1 n x′ n Xi=1 (cid:13)(cid:13)n−1/2 n Xi=1 x′ i,ki(cid:1)(cid:13)(cid:13)Y = lim n→∞kµnkp = kµkp ≥ 1. 1/2 p/2 → M (2−p)/4, as n → ∞, Comparing this with (7.5), we conclude that kTk ≥ M (p−2)/4 as claimed. (cid:3) 8. Concluding Remarks A natural question remains, namely to characterize when a subspace X ⊆ Lp (2 < p < ∞) embeds into ℓp(ℓ2). We do not know the answer. In light of the [JO2] ℓp⊕ ℓ2 quotient result (see paragraph 7.1 above) we ask the following. Problem 8.1. Let X ⊆ Lp (2 < p < ∞). If X is a quotient of ℓp(ℓ2) does X embed into ℓp(ℓ2)? Extensive study has been made of the Lp spaces, i.e., the complemented subspaces of Lp which are not isomorphic to ℓ2 (see e.g., [LP] and [LR]). In particular there are uncountably many such spaces [BRS] and even infinitely many which embed into ℓp (ℓ2) [S1]. Thus it seems that a deeper study of the index in [BRS] will be needed for further progress. However some things, which we now recall, are known. Theorem 8.2. [P] If Y is complemented in ℓp then Y is isomorphic to ℓp. Theorem 8.3. [JZ] If Y is a Lp subspace of ℓp then Y is isomorphic to ℓp. Theorem 8.4. [EW] If Y is complemented in ℓp ⊕ ℓ2 then Y is isomorphic to ℓp, ℓ2 or ℓp ⊕ ℓ2. Theorem 8.5. [O] If Y is complemented in ℓp(ℓ2) then Y is isomorphic to ℓp, ℓ2, ℓp ⊕ ℓ2 or ℓp (ℓ2). SMALL SUBSPACES OF Lp 31 We recall that Xp is the Lp discovered by H. Rosenthal [R]. For p > 2, Xp may be defined to be the subspace of ℓp ⊕ ℓ2 spanned by (ei + wifi), where (ei) and (fi) are the unit vector bases of ℓp and ℓ2, respectively, and where wi → 0 with P w2p/p−2 = ∞. Since ℓp ⊕ ℓ2 embeds into Xp, the subspaces of Xp and of ℓp ⊕ ℓ2 are (up to isomorphism) the same. For 1 < p < 2 the space Xp is defined to be the dual of Xp′ where 1/p + 1/p′ = 1. When restricted to Lp-spaces, the results of this paper lead to a dichotomy valid for 1 < p < ∞. Proposition 8.6. Let Y be a Lp-space (1 < p < ∞). Either Y is isomorphic to a comple- mented subspace of Xp or Y has a complemented subspace isomorphic to ℓp(ℓ2). Proof. For p > 2 it is shown in [JO2] that a Lp-space which embeds in ℓp ⊕ ℓ2 embeds complementedly in Xp. Combining this with the main theorem of the present paper gives what we want for p > 2. When 1 < p < 2, the space Xp is defined to be the dual of Xp′ and so a simple duality argument extends the result to the full range 1 < p < ∞. (cid:3) i It remains a challenging problem to understand more deeply the structure of the Lp- subspaces of Xp and ℓp ⊕ ℓ2. Theorem 8.7. [JO2] If Y is a Lp subspace of ℓp ⊕ ℓ2 (or Xp), 2 < p < ∞, and Y has an unconditional basis then Y is isomorphic to ℓp, ℓp ⊕ ℓ2 or Xp. It is known [JRZ] that every Lp space has a basis but it remains open if it has an unconditional basis. Theorem 8.8. [JO2] If Y is a Lp subspace of ℓp ⊕ ℓ2 (1 < p < 2) with an unconditional basis then Y is isomorphic to ℓp or ℓp ⊕ ℓ2. So the main open problem for small Lp spaces is to overcome the unconditional basis requirement of 8.7 and 8.8. Problem 8.9. (a) Let X be a Lp subspace of ℓp ⊕ ℓ2 (2 < p < ∞). Is X isomorphic to ℓp, ℓp ⊕ ℓ2 or Xp? (b) Let X be a Lp subspace of ℓp ⊕ ℓ2 (1 < p < 2). Is X isomorphic to ℓp or ℓp ⊕ ℓ2? References [Ald] [Als] [AO] [BRS] D. Aldous, Subspaces of L1, via random measures, Trans. Amer. Math. Soc. 267 (2) (1981), 445 -- 463. D. Alspach, Tensor products and independent sums of Lp-spaces, 1 < p < ∞, Mem. Amer. Math. Soc. 660 (1999). D. Alspach and E. Odell, Lp spaces, Handbook of Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds., Elsevier, Amsterdam (2001), 123 -- 159. J. Bourgain, H.P. Rosenthal, and G. Schechtman, An ordinal Lp-index for Banach spaces, with application to complemented subspaces of Lp, Ann. of Math. 114 (2) (1981), 193 -- 228. D.L. Burkholder, Distribution function inequalities for martingales, Ann. Prob. 1 (1973), 19 -- 42. [B] [BDG] D.L. Burkholder, B.J. Davis, and R.F. Gundy, Integral inequalities for convex functions of operators on martingales, Proc. of the 6-th Berkeley Symp. on Math. Stat. and Probl. Vol. 2 (1972), 223 -- 240. I.S. Edelstein and P. Wojtaszczyk, On projections and unconditional bases in direct sums of Banach spaces, Studia Math. 56 (1976), no. 3, 263 -- 276. D.J.H. Garling, Stable Banach spaces, random measures and Orlicz spaces, Probability Measures on Groups, Lecture Notes in Mathematics 928, H. Heyer ed., Springer-Verlag 1982. [EW] [G] [GLR] Y. Gordon, D.R. Lewis and J.R. Retherford, Banach ideals of operators with applications. J. Func- [G-D] [H] tional Analysis 14 (1973), 85 -- 129. S. Guerre-Delabri`re, Classical Sequences in Banach Spaces, Marcel Dekker, 1992. P. Hitczenko, Best possible constants in martingale version of Rosenthal's inequality, Annals of Prob. 18 (1990), 1656 -- 1668. 32 [J] R. HAYDON, E. ODELL AND TH. SCHLUMPRECHT W.B. Johnson, On quotients of Lp which are quotients of ℓp, Compositio Math. 34 (1) (1977), 69 -- 89. [JMST] W.B. Johnson, B. Maurey, G. Schechtman and L. Tzafriri, Symmetric structures in Banach spaces, Mem. Amer. Math. Soc. 19 (217) (1979), v+298. [JO1] W.B. Johnson and E. Odell, Subspaces of Lp which embed into ℓp, Compositio Math. 28 (1974), 37 -- 49. [JO2] W.B. Johnson and E. Odell, Subspaces and quotients of ℓp ⊕ ℓ2 and Xp, Acta Math. 147 (1 -- 2) [JL] (1981), 117 -- 147. W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds., Elsevier, Amsterdam (2001), 123 -- 159. [JRZ] W.B. Johnson, H.P. Rosenthal, and M. Zippin, On bases, finite dimensional decompositions and [JZ] weaker structures in Banach spaces, Israel J. Math. 9 (1971), 488 -- 506. W.B. Johnson and M. Zippin, On subspaces of quotients of (P Gn)ℓp and (P Gn)c0 , Israel J. Math. 13 (1972), 311 -- 316. [KP] M.I. Kadets and A. Pe lczy´nski, Bases lacunary sequences and complemented subspaces in the spaces Lp, Studia Math. 21 (1961/1962), 161 -- 176. [KW] 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. [KOS] H. Knaust, E. Odell, and Th. Schlumprecht, On asymptotic structure, the Szlenk index and UKK [KK] [KM] [LP] properties in Banach spaces, Positivity 3 (1999), 173 -- 199. A. Koldobsky and H. Konig, Aspects of the isometric theory of Banach spaces, Handbook of Ge- ometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds., Elsevier, Amsterdam (2001), 899 -- 939. J.L. Krivine and B. Maurey, Espaces de Banach stables, Israel J. Math. 39 (4) (1981), 273 -- 295. J. Lindenstrauss and A. Pe lczy´nski, Absolutely summing operators in Lp spaces and their applica- tions, Studia Math. 29 (1968), 275 -- 321. J. Lindenstrauss and H.P. Rosenthal, The Lp spaces, Israel J. Math. 7 (1969), 325 -- 349. D.A. Martin, Borel determinacy, Annals of Math. 102 (1975), 363 -- 371. [LR] [Ma] [MMT] B. Maurey, V.D. Milman, and N. Tomzczak-Jaegermann, Asymptotic infinite dimensional theory of [O] [OS1] [OS2] Banach spaces, Oper. Theory: Adv. Appl. 77 (1994), 149 -- 175. E. Odell, On complemented subspaces of (P ℓ2)ℓp, Israel J. Math. 23 (3 -- 4) (1976), 353 -- 367. E. Odell and Th. Schlumprecht, Trees and branches in Banach spaces, Trans. Amer. Math. Soc. 354, no.10 (2002), 4085 -- 4108. E. Odell and Th. Schlumprecht, Embedding into Banach spaces with finite dimensional decomposi- tions, Rev. R. Acad. Cien. Serie A. Math., vol. 100 (1 -- 2) (2006), 295 -- 323. [OSZ] E. Odell, Th. Schlumprecht, and M. Zsak, On the structure of asymptotic ℓp spaces, On the structure [P] [PR] [Ro] [R] [S1] [S2] [Z] of asymptotic lp spaces. Q. J. Math. 59 (2008), no. 1, 85 -- 122. A. Pe lczy´nski, Projections in certain Banach spaces, Studia Math. 19 (1960), 209 -- 228. A. Pe lczy´nski and H. P. Rosenthal, Localization techniques in Lp spaces, Studia Mathematica 52 (1975), 263 -- 289. C. Rosendal, Infinite asymptotic games, Ann. Inst. Fourier (Grenoble) 59 (2009), no. 4, 1359 -- 1384. H.P. Rosenthal, On the subspaces of Lp (p > 2) spanned by sequences of independent random variables, Israel J. Math. 8 (1970), 273 -- 303. G. Schechtman, Examples of Lp spaces (1 < p 6= 2 < ∞), Israel J. Math. 22 (1975), 138 -- 147. G. Schechtman, personal communication. M. Zippin, Banach spaces with separable duals, Trans. Amer. Math. Soc. 310, Nr. 1 (1988), 371 -- 379. Brasenose College, Oxford OX1 4AJ, U.K. E-mail address: [email protected] Department of Mathematics, The University of Texas at Austin, Austin, TX 78712-0257 E-mail address: [email protected] Department of Mathematics, Texas A&M University, College Station, TX 77843-3368 E-mail address: [email protected]
1212.1996
2
1212
2013-09-20T16:49:41
The numerical range and the spectrum of a product of two orthogonal projections
[ "math.FA" ]
The aim of this paper is to describe the closure of the numerical range of the product of two orthogonal projections in Hilbert space as a closed convex hull of some explicit ellipses parametrized by points in the spectrum. Several improvements (removing the closure of the numerical range of the operator, using a parametrization after its eigenvalues) are possible under additional assumptions. An estimate of the least angular opening of a sector with vertex 1 containing the numerical range of a product of two orthogonal projections onto two subspaces is given in terms of the cosine of the Friedrichs angle. Applications to the rate of convergence in the method of alternating projections and to the uncertainty principle in harmonic analysis are also discussed.
math.FA
math
The numerical range and the spectrum of a product of two orthogonal projections Hubert Klaja ∗ Abstract The aim of this paper is to describe the closure of the numerical range of the product of two orthogonal projections in Hilbert space as a closed convex hull of some explicit ellipses parametrized by points in the spectrum. Several improvements (removing the closure of the numerical range of the operator, using a parametrization after its eigenvalues) are possible under additional assumptions. An estimate of the least angular opening of a sector with vertex 1 containing the numerical range of a product of two orthogonal projections onto two subspaces is given in terms of the cosine of the Friedrichs angle. Applications to the rate of convergence in the method of alternating projections and to the uncertainty principle in harmonic analysis are also discussed. Keywords: Numerical range; orthogonal projections; Friedrich angle; method of alternating projections; uncertainty principle; annihilating pair. MSC 2010: 47A12, 47A10. 1 Introduction Background. The numerical range of a Hilbert space operator T ∈ B(H) is defined as W (T ) = {hT x, xi , x ∈ H,kxk = 1}. It is always a convex set in the complex plane (the Toeplitz-Hausdorff theorem) containing in its closure the spectrum of the operator. Also, the intersection of the closure of the numerical ranges of all the operators similar to T is precisely the convex hull of the spectrum of T (Hildebrandt's theorem). We refer to the book [GR97] for these and other facts about numerical ranges. Another useful property the numerical ranges have is the following recent result of Crouzeix [Cro07]: for every T ∈ B(H) and every polynomial p , we have kp(T )k ≤ 12 supz∈W (T ) p(z). The problem. The main aim of this paper is to study the numerical range W (T ) and the numerical radius, defined by ω(T ) = sup{z , z ∈ W (T )}, of a product of two orthogonal projections T = PM2 PM1 . In what follows we denote by PM the orthogonal projection onto the closed subspace M of a given Hilbert space H. We prove a representation of the closure of W (T ) as a closed convex hull of some explicit ellipses parametrized by points in the spectrum σ(T ) of T and we discuss several applications. We also study the relationship between the numerical range (numerical radius) of a product of two orthogonal projections and its spectrum (resp. spectral radius). Recall that the spectral radius r(T ) of T ∈ B(H) is defined as r(T ) = sup{z , z ∈ σ(T )}. ∗Laboratoire Paul Painlevé, Université Lille 1, CNRS UMR 8524, Bât. M2, F-59655 Villeneuve d'Ascq, France ; [email protected] 1 Previous results. Orthogonal projections in Hilbert space are basic objects of study in Operator theory. Products or sums of orthogonal projections, in finite or infinite dimensional Hilbert spaces, appear in various problems and in many different areas, pure or applied. We refer the reader to a book [Gal04] and two recent surveys [Gal08, BS10] for more information. The fact that the numerical range of a finite product of orthogonal projections is included in some sector of the complex plane with vertex at 1 was an essential ingredient in the proof by Delyon and Delyon [DD99] of a conjecture of Burkholder, saying that the iterates of a product of conditional expectations are almost surely convergent to some conditional expectation in an L2 space (see also [Cro08, Coh07]). For a product of two orthogonal projections we know that the numerical range is included in a sector with vertex one and angle π/6 ([Cro08]). The spectrum of a product of two orthogonal projections appears naturally in the study of the rate of convergence in the strong operator topology of (PM2 PM1 )n to PM1∩M2 (cf. [Deu01, BDH09, DH10a, DH10b, BGM, BGM10, BL10]). This is a particular instance of von Neumann-Halperin type theorems, sometimes called in the literature the method of alternating projections. The following dichotomy holds (see [BDH09]): either the sequence (PM2 PM1 )n converge uniformly with an expo- nential speed to PM1∩M2 (if 1 /∈ σ(PM2 PM1 )), or the sequence of alternating pro- jections (PM2 PM1 )n converges arbitrarily slowly in the strong operator topology (if 1 ∈ σ(PM2 PM1 )). We refer to [BGM, BGM10] for several possible meanings of "slow convergence". An occurrence of the numerical range of operators related to sums of orthogonal projections appears also in some Harmonic analysis problems. The uncertainty prin- ciple in Fourier analysis is the informal assertion that a function f ∈ L2(R) and its Fourier transform F (f ) cannot be too small simultaneously. Annihilating pairs and strong annihilating pairs are a way to formulate this idea (precise definitions will be given in Section 5). Characterizations of annihilating pairs and strong annihilating pairs (S, Σ) in terms of the numerical range of the operator PS +iPΣ, constructed using some associated orthogonal projections PS and PΣ, can be found in [HJ94, Len72]. Main Results. Our first contribution is an exact formula for the closure of the numerical range W (PM2 PM1 ), expressed as a convex hull of some ellipses E (λ), parametrized by points in the spectrum (λ ∈ σ(PM2 PM1 )). Definition 1.1. Let λ ∈ [0, 1]. We denote E (λ) the domain delimited by the ellipse with foci 0 and λ, and minor axis length pλ(1 − λ). Theorem 1.2. Let M1 and M2 be two closed subspaces of H such that M1 6= H or M2 6= H. Then the closure of the numerical range of PM2 PM1 is the closure of the convex hull of the ellipses E (λ) for λ ∈ σ(PM2 PM1 ), i.e.: We refer to Remark 3.3 and to Figure 1 for more information about these ellipses. W (PM2 PM1 ) = conv{∪λ∈σ(PM2 PM1 )E (λ)}. The proof uses in an essential way Halmos' two subspaces theorem recalled in the next section. We will use a completely different approach to describe the numerical range (without the closure) of T = PM2 PM1 under the additional assumption that the self-adjoint operator T ∗T = PM1 PM2 PM1 is diagonalisable (see Definition 3.7). In this case the numerical range W (T ) is the convex hull of the same ellipses as before but this time parametrized by the point spectrum σp(T ) (=eigenvalues) of T = PM2 PM1 . Theorem 1.3. Let H be a separable Hilbert space. Let M1 and M2 be two closed subspaces of a Hilbert space H such that M1 6= H or M2 6= H. If PM1 PM2 PM1 is 2 diagonalizable, then the numerical range W (PM2 PM1 ) is the convex hull of the ellipses E (λ), with the λ's being the eigenvalues of PM2 PM1 , i.e.: W (PM2 PM1 ) = conv{∪λ∈σp(PM2 PM1 )E (λ)}. Concerning the relationship between the numerical radius and the spectral radius of a product of two orthogonal projections we prove the following result. Proposition 1.4. Let M1, M2 be two closed subspaces of H. The numerical radius and the spectral radius of PM2 PM1 are linked by the following formula: ω(PM2 PM1 ) = 1 2(cid:16)pr(PM2 PM1 ) + r(PM2 PM1 )(cid:17) . The proof is an application of Theorem 1.2 and the obtained formula is better than Kittaneh's inequality [Kit03] whenever the Friedrichs angle (Definition 2.7) between M1 and M2 is positive. Theorems 1.2 and 1.3 can be used to localize W (PM2 PM1 ) even if the spectrum of PM2 PM1 is unknown. We mention here the following important consequence about the inclusion of W (PM2 PM1 ) in a sector of vertex 1 whose angular opening is expressed in terms of the cosine of the Friedrichs angle cos(M1, M2) between the subspaces M1 and M2. This is a refinement of the Crouzeix's result [Cro08] for products of two orthogonal projections. Proposition 1.5. Let M1 and M2 be two closed subspaces of a Hilbert space H. We have the following inclusion: W (PM2 PM1 ) ⊂(z ∈ C,arg(1 − z) ≤ arctan(s cos2(M1, M2) 4 − cos2(M1, M2) )) . We next consider some inverse spectral problems and construct examples of pro- jections such that the spectrum of their product is a prescribed compact set included in [0, 1]. These examples will generalize to the infinite dimensional setting a result due to Nelson and Neumann [NN87]. We will also give examples that answer two open questions stated in a article of Nees [Nee99]. The following result allows to find σ(PM2 PM1 ) ∩ [ 1 4 , 1], the points of the spectrum which are larger than 1/4, whenever the closure W (PM2 PM1 ) of the numerical range is known. Theorem 1.6. Let α ∈ [ π 3 , π]. The following assertions are equivalent: 2(1−cos(α)) ∈ σ(PM2 PM1 ); 1. 2. sup{Re(z exp(−iα)), z ∈ W (PM2 PM1 )} = Actually it is possible to obtain a description of the entire spectrum σ(PM2 PM1 ) 4(1−cos(α)) . 1 1 starting from W (PM2 PM1 ) and W (PM2 (I − PM1 )). Finally, we will explain how the relation 1 ∈ W (PM2 PM1 ) is related to arbitrarily slow convergence in the von Neumann-Halperin theorem and we will give new charac- terizations of annihilating pairs and strong annihilating pairs in terms of W (PSPΣ). Organization of the paper. The rest of the paper is organized as follows. We recall in Section 2 several preliminary notions and known facts that will be useful in the sequel. In Section 3 we discuss the results concerning the exact computation of the numerical range of a product T of two orthogonal projections assuming that the spectrum, or the point spectrum, of T is known. Then we will give some "localiza- tion" results about the numerical range of T that require less informations about the 3 spectrum of T . Several examples are also given, some of them leading to an answer of two open questions from [Nee99]. In Section 4 we discuss the inverse problem of describing the spectrum of T knowing its numerical range, and the relationship be- tween the numerical and spectral radii of T . The paper ends with two applications of these results, one concerning the rate of convergence in the method of alternating projections and the second one concerning the uncertainty principle. 2 Preliminaries In this section we introduce some notations and recall several useful facts and results. Definition 2.1. Let E be a bounded subset of the complex plane C. We denote by conv{E} the convex hull of E, which is the set of all convex combinations of the points in E, i.e. conv{E} = {Xn∈N xnεn, εn ∈ E, xn ∈ [0, 1],Xn∈N xn = 1}. We refer the reader to [TUZ03] for a proof that this definition coincides with the classical one (the smallest convex subset which contains E). We will also denote by conv{E} the closure of the convex hull of E. 2.1 Halmos' two subspaces theorem For a fixed Hilbert space H and a closed subspace M of H we denote by M⊥ the orthogonal complement of M in H and by PM the orthogonal projection onto M . Let now M1 and M2 be two closed subspaces of a Hilbert space H. Consider the following orthogonal decomposition: H = (M1 ∩ M2) ⊕ (M1 ∩ M⊥2 ) ⊕ (M⊥1 ∩ M2) ⊕ (M⊥1 ∩ M⊥2 ) ⊕ H, (1) where H is the orthogonal complement of the first 4 subspaces. With respect to this orthogonal decomposition we can write: PM1 = I ⊕ I ⊕ 0 ⊕ 0 ⊕ P1 PM2 = I ⊕ 0 ⊕ I ⊕ 0 ⊕ P2 PM2 PM1 = I ⊕ 0 ⊕ 0 ⊕ 0 ⊕ P2 P1. 1 ∩ M (⊥) Suppose that the subspaces M (⊥) and H are not equal to {0}. Then us- ing the formula W (T ⊕ S) = conv{W (T ), W (S)} (see for instance [GR97]) we have W (PM2 PM1 ) = conv{{1}∪{0}∪W ( P2 P1)}. If M1∩M2 = {0} and the other subspaces are not equal to {0}, then we have that W (PM2 PM1 ) = conv{{0} ∪ W ( P2 P1)}. The other cases when the others subspaces are equal to {0} can be handle easily in the same way. 2 Definition 2.2. Let N1, N2 be two closed subspaces of an Hilbert space H. We say that (N1, N2) are in generic position if: N1 ∩ N2 = N⊥1 ∩ N2 = N1 ∩ N⊥2 = N⊥1 ∩ N⊥2 = {0}. In Sections 2 and 3 we will denote pairs of subspaces in generic position by (N1, N2), in order to distinguish them from pairs of general closed subspaces (M1, M2). We say that A is unitary equivalent to B (and write A ∼ B) if there exists a unitary operator U such that A = U BU∗. The following result, Halmos' two subspace theorem [Hal69], is a useful description of orthogonal projections of two subspaces in generic position. 4 Theorem 2.3. If (N1, N2) are in generic position, then there exists a subspace K of H such that H is unitary equivalent to K ⊕ K. Also, there exist two operators C, S ∈ B(K) such that 0 ≤ C ≤ I, 0 ≤ S ≤ I and C2 + S2 = I, and such that P1 and P2 are simultaneously unitary equivalent to the following operators: 0 P1 ∼(cid:18) I 0 0 (cid:19) , P2 ∼(cid:18) C2 CS S2 (cid:19) . CS Moreover, there exists a self adjoint operator T verifying 0 ≤ T ≤ π cos(T ) = C and sin(T ) = S. 2 I such that For a historical discussion and several applications of Halmos' two subspace theo- rem we refer the reader to [BS10]. 2.2 Support functions The notion of support functions is classical in convex analysis. Definition 2.4. Let S be a bounded convex set in C. Let α ∈ R. The support function of S , of angle α, is defined by the following formula: ρS (α) = sup{Re(zexp(−iα)), z ∈ S }. The following proposition shows that the support function characterizes the closure of convex sets. Proposition 2.5. We denote by S the closure of S . We have: S = {z ∈ C,∀α, Re(z exp(−iα)) ≤ ρS (α)}. We will need in this paper the following result about support functions. Lemma 2.6. Let S1, S2 be two bounded convex sets of the plane with support func- tions ρS1(α) and, respectively, ρS2(α). Let S be such that ρS (α) = maxi=1,2 ρSi(α). Then we have S = conv{S1, S2}. A proof of the above propositions and more information about support functions are available in [Roc70]. 2.3 Cosine of Friedrichs angle of two subspaces We now introduce the cosine of the Friederichs angle between two subspaces. We refer to [Deu01] as a source for more information. Definition 2.7. Let M1, M2 be two closed subspaces of H, with intersection M = M1 ∩ M2. We define the cosine of the Friederichs angle between M1 and M2 by the following formula: cos(M1, M2) = sup{hx, yi , x ∈ M1 ∩ M⊥, y ∈ M2 ∩ M⊥,kxk = kyk = 1}. An equivalent way ([KW88, Deu01]) to express the above cosine is given by the formula cos2(M1, M2) = kPM1 PM2 PM1 − PM1∩M2k. The following result, which will be helpful later on, offers a spectral interpretation of cos(M1, M2). Lemma 2.8. Let M1 and M2 be two closed subspaces of H. Then cos(M1, M2) = sup{ √λ : λ ∈ σ(PM2 PM1 ) \ {1}}. 5 This result can be seen as a consequence of Halmos' two subspace theorem (see [BS10]). We present here a different proof. Proof. We start by remarking that σ(PM2 PM1 ) is a compact subset of [0, 1]. Indeed, we have σ(PM1 PM2 PM1 ) \ {0} = σ((PM2 PM1 )PM1 ) \ {0} = σ(PM2 PM1 ) \ {0} and PM1 PM2 PM1 is a self-adjoint operator which is positive and of norm less or equal to one. Using the decomposition H = (M1 ∩ M2) ⊕ (M1 ∩ M2)⊥ we can write PM1 PM2 PM1 = PM1∩M2 ⊕ (PM1 PM2 PM1 − PM1∩M2), so we get σ(PM1 PM2 PM1 ) = σ(PM1∩M2 ) ∪ σ(PM1 PM2 PM1 − PM1∩M2). Since cos2(M1, M2) = kPM1 PM2 PM1 − PM1∩M2k = sup σ(PM1 PM2 PM1 − PM1∩M2), we obtain cos2(M1, M2) = sup σ(PM1 PM2 PM1 ) \ {1} = sup σ(PM2 PM1 ) \ {1}. 3 Description of the numerical range knowing the spectrum 3.1 The closure of the numerical range as a convex hull of ellipses The goal of this section is to prove Theorem 1.2 using a description of the support function of W (P2P1), which is a closed convex set of C. This idea appeared for instance in [Len72] in a different context. We will first assume that we are in generic position; the general case will be easily deduced from this particular one. The reader could see [RSN90] for more details about borelian functional calculus on self adjoint operators. Lemma 3.1. Suppose that (N1, N2) is in generic position. Denote Pi = PNi , i = 1, 2, the orthogonal projection on Ni. Then the support function of the numerical range of P2P1 is: ρW (P2P1)(α) = sup λ∈σ(P2P1) 1 2 (cos(α)λ +pλ(1 − sin(α)2λ)). Proof. We fix α ∈ [0, 2π]. We have that ρW (P2P1)(α) = sup{Re(hP2P1h, hi exp(−iα)), h ∈ H,khk = 1} = sup{Re(hexp(−iα)P2P1h, hi), h ∈ H,khk = 1} = sup{hRe(exp(−iα)P2P1)h, hi , h ∈ H,khk = 1}. Applying Halmos' two subspace theorem, there exists a self adjoint operator T such that P2P1 ∼(cid:18) So we have that cos(T )2 cos(T ) sin(T ) 0 (cid:19) , P1P2 ∼(cid:18) cos(T )2 0 0 cos(T ) sin(T ) 0 (cid:19) . Re(exp(−iα)P2P1) ∼ cos(α) cos(T )2 exp(−iα) 2 cos(T ) sin(T ) exp(iα) 2 cos(T ) sin(T ) 0 ! . 6 We set M (t, α) = cos(α) cos(t)2 exp(−iα) 2 cos(t) sin(t) exp(iα) 2 cos(t) sin(t) 0 ! . Then we have that Re(exp(−iα)P2P1) ∼ M (T, α). After some computations we get that M (t, α) = U∗(t, α)D(t, α)U (t, α) with ! , v2(t, α) (cid:19) , U (t, α) = D(t, α) =(cid:18) v1(t, α) exp(iα) cos(t) sin(t) exp(iα) cos(t) sin(t) 2v2(t,α) u2(t,α) 2v1(t,α) u1(t,α) u1(t,α) u2(t,α) 0 0 and v1(t, α) = 1 2 (cos(α) cos(t)2+cos(t)p1 − sin(α)2 cos(t)2) and v2(t, α) = 1 cos(t)p1 − sin(α)2 cos(t)2) and ui(t, α) = p4(vi(t, α))2 + cos(t)2 sin(t)2. One can easily check by passing to the limit when t goes to π 2 (cos(α) cos(t)2− U ( π 2 , α) = 1√2 √2 exp(iα) 2 that: −1√2 √2 ! . exp(iα) u1(T,α) We also have that U (t, α)U∗(t, α) = U∗(t, α)U (t, α) = I. As all entries of U (t, α) are borelians functions and T is a self adjoint operator, one can define 2v1(T,α) u1(T,α) , u2(T,α) , exp(iα) cos(T ) sin(T ) 2v2(T,α) . So we can define D(T, α) and U (T, α), and we have that M (T, α) = U∗(T, α)D(T, α)U (T, α) and U (T, α)U∗(T, α) = U∗(T, α)U (T, α) = I. So M (T, α) ∼ D(T, α) = v1(T, α) ⊕ v2(T, α). Note also that v1(t, α) ≥ 0 and v2(t, α) ≤ 0 for every t ∈ [0, π 2 ], we obtain the following order relations v2(T, α) ≤ 0 ≤ v1(T, α). Note that the operators vi(T, α) are self-adjoint. Therefore 2 ] and α ∈ [0, 2π] . As σ(T ) ⊂ [0, π and exp(iα) cos(T ) sin(T ) u2(T,α) ρW (P2P1)(α) = sup = sup khk=1hRe(exp(−iα)P2P1)h, hi kxk=1hv1(T, α)x, xi = kv1(T, α)k = sup v1(t0, α). Halmos' theorem implies that t0∈σ(T ) P1P2P1 ∼(cid:18) cos(T )2 0 (cid:19) . We have σ(P2P1)\{0} = σ((P2P1)P1)\{0} = σ(P1P2P1)\{0}, and cos2(σ(T ))∪{0} = σ(P1P2P1). Denoting λ = cos(t)2 and vi(λ, α) = 1 2 (cos(α)λ ±pλ(1 − sin(α)2λ)), we get ρW (P2P1)(α) = supλ∈σ(P2P1) v1(λ, α). Remark 3.2. Using a formula due to Lumer [Lum61, Lemma 12], we obtain 0 0 ρW (P2P1)(α) = sup Re(W (exp(−iα)P2P1)) = lim t→0+ = lim t→0+ kI − tRe(exp(−iα)P2P1)k − 1 t kI − t exp(−iα)P2P1k − 1 t . In order to make the formula of W (P2P1) more explicit, we will describe it as the convex hull of ellipses E (λ). Recall that for λ ∈ [0, 1], E (λ) denote the domain delimited by the ellipse with foci 0 and λ, and minor axis length pλ(1 − λ). Several of these ellipses are represented in Figure 1. 7 1.0 0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8 -1.0 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 Figure 1: Ellipse E (λ) for λ = 0.1, 0.2, . . . , 0.9 Remark 3.3. Other descriptions for E (λ) are possible. The Cartesian equation of the boundary of E (λ) is given by: (xλ − λ 2 )2 λ 4 + y2 λ λ(1−λ) 4 = 1, while the parametric equation of the boundary of E (λ) is given by: xλ(t) = √λ 2 cos(t) + λ 2 , yλ(t) = pλ(1 − λ) 2 sin(t). Lemma 3.4. Let λ ∈ [0, 1]. The support function of the ellipse E (λ) is: ρE (λ)(α) = 1 2 (cos(α)λ +pλ(1 − sin(α)2λ)). Proof. Let λ ∈ [0, 1]. The support function of E (λ) relative to the point 0 is given by ρE (λ)(α) = supt∈R xλ(t) cos(α)+yλ(t) sin(α), where xλ(t) and yλ(t) are the parametriza- tion of the boundary of E (λ). Let g = gλ,α be the function defined by the following formula: gλ,α(t) = xλ(t) cos(α) + yλ(t) sin(α) cos(α) cos(t) + pλ(1 − λ) 2 sin(α) sin(t). = λ 2 cos(α) + √λ 2 In order to compute ρE (λ)(α) we only need to study this function for α ∈ [0, π] because E (λ) has y = 0 as a symmetry axis. Suppose that cos(α) 6= 0. We have g′λ,α(t0) = 0 if and only if tan(t0) = √1 − λ tan(α). So the critical points of gλ,α are t0 = arctan(√1 − λ tan(α)) and t1 = arctan(√1 − λ tan(α))+ π. We denote ǫ0 = 1, ǫ1 = −1. Using standard trigonometric identities, we get cos(ti) = ǫi 1 p1 + (1 − λ) tan(α)2 , sin(ti) = ǫi √1 − λ tan(α) p1 + (1 − λ) tan(α)2 . 8 We denote ǫα = cos(α) cos(α) 2gλ,α(ti) =λ cos(α) + ǫi√λ cos(α) 1 . Using again some trigonometry formulas, we have: We finally obtain that p1 + (1 − λ) tan(α)2 √1 − λ tan(α) p1 + (1 − λ) tan(α)2 + ǫipλ(1 − λ) sin(α) =λ cos(α) + ǫiǫα√λp1 − λ sin(α)2. ρE (λ)(α) = √λ(1−λ) 2 1 2(cid:16)λ cos(α) + √λp1 − λ sin(α)2(cid:17) . √λ(1−λ) 2 . We obtain ρE (λ)(α) = 1 Suppose now that cos(α) = 0. Then gλ,α(t) = sin(t). So we get that in all situations ρE (λ)(α) = for every α. 2 (cos(α)λ +pλ(1 − sin(α)2λ)) Now we can easily prove Theorem 1.2 in the "generic position" case. Theorem 3.5. If (N1, N2) are in generic position, then: W (P2P1) = conv{∪λ∈σ(P2P1)E (λ)}. Proof. We first notice that: ρW (P2P1)(α) = sup λ∈σ(P2P1) 1 2 (cos(α)λ ±pλ(1 − sin(α)2λ)) = sup λ∈σ(P2P1) ρE (λ)(α). As the support function characterizes the closure of a convex bounded set, we simply use Lemma 2.6 to conclude. The proof of the general case follows now by combining the previous theorem with the decomposition (1). Proof of Theorem 1.2. Recall that M1 6= H or M2 6= H. We use the notation of the orthogonal decomposition (1) of H. Suppose that H = {0}. Then PM2 PM1 is the direct sum of 0 and I (or is zero if M1 ∩ M2 = {0}). Then it is easy to see that E (0) = {0} and E (1) = [0, 1]. So we have W (PM2 PM1 ) = [0, 1] = conv{E (0) ∪ E (1)}. When M1 ∩ M2 = {0}, we have W (PM2 PM1 ) = {0} = E (0). Suppose H 6= {0}, and M⊥1 ∩M⊥2 6= {0} (the cases M⊥1 ∩M2 6= {0} and M1∩M⊥2 6= {0} are similar). On the space M⊥1 ∩ M⊥2 , we have PM2 PM1 = 0. The numerical range of PM2 PM1 on (M⊥1 ∩ M⊥2 ) ⊕ H is conv{{0} ∪ conv{∪λ∈σ(P2P1)E (λ)}}. As E (0) = {0} ⊂ E (λ) for all λ ∈ [0, 1], the numerical range of PM2 PM1 on (M⊥1 ∩M⊥2 )⊕ H is given by conv{∪λ∈σ(P2P1)E (λ)}. Suppose M1 ∩ M2 6= {0}. As PM2 PM1 = I on the intersection M1 ∩ M2, the numerical range of PM2 PM1 on (M1 ∩ M2)⊕ H is conv{{1}∪ conv{∪λ∈σ(P2P1)E (λ)}}. For every λ ∈ [0, 1] we have 0 ∈ E (λ). As H 6= {0}, the numerical range of PM2 PM1 on (M1 ∩ M2) ⊕ H is conv{[0, 1]∪ conv{∪λ∈σ(P2P1)E (λ)}}. But E (1) = [0, 1]. So, finally, the numerical range of PM2 PM1 on (M1 ∩ M2)⊕ H is conv{∪λ∈σ(PM2 PM1 )E (λ)}. This proves the theorem. In the case when PM1 = I and PM2 = I, we have of course that W (PM2 PM1 ) = {1}. 9 Remark 3.6. In [CM11], Corach and Maestripieri proved that the Moore-Penrose pseudoinverse of a product of two orthogonal projections is idempotent (possibly un- bounded). Conversely, the Moore-Penrose pseudoinverse of an idempotent is a product of two orthogonal projections. It is well known that the numerical range of a (bounded) idempotent is an ellipse (see [SS10]). By using Halmos' theorem in a similar way as before, it is possible to prove that the closure of the numerical range of an idempotent E is the convex hull of the domains delimited by the ellipses E +(λ) of foci 0, 1 and of minor axis length q 1−λ λ , for λ describing the spectrum σ(E+) of the Moore-Penrose pseudoinverse E+ of E, i.e.: W (E) = conv{∪λ∈σ(E+)E +(λ)}. As E +(λ1) ⊂ E +(λ2), if λ1 ≤ λ2, the convex hull of all these ellipses will be just the biggest one, and we find another proof that W (E) is an ellipse. 3.2 W (P2P1) when P1P2P1 is diagonalizable Let (N1, N2) be a pair of closed subspaces of H. Denote Pi = PNi . Suppose that (N1, N2) is in generic position. As we have seen in the proof of Theorem 1.2, if we get W (P2P1) when (N1, N2) is in generic position, we can manage to get W (P2P1) in the general case. In this section we always assume for simplification that H is separable and make the hypothesis that P1P2P1 is diagonalizable, according to the following definition. Definition 3.7. We say that P1P2P1 is diagonalizable if there exists an orthonormal basis (cid:0)hn(cid:1)n∈N of H and a sequence of scalars (cid:0)λn(cid:1)n∈N such that: (x ∈ H). P1P2P1x = Xn∈N λn(cid:10)x, hn(cid:11) hn This happens for instance when P2P1 is a compact operator. Using our diagonal- izability assumption, it will be possible to decompose P2P1 as a direct sum of 2 × 2 matrices. As we know that the numerical range of such a matrix is an ellipse, this will permit to deduce the numerical range of P2P1. We first notice that 0 ≤ P1P2P1 ≤ I. Therefore 0 ≤ λn ≤ 1. The next lemma characterizes when hn ∈ N1. Lemma 3.8. Suppose that (N1, N2) is in generic position. We have: 1. hn ∈ N1 ⇔ λn 6= 0 2. hn ∈ N⊥1 ⇔ λn = 0. Proof. We know that P1P2P1hn = λnhn. If λn 6= 0, then hn = 1 P1P2P1hn ∈ N1. If λn = 0, then P1P2P1hn = 0. So P2P1hn ∈ N⊥1 ∩ N2 = {0}, because we are in generic position. So P2P1hn = 0. We get P1hn ∈ N⊥2 ∩ N1 = {0}, P1hn = 0 and thus hn ∈ N⊥1 . λn From now on, we just need those vectors hn which are in N1. For simplification, we denote these vectors as (hn)n∈N, each one correspond to a nonzero λn. This means that P1P2P1hn = λnhn. As we have hn ∈ N1, we get P1hn = hn. We denote (see Figure 2) wn = P2hn, wn = wn kwnk , fn = (I − P1)P2hn, fn = . fn kfnk Lemma 3.9. We have hwn, wki = δn,kλn and hwn, hki = δn,kλn, where δn,k is the Kronecker symbol, whose value is 1 if n = k, and 0 otherwise. 10 fn ✻ fn ✻ wn ✒ wn ✒ ❅ θn ❅ ✲ ❅ hn Figure 2: Proof. For the first equality, we have that hwn, wki = hP2P1hn, P2P1hki = hP1P2P1hn, hki = λn hhn, hki = δn,kλn. For the other one, we have hwn, hki = hP2P1hn, P1hki = hP1P2P1hn, hki = λn hhn, hki = δn,kλn. Corollary 3.10. Let span{h, w} be the closed subspace of H generated by h and w. If n 6= k, then span{hn, wn} is orthogonal to span{hk, wk}. Proposition 3.11. The range of span{hn, wn} by P2P1 verifies P2P1(span{hn, wn}) = span{wn} ⊂ span{hn, wn}. Proof. We just need to prove that P2P1(hn) and P2P1(wn) are collinear with wn. We have P2P1(hn) = wn. As hn is an eigenvector of P1P2P1, we obtain P2P1(wn) = P2P1P2P1(hn) = P2(λnhn) = λnwn. Lemma 3.12. We have span{hn, wn} = span{hn, fn}. Proof. As both of them are subspaces of dimension 2, it will be enough to show that span{hn, wn} ⊂ span{hn, fn}. As hn ∈ span{hn, fn}, we just need to prove that wn ∈ span{hn, fn}. We have wn = P2P1hn = P1P2P1hn + (I − P1)P2P1hn = λnhn + fn. So wn ∈ span{hn, fn}. Corollary 3.13. If n 6= k, then span{hn, fn} is orthogonal to span{hk, fk}. More- over, P2P1(span{hn, fn}) = span{wn} ⊂ span{hn, fn}. Proposition 3.14. We have P2(N1) = N2. Proof. The inclusion P2(N1) ⊂ N2 is obvious. In order to prove that P2(N1) ⊃ N2, it is enough to show that P2(N1)⊥ ⊂ N⊥2 . Let y ∈ P2(N1)⊥. Then, for every x ∈ N1, we have 0 = hy, P2(x)i = hP2(y), xi. So P2(y) ∈ N⊥1 . As P2(y) ∈ N2 and N⊥1 ∩ N2 = {0}, we obtain P2(y) = 0. So y ∈ N⊥2 . Corollary 3.15. The vectors ( wn)n∈N forms an orthonormal basis of N2. Proof. We know from Lemma 3.9 that ( wn)n∈N is an orthonormal system in N2. It remains to show that it is a generating system. We notice that the inclusion P2(N1) ⊂ span{wn, n ∈ N} implies, using P2(N1) = N2 and span{wn, n ∈ N} = span{ wn, n ∈ N} ⊂ N2, that N2 = P2(N1) ⊂ span{ wn, n ∈ N} ⊂ N2, and then N2 = span{ wn, n ∈ N}. Let us show that P2(N1) ⊂ span{wn, n ∈ N}. For x ∈ N1, there exists a sequence (νn) such that x = Pn νnhn. Therefore P2(x) = P2(Pn νnhn) =Pn νnP2(hn) =Pn νnwn. Finally P2(x) ∈ span{wn, n ∈ N}. 11 Similarly, we can also show the following proposition. Proposition 3.16. We have (I − P1)(N2) = N⊥1 . Moreover, ( fn)n∈N is an orthonor- mal basis of N⊥1 . Corollary 3.17. The operator P2P1 can be written as a direct sum of 2× 2 matrices, i.e.: P2P1 =Mn∈N P2P1 span{hn, fn} . Proof. As fn = fn , we have span{hn, fn} = span{hn, fn}, and P2P1(span{hn, fn}) ⊂ kfnk span{hn, fn}. Also, span{hn, fn} is orthogonal to span{hk, fk} whenever n 6= k. Moreover, ( fn)n∈N is an orthonormal basis of N⊥1 . We can write H as H = N1⊕N⊥1 = span{hn, n ∈ N} ⊕ span{ fn, n ∈ N} = ⊕nspan{hn, fn} which proves the result. Lemma 3.18. With respect to the orthonormal basis (hn, fn), the restriction of P2P1 to its invariant subspaces span{hn, fn} is given by: λn 0 P2P1 span{hn, fn} =(cid:18) pλn(1 − λn) 0 (cid:19) . Proof. As fn ∈ N⊥1 , we have P1 fn = 0, so P2P1 fn = 0. We can represent P2P1hn as: P2P1hn = P1P2P1hn + (I − P1)P2P1hn = λnhn + fn = λnhn + kfnk fn. In order to complete the proof, we have to show that kfnk = pλn(1 − λn). We have kfnk2 = k(I − P1)P2P1hnk2 = kP2P1hnk2 − kP1P2P1hnk2 = hP1P2P1hn, hni − kλnhnk2 = λn − λ2 n . Remark 3.19. As 0 ≤ P1P2P1 ≤ I, we have 0 ≤ λn ≤ 1 for every n. There exists θn such that 0 ≤ θn ≤ π 2 and cos(θn)2 = λn. Now we can rewrite P2P1 span{hn, fn} as: P2P1 span{hn, fn} =(cid:18) cos(θn)2 cos(θn) sin(θn) 0 (cid:19) . 0 This corresponds to the matrix of the composition of two orthogonal projections in the plane, projecting onto two lines of angle θn. Corollary 3.20. The numerical range W (P2P1 span{hn, fn} Proof. This is consequence of the classical ellipse lemma for the numerical range of a 2 × 2 matrix (see for instance [GR97]). ) is the ellipse E (λn). The following corollary is a "generic position" version of Theorem 1.3. Corollary 3.21. Let (N1, N2) be two subpsaces in generic position such that P1P2P1 is diagonalizable, then the numerical range W (P2P1) is the convex hull of the ellipses E (λ) for all the λ's which are non zero eigenvalues of P2P1, i.e.: W (P2P1) = conv{∪λ∈σp(P2P1)\{0}E (λ)}. have that hP2P1xn,xni Proof. From Corollary 3.17, we have that H = ⊕n∈Nspan{hn, fn}. Let x = ⊕n∈Nxn be a vector in H such that xn ∈ span{hn, fn} and kxk2 = Pn∈N kxnk2 = 1. Then hP2P1x, xi =Pn∈N hP2P1xn, xni =Pn∈N kxnk2 hP2P1xn,xni Let (αn)n∈N be a sequence such that αn ∈ [0, 1] and Pn∈N αn = 1. Let (ǫn)n∈N be a sequence such that ǫn ∈ E (λn). From Corollary 3.20, there exist some xn ∈ span{hn, fn} such that kxnk = 1 and ǫn = hP2P1xn, xni. Let x = Pn∈N αnxn, then hP2P1x, xi =Pn∈N αnǫn. So conv{∪λ∈σp(P2P1)\{0}E (λ)} ⊂ W (P2P1). ∈ E (λn). So W (P2P1) ⊂ conv{∪λ∈σp(P2P1)\{0}E (λ)}. . From Corollary 3.20, we kxnk2 kxnk2 12 Using the same idea as in the proof of Theorem 1.2, we can deduce Theorem 1.3 from Corollary 3.21. With this Corollary, we can see that the numerical range of a product of two orthogonal projections is not closed in general. Example 3.22. Let (N1, N2) be two subspaces in generic position and denote PNi = Pi. Suppose that P1P2P1 is diagonalizable. Moreover suppose that there exists an orthonormal basis (hn)n∈N∗ of N1 such that for all x ∈ H we have P1P2P1x = Xn∈N∗(cid:18)1 − 1 n + 1(cid:19)hx, hni hn. Then we have that σp(P2P1) = {1− 1 Therefore by Corollary 3.21 and Theorem 1.2, we have that W (P2P1) = conv{∪λ∈σp(P2P1)\{0}E (λ)} and W (P2P1) = conv{∪λ∈σ(PM2 PM1 )E (λ)}. n+1 , n ∈ N∗}∪{0} and σ(P2P1) = σp(P2P1)∪{1}. We have that 1 ∈ W (P2P1) but 1 /∈ W (P2P1). Note that 1 ∈ E (λ) if and only if λ = 1. We have that (see Remark 3.3) 1 n + 1 ) ⊂ W (P2P1). x1− 1 n+1 (0) = As limn→∞ x1− 1 n+1 1 1 1 2 n + 1 (r1 − (0) = 1 we have that 1 ∈ W (P2P1) . + 1 − n + 1 ) ∈ E (1 − Suppose that 1 ∈ W (P2P1). Then there exists x ∈ H such that kxk = 1 and hP2P1x, xi = 1. As 1 = hP2P1x, xi ≤ kP2P1xkkxk ≤ 1, we have that hP2P1x, xi = kP2P1xk kxk, so there exists λ such that P2P1x = λx. We get that 1 = hP2P1x, xi = λhx, xi = λ. So λ = 1 ∈ σp(P2P1). This is a contradiction with 1 /∈ σp(P2P1), so 1 /∈ W (P2P1). Example 3.23. There are non-trivial examples where P1P2P1 admits only 0 as eigen- value (hence P1P2P1 is not diagonalizable). Let T ∈ B(L2([0, 1])) be defined by T f (x) = xf (x). One can easily show that T is an injective positive contraction that has no eigenvalues, with Ker(I − T ) = {0} and σ(T ) = [0, 1]. If we set C = T 1/2 and S = (I − T )1/2, we easily see that C and S are injective and positive contractions with no eigenvalues such that C2 + S2 = I. Moreover C and S commute. We set H = L2([0, 1]) ⊕ L2([0, 1]) and P1 =(cid:18) I 0 (cid:19) , P2 =(cid:18) C2 CS S2 (cid:19) . CS 0 0 Then P1 and P2 are orthogonal projections onto subspaces sitting in generic position, and P1P2P1 =(cid:18) C2 0 0 0 (cid:19) . Suppose there exist f ⊕ g ∈ H and λ ∈ σ(P2P1) such that P1P2P1(f ⊕ g) = λ(f ⊕ g). Then xf (x) = λf (x) almost everywhere, and 0 = λg(x). This implies that λ = 0 and f = 0. So 0 is the only eigenvalue of P1P2P1. However, we have σ(P1P2P1) = σ(T ) ∪ {0} = [0, 1]. Remark 3.24. At the end of [Nee99], the author asks if kP2P1k2 is an accumulation point of eigenvalues, and if the spectrum P2P1 without zero consists only of eigenval- ues. The previous example answers these two questions negatively. 13 1.0 0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8 -1.0 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 Figure 3: convλ∈[0,1]{E (λ)} 3.3 Localization of W (P2P1) First we have this simple consequence of Theorem 1.2. Corollary 3.25. Let P1, P2 be two orthogonal projections. We have: W (P2P1) ⊂ convλ∈[0,1]{E (λ)}. Proof. If P1 = P2 = I this is clear since W (I) = {1}. Now suppose that P1 6= I or P2 6= I. We use Theorem 1.2 and the fact that σ(P2P1) ⊂ [0, 1], so we have the inclusion convλ∈σ(PM2 PM1 ){E (λ)} ⊂ convλ∈[0,1]{E (λ)}. This corollary says that if we can include convλ∈[0,1]{E (λ)} (see Figure 3) in a subset of C, then for any pair of projection P1, P2 we can include W (P2P1) in the same subset. The next lemma is an example of localization of the numerical range using Corollary 3.25. 8 , x = 1, y = 1 Lemma 3.26. Let P1 and P2 be two orthogonal projections. Then W (P2P1) is a subset of the rectangle whose sides are x = − 1 Proof. Using Corollary 3.25 and the parametric equation of the boundary of E (λ) (see Remark 3.3), we can prove that for all t ∈ R and for all λ ∈ [0, 1], we have − 1 8 ≤ xλ(t) ≤ 1 and − 1 Proof of Proposition 1.5. Suppose that we have found θλ such that E (λ) ⊂ {z ∈ C,arg(1 − z) ≤ θλ} for every λ. Taking θ = sup{θλ : λ ∈ σ(P2P1)}, we will have that 4 and y = − 1 4 . 4 ≤ yλ(t) ≤ 1 4 . W (P2P1) ⊂ convλ∈σ(P2P1){E (λ)} ⊂ {z ∈ C,arg(1 − z) ≤ θ}. First we note that E (0) = {0} and E (1) = [0, 1]. So we have θ0 = θ1 = 0. For λ ∈]0, 1[, we denote (xλ(t), yλ(t)) the parametrization of the boundary of E (λ) given in Remark 3.3. We denote θλ(t) the angle between the line connecting the points 0 and 1, and the one connecting points 1 and (xλ(t), yλ(t)). We have that θλ = supt∈R θλ(t), and tan(θλ(t)) = yλ(t) 1 − xλ(t) = pλ(1 − λ) sin(t) 2 − λ − √λ cos(t) . 14 By differentiating tan(θλ(t)), we can see that t0 is a critical point if cos(t0) = we have that √λ 2−λ . So tan(θλ) = pλ(1 − λ)q1 − λ √λ (2−λ)2 √λ (2−λ) 2 − λ − (2 − λ)2 − λ = pλ(1 − λ)p(2 − λ)2 − λ √λ√4−λ = √λ √4 − λ . As θ = supλ∈σ(P2P1) θλ, we get that tan(θ) = supλ∈σ(P2P1)\{1} clude using Lemma 2.8. . Then we con- Remark 3.27. We obtain as a consequence the result that the numerical range of a product of two orthogonal projections is included in a sector with vertex 1 and angle π/6 ([Cro08]). Also, the result of Proposition 1.5 is sharp, in the sense that if θ < 4−cos2(M1,M2) ), then W (P2P1) is not included in {z ∈ C,arg(1 − z) ≤ θ}. arctan(q cos2(M1,M2) 3.4 Some examples Let P1, P2 be two orthogonal projections. The spectrum σ(P2P1) is always a compact subset of [0, 1]. In this section, we study the following inverse spectral problem : let K be a compact subset of [0, 1]; when two orthogonal projections P1 and P2 exist such that σ(P2P1) = K? We will show that the answer is positive if and only if 0 ∈ K or K = {1}. We start with the case K = {1}. Proposition 3.28. Let M1 and M2 be two subspaces of H. If 0 does not belong to σ(PM2 PM1 ), then we have that M1 = M2 = H, PM1 = PM2 = I and σ(PM2 PM1 ) = {1}. Proof. We decompose H as in (1): H = (M1 ∩ M2) ⊕ (M1 ∩ M⊥2 ) ⊕ (M⊥1 ∩ M2) ⊕ (M⊥1 ∩ M⊥2 ) ⊕ H. Then PM2 PM1 = I ⊕ 0 ⊕ 0 ⊕ 0 ⊕ P2P1. As 0 does not belong to σ(P2P1), we obtain M1 ∩ M⊥2 = M⊥1 ∩ M2 = M⊥1 ∩ M⊥2 = H = {0} (otherwise PM2 PM1 would have a non trivial kernel). So we have H = M1 ∩ M2 and M1 = M2 = H. Therefore PM1 = PM2 = I and σ(PM2 PM1 ) = σ(I) = {1}. Now, we suppose that 0 ∈ K. Theorem 3.29. Let H be a separable Hilbert space. Let K be a compact subset of [0, 1] such that 0 ∈ K. Then there exist two orthogonal projections P1, P2 on H such that σ(P2P1) = K. Moreover, P1P2P1 is diagonalisable. Proof. As K is a compact subset of [0, 1], there exists a sequence (λn) in K such that {λn, n ∈ N} = K. For all n ∈ N, there exists a unique θn ∈ [0, π 2 ] such that λn = cos(θn)2. Let (en)n∈N be an orthonormal basis of H. We denote hn = e2n, fn = e2n+1 and wn = cos(θn)e2n + sin(θn)e2n+1. Let N1 = span{hn, n ∈ N} and N2 = span{ wn, n ∈ N} (see Figure 2). Then we have that P1hn = hn, P1 fn = 0 and P2hn = cos(θn)2hn + cos(θn) sin(θn) fn, P2 fn = cos(θn) sin(θn)hn + sin(θn)2 fn. Hence P2P1hn = cos(θn)2hn + cos(θn) sin(θn) fn and P2P1 fn = 0. Thus we get P2P1 =Mn∈N =Mn∈N(cid:18) P2P1 span{hn, fn} cos(θn) sin(θn) 0 (cid:19) . cos(θn)2 0 15 Also, σ(P2P1) = {cos(θn)2, n ∈ N} ∪ {0} = {λn, n ∈ N} ∪ {0} = K. Remark 3.30. We have proved in the previous section that W (P2P1) ⊂ conv{∪λ∈[0,1]E (λ)}. There are examples where this inclusion is an equality. According to Theorem 1.2, we just need two projections that satisfy σ(P2P1) = [0, 1]. The projections of Example 3.23 satisfy this condition, but P1P2P1 is not diagonalisable. With Theorem 3.29, we can also construct an example such that P1P2P1 is diagonalisable and σ(P2P1) = [0, 1]. Remark 3.31. As we now know all the possible shapes of σ(P2P1), Theorem 1.2 gives all the possible shapes of W (P2P1). Remark 3.32. Using the parametrization of the boundary of E (λ) (see Remark 3.3), we can prove that for all λ ∈ [0, 1 4}. It follows from Theorem 3.29 that there exist orthogonal projections P1, P2, Q1, Q2 such that σ(P2P1) = K1 and σ(Q2Q1) = K2. Moreover, we have that: 4 ] and K2 = {0, 1 4 ], E (λ) ⊂ E ( 1 4 ). Let K1 = [0, 1 W (P2P1) = conv{∪λ∈[0, 1 4 ] E (λ)} = E ( 1 4 ) = conv{E (0) ∪ E ( 1 4 )} = W (Q2Q1). This shows that the points of the spectrum of P2P1 which are less that 1 4 are not uniquely determined by the numerical range. We will see in the next section that the situation is different for spectral values greater than 1 4 . 4 The spectrum of P2P1 in terms of the numerical range 4.1 The relationship between the spectral and numerical radii In this section, we will prove proposition 1.4, and compare this result with an inequality from [Kit03]. Proof of Proposition 1.4. If M1 = M2 = H, this is true. Now we suppose that M1 6= H or M2 6= H. By combining the definition of the numerical radius with the Theorem 1.2, we obtain: ω(P2P1) = sup w∈W (P2P1)w = sup w∈E (λ),λ∈σ(P2P1)w . First, we compute supw∈E (λ) w for a fixed λ. We denote by (xλ(t), yλ(t)) the parametrization of the boundary of E (λ) given in Remark 3.3. We have supw∈E (λ) w = 4 (λ2 cos(t)2 + 2λ√λ cos(t) + λ). There- supt∈Rpxλ(t)2 + yλ(t)2 and xλ(t)2 + yλ(t)2 = 1 fore sup w∈E (λ)w =r 1 (λ + √λ). 1 2 4(cid:16)λ2 + 2λ√λ + λ(cid:17) = (λ + √λ) = 1 2 sup w∈E (λ),λ∈σ(P2P1)w = sup λ∈σ(P2P1) Finally, ω(P2P1) = 1 2 (r(P2P1) +pr(P2P1)). Remark 4.1. In [Kit03], Kittaneh proved that for any operator T , we have the following inequality: ω(T ) ≤ 1 2(cid:16)kTk +(cid:13)(cid:13)T 2(cid:13)(cid:13) 1 2(cid:17) . (2) 16 1 1 3 1 2 ) 2 ) = 1 ω(P2P1) < 1 2 (cos(M1, M2)+cos(M1, M2) Let us compare Proposition 1.4 with Kittaneh's inequality when T = P2P1. If M1 ∩ ω(P2P1) = 1. Thus ω(P2P1) = 1 and in this case, (2) is an equality. 2 (pr(P2P1) + r(P2P1)) = 1 2 (pr(P2P1) + r(P2P1)) = 1 2(cid:16)kP2P1k +(cid:13)(cid:13)(P2P1)2(cid:13)(cid:13) M2 6= {0}, then 1 is eigenvalue of P2P1. So kP2P1k =(cid:13)(cid:13)(P2P1)2(cid:13)(cid:13) = 1, r(P2P1) = 1 and 2 (kP2P1k +(cid:13)(cid:13)(P2P1)2(cid:13)(cid:13) If M1 ∩ M2 = {0}, then according to [KW88, Deu01] we have k(P2P1)nk = cos(M1, M2)2n−1 and kP1P2P1k = cos(M1, M2)2 = r(P1P2P1) = r(P2P1). So we have ω(P2P1) = 1 2 (cos(M1, M2) + cos(M1, M2)2) and also 1 2 (kP2P1k+(cid:13)(cid:13)(P2P1)2(cid:13)(cid:13) 2 ). If cos(M1, M2) < 1, then 2(cid:17). So in this case, (2) is a strict inequality. 4.2 How to find σ(P2P1) from W (P2P1) (and W (P2(I − P1))) Contrarily to Sections 2.1 and 2.2, where we have described W (P2P1) in terms of σ(P2P1), the aim of this section is to obtain information about the spectrum of P2P1 from its numerical range. We give an informal idea about how we do this. Denote gα(λ) = 1 2 (cos(α)λ+pλ(1 − sin(α)2λ)), then we have ρW (P2P1)(α) = supλ∈σ(P2P1) gα(λ). We will use the support function as a tool to identify if the ellipse E (λ) is in the nu- merical range. If this is the case, then λ will be in the spectrum. Denote by S the closure of convλ∈[0,1]{E (λ)}. By Corollary 3.25, we have W (P2P1) ⊂ S , so supλ∈σ(P2P1) gα(λ) = ρW (P2P1)(α) ≤ ρS (α) = supλ∈[0,1] gα(λ). Using the continu- ity of the function gα(·) and the compacity of σ(P2P1), we get the existence of a point λ0 ∈ σ(P2P1) such that ρW (P2P1)(α) = gα(λ0). With this information we are able to find an explicit formula for ρS (α). Moreover, we will see that the equality ρW (P2P1)(α) = ρS (α) is equivalent to the presence of a unique point λ0 (depending only on α) in the spectrum of P2P1. We begin by giving a necessary and sufficient condition such that λ is a critical point of gα(λ). Lemma 4.2. Let λ0 ∈]0, 1[ and α ∈]0, π[. Then λ0 is a critical point for gα if and only if we have: α = 2 arcsin(p1 − λ0 sin(α)2). 2 (cos(α) + 1−2λ sin(α)2 2√λ(1−sin(α)2λ) 1 1−2 sin(γ)2 sin(α)2 cos(α) = 1 ). Thus g′α(λ) = 0 if and only if Proof. We have g′α(λ) = 1 √λ cos(α) = 2√1−sin(α)2λ −p1 − λ sin(α)2. Denoting X = p1 − sin(α)2λ, we have that g′α(λ) = 0 if and only ifq 1−X 2 2X − X, or, equivalently, if and only if cot(α) = 1−2X 2 2 sin(γ)√1−sin(γ)2 = cot(2γ). As λ ∈ [0, 1], we have that X ∈ [cos(α) , 1], and γ ∈ [arcsin(cos(α)), π 2 ] ⊂ [0, π 2 ]. So 2γ ∈ [0, π]. Therefore g′α(λ) = 0 if and only if α = 2γ, if and only if α = 2 arcsin(p1 − λ0 sin(α)2). 2X√1−X 2 . We denote X = sin(γ) and get that cot(α) = The next corollary says that the support functions of W (P2P1) for α ∈]0, π 3 ], and λ0 ∈]0, 1[, then λ0 is not a critical point of gα. not give us useful information about σ(P2P1). Corollary 4.3. If α ∈ [0, π Proof. We just need to check that Lemma 4.2 fails in this case. If λ0 ∈]0, 1[, then 2 arcsin(p1 − λ0 sin(α)2) ∈] arcsin(cos(α)), π[. If α satisfies the condition of Lemma 4.2, then α ∈] arcsin(cos(α)), π[. We want to know when we have α = 2 arcsin(cos(α)). If α = 2 arcsin(cos(α)), using some trigonometric formulas, we get that sin(α) = 2 cos(α) sin(α). So cos(α) = 1 If 2 . 3 , then 2 arcsin(cos(α)) = π If α = π 3 = α. 3 ] do 17 α = 2π only if α = π no critical point on ]0, 1[. 3 , then 2 arcsin(cos(α)) = π 3 . Moreover, if α ∈ [0, π 3 6= α. In other words, α = 2 arcsin(cos(α)) if and 3 ], then we have α < 2 arcsin(cos(α)), so gα has The following proposition says that ρW (P2P1)(α) can give information on σ(P2P1) 3 , π]. if α ∈ [ π Proposition 4.4. If α ∈ [ π Proof. From Lemma 4.2, we know that λ is a critical point of gα if and only if α = 3 , π], then the only critical point of gα is λα = 1+cos(α) 2 sin(α)2 . 2 arcsin(p1 − λ sin(α)2). Compose with sinus on each side of the equality and use some trigonometric formulas to get that sin(α) = 2p1 − λ sin(α)2√λ sin(α). Dividing each side by sin(α) and raising to the square, we get that 4λ2 sin(α)2 − 4λ + 1 = 0. Therefore if λ is a critical point of gα, then λ = 1+cos(α) 2 sin(α)2 . If λ = 1−cos(α) 2 sin(α)2 , then 2 sin(α)2 or λ = 1−cos(α) 2 arcsin(p1 − λ sin(α)2) = 2 arcsin(r 1 2 = 2 arcsin(cos( (1 + cos(α))) α 2 )) Lemma 4.2 says that λ is not a critical point of gα. If λ = 1+cos(α) 2 sin(α)2 , then 6= α. 2 arcsin(p1 − λ sin(α)2) = 2 arcsin(r 1 2 = 2 arcsin(sin( (1 − cos(α))) α )) 2 = α. According to Lemma 4.2, λ is a critical point of gα. Remark 4.5. The condition α ∈ [ π If we have π 3 ≤ α ≤ π, then 1 4 ≤ We give now an explicit formula for ρS (α). λα = 3 , π] ensures that λα ∈ [0, 1]. We remark that 1 + cos(α) 2 sin(α)2 = 2(1 − cos(α)) 2(1−cos(α)) ≤ 1. So λα ∈ [ 1 4 , 1]. 1 1 . Corollary 4.6. The support function of S = conv{∪λ∈[0,1]E (λ)} is given by the following formula: ρS (α) =(cid:26) cos(α) 4(1−cos(α)) 1 if α ∈ [0, π 3 ] if α ∈ [ π 3 , π] . 1 3 ], then ρS (α) = max{gα(0), gα(1)} and if α ∈ [ π Proof. We know that ρS (α) = maxλ∈[0,1] gα(λ). We proved previously that if α ∈ [0, π 2(1−cos(α)) . We have that gα(0) = 0 and gα(1) = cos(α) and also with λα = 4(1−cos(α)) . Now it remains to show that for any α ∈ [ π gα(λα) = 3 , π], we have 4(1−cos(α)) , and the gα(λα) ≥ gα(1). As 4(1−cos(α)) last term is always positive, we get the announced result. 4(1−cos(α)) − cos(α) = 1−4 cos(α)+4 cos(α)2 = (1−2 cos(α))2 1 1 3 , π] then ρS (α) = max{gα(0), gα(λα), gα(1)}, Now we have enough material to prove Theorem 1.6. 18 Proof of Theorem 1.6. Let α ∈ [ π 3 , π]. We know that ρW (P2P1)(α) = supλ∈σ(P2P1) gα(λ). As σ(P2P1) is a compact set and gα is a continuous function, there exists a λ0 ∈ σ(P2P1) such that: ρW (P2P1)(α) = maxλ∈σ(P2P1) gα(λ) = gα(λ0). According to Propo- sition 4.4, we have gα(λ0) = 4(1−cos(α)) if and only if λ0 = λα = 2(1−cos(α)) . 1 1 If ρW (P2P1)(α) = 1 4(1−cos(α)) = gα(λ0), then we have λ0 = λα = "1 ⇒ 2": 1 2(1−cos(α)) . As λ0 ∈ σ(P2P1), we get that λα ∈ σ(P2P1). "2 ⇒ 1": If λα ∈ σ(P2P1), then we have that: gα(λ) ≤ max λ∈[0,1] gα(λα) ≤ max λ∈σ(P2P1) gα(λ) = gα(λα). Therefore ρW (P2P1)(α) = max λ∈σ(P2P1) gα(λ) = gα(λα) = 1 4(1 − cos(α)) . Given α, Theorem 1.6 tells us whether λα is in the spectrum or not by looking at the support function of W (P2P1) in the direction α. Given λ, the next corollary tell us in which direction αλ we have to look to know whether λ is in σ(P2P1) or not. Corollary 4.7. Let λ ∈ [ 1 assertions are equivalent: 4 , 1]. We denote αλ = arccos(1 − 1 2λ ). The following 1. ρW (P2P1)(αλ) = 2. λ ∈ σ(P2P1). 1 4(1−cos(αλ)) ; 3 , π] −→ [ 1 Proof. We denote f : [ π 2(1−cos(α)) . The equivalence follows from Theorem 1.6, and the facts that f is bijective with inverse function given by λ 7→ arccos(1 − 1 4 , 1] the function given by f (α) = 2λ ). 1 The next proposition is a "trick" to deduce most of the spectrum of P2P1 from σ(P2(I − P1)). As P2(I − P1) is again a product of two orthogonal projections, all the results of this paper apply also to this operator. Proposition 4.8. Let λ 6= 0. If λ ∈ σ(P2(I − P1)), then 1 − λ ∈ σ(P2P1). Proof. We decompose H as in (1). Therefore we have H = (M1 ∩ M2) ⊕ (M1 ∩ M⊥2 ) ⊕ (M⊥1 ∩ M2) ⊕ (M⊥1 ∩ M⊥2 ) ⊕ H and 0 0 PM1 ∼ I ⊕ I ⊕ 0 ⊕ 0 ⊕(cid:18) I 0 (cid:19) PM2 ∼ I ⊕ 0 ⊕ I ⊕ 0 ⊕(cid:18) C2 CS S2 (cid:19) 0 I (cid:19) CS 0 (cid:19) PM2 (I − PM1 ) ∼ 0 ⊕ 0 ⊕ I ⊕ 0 ⊕(cid:18) 0 CS 0 S2 (cid:19) I − PM1 ∼ 0 ⊕ 0 ⊕ I ⊕ I ⊕(cid:18) 0 PM2 PM1 ∼ I ⊕ 0 ⊕ 0 ⊕ 0 ⊕(cid:18) C2 CS 0 0 19 2 1 ∩ M (⊥) We remind that C2 + S2 = I, so we have that σ(S2) = 1 − σ(C2). Suppose that the subspaces M (⊥) and H are not equal to {0}. Then σ(PM2 PM1 ) = ∪{{1},{0}, σ(C2) ∪ {0}}. If M1 ∩ M2 = {0}, then we have to remove {1} of the former union to get σ(PM2 PM1 ). If M1 ∩ M⊥2 = M⊥1 ∩ M2 = M⊥1 ∩ M⊥2 = {0}, then we have to remove {0} of the former union to get σ(PM2 PM1 ). If H = {0}, then we have to remove σ(C2) ∪ {0} of the former union to get σ(PM2 PM1 ). In a similar way, σ(PM2 (I − PM1 )) = ∪{{0},{1}, (1− σ(C2)) ∪ {0}} depending on whether the corresponding subspaces are not reduced to {0}. Let λ 6= 0 be such that λ ∈ σ(PM2 (I − PM1)). Suppose that λ = 1 and M⊥1 ∩ M2 6= {0}. Then we get that PM2 PM1 = 0 on M⊥1 ∩ M2. So 1 − λ = 0 is an eigenvalue of PM2 PM1 . In the other cases, we get that λ ∈ 1− σ(C2) and H 6= {0}, hence 1− λ ∈ σ(C2) ⊂ σ(PM2 PM1 ). Example 4.9. There exist orthogonal projections such that 1 − σ(PM2 (I − PM1 )) 6= σ(PM2 PM1 ). We will exhibit an example in H = C3. Let (e1, e2, e3) be an orthonormal basis of C3. We set M1 = span{e1} and M2 = span{e2}. Then we get that M1 ∩ M2 = {0}, M1 ∩ M⊥2 = span{e2}, M⊥1 ∩ M2 = span{e1}, M⊥1 ∩ M⊥2 = span{e3} and H = {0}. So PM2 PM1 = 0, PM2 (I − PM1 ) = PM2 , σ(PM2 PM1 ) = {0} and σ(PM2 (I − PM1 )) = {0, 1}. Remark 4.10. Theorem 1.6 allows us to deduce σ(P2P1) ∩ [ 1 I − P1 is also an orthogonal projection, we can also deduce σ(P2(I − P1))∩ [ 1 W (P2(I − P1)). Moreover, Proposition 4.8 allows us to deduce σ(P2P1) ∩ [0, 3 σ(P2(I − P1)) ∩ [ 1 In other words, we can deduce σ(P2P1) from W (P2P1) and W (P2(I − P1)). 4 , 1] from W (P2P1). As 4 , 1] from 4 ] from 4 , 1]. Proposition 4.11. Let P1, P2 be two orthogonal projections. If α ∈ [0, π have that 2 ], then we ρW (P2P1)(α) = r(Re(exp(−iα)P2P1)) = kRe(exp(−iα)P2P1)k = ω(Re(exp(−iα)P2P1)). This proposition is significant because if we know r(Re(exp(−iα)P2P1)) and r(Re(exp(−iα)P2(I− 2 ] then, by using Theorem 1.6 and Proposition 4.8, we can 3 , π P1))) for every α ∈ [ π deduce σ(P2P1). Proof. Notice that Re(exp(−iα)P2P1) is an hermitian operator, so r(Re(exp(−iα)P2P1)) = kRe(exp(−iα)P2P1)k = ω(Re(exp(−iα)P2P1)) and the highest positive spectral value of Re(exp(−iα)P2P1) is the highest positive value in the numerical range. In other words, we just need to prove that for all α ∈ [0, π 2 ], the highest positive spectral value of Re(exp(−iα)P2P1) is greater than its lowest negative spectral value. 2 (cos(α)λ ±pλ(1 − sin(α)2λ)) ), we have that According to the notation of the end of the proof of Lemma 3.1 (i.e. vi(λ, α) = 1 Re(exp(−iα)P2P1) ∼(cid:18) v1(C2, α) 0 0 v2(C2, α) (cid:19) . We also have that for all λ ∈ [0, 1] and for all α ∈ [0, π], v1(λ, α) ≥ 0 and v2(λ, α) ≤ 0. Moreover λ ∈ σ(C2) if and only if v1(λ, α) and v2(λ, α) ∈ σ(Re(exp(−iα)P2P1)). Therefore v1(λ, α) − v2(λ, α) = v1(λ, α) + v2(λ, α) = λ cos(α). This last term is positive if α ∈ [0, π 2 ] implies that ρW (P2P1)(α) = r(Re(exp(−iα)P2P1)). 2 ] and negative if α ∈ [ π 2 , π]. So α ∈ [0, π 20 5 Applications to the rate of convergence in the von Neumann-Halperin theorem and to the uncertainty principle 5.1 Applications to the method of alternating projections Von Neumann proved (cf. [Deu01, Chapter 9]) the following theorem: Theorem 5.1. Let M1, M2 be two closed subspaces of H. Then for every x ∈ H we have that: lim n→∞k(PM2 PM1 )nx − PM1∩M2xk = 0. If we set N1 = M1 ∩ (M1 ∩ M2)⊥ and N2 = M2 ∩ (M1 ∩ M2)⊥, we have that N1 ∩ N2 = {0}. In addition, we have (PM2 PM1 )n − PM1∩M2 = (PN2 PN1)n for every n ∈ N. Therefore, the study of the convergence of (PM2 PM1 )n to PM1∩M2 reduces to studying the convergence of (PN2PN1 )n to 0. If one looks at the speed of convergence of (PN2 PN1 )n to 0, we have the dichotomy that either (PN2PN1 )n converges linearly to 0, or (PN2 PN1)n converges arbitrarily slowly to 0. We can characterize arbitrarily slow convergence in many ways; see [BDH09, BGM, DH10a, DH10b] and the references therein. The novelty of the following characterization of arbitrarily slow convergence is in the use of the numerical range of PN2PN1 in items 6 through 8. Proposition 5.2. Let N1, N2 be two closed subspaces of H such that N1 ∩ N2 = {0}. The following assertions are equivalent: 1. (PN2 PN1)n converges arbitrarily slowly to 0 2. kPN2PN1k = 1 3. N⊥1 + N⊥2 is not closed 4. 1 ∈ σ(PN2 PN1 ) 5. cos(N1, N2) = 1 6. 1 ∈ W (PN2 PN1) 7. there exists a sequence (λn) in [0, 1[ such that lim λn = 1 and for every n ∈ N, E (λn) ⊂ W (P2P1) 8. there exists θ < π 6 such that W (PN2PN1 ) ⊂ {z ∈ C,arg(1 − z) ≤ θ}. Proof. We refer to [BDH09, BGM] (see also [Deu01, Chapter 9]) for a proof of the equivalences of the first five assertions. "6 ⇒ 2". As 1 ∈ W (PN2PN1 ), we can find a sequence (xn) such that kxnk = 1 and limn→∞ hPN2PN1 xn, xni = 1. Since we have that hPN2PN1 xn, xni ≤ kPN2 PN1xnkkxnk ≤ kPN2 PN1xnk ≤ kPN2 PN1k ≤ 1, we have that kPN2 PN1k = 1. W (PN2 PN1). "4 ⇒ 6". As 1 ∈ σ(PN2 PN1) and σ(PN2 PN1) ⊂ W (PN2 PN1), we have that 1 ∈ 21 "7 ⇒ 6". This is clear as xλn (0) = "4 ⇒ 7". As N1 ∩ N2 = {0}, 1 is not an eigenvalue of PN2 PN1. So there exist "5 ⇔ 8". This is a consequence of Lemma 1.5. λn ∈ σ(PN2 PN1 ) such that limn λn = 1. The assertion 7 follows from Theorem 1.2. √λn 2 + λn 2 ∈ E (λn) ⊂ W (P2P1). Remark 5.3. In the spirit of [BGM], we can extend "1 ⇔ 6" to a finite number of pro- jection, to obtain the following statement: If PN1 , . . . , PNr are orthogonal projections i=1Ni = {0}, then (PNr . . . PN1)n converges arbitrarily slowly to 0 if and such that ∩r only if 1 ∈ W (PNr . . . PN1). The proof is similar. Remark 5.4. The equivalences between items 5 through 8 still hold if we drop the assumption that N1 ∩ N2 = {0}. 5.2 Applications to annihilating pairs In this section we will give new characterizations of annihilating pairs. First we recall the context. We denote by F the Fourier transform on L2(R). Let S and Σ be two measurable subsets of R. We denote by Mg the operator of multiplication by g ∈ L∞(R) (i.e.: Mg(f ) = gf for f ∈ L2(R)). We denote by 1S the indicator function of the subset S. Set PS = M1S and PΣ = F∗M1ΣF . Definition 5.5. We say that (S, Σ) is an annihilating pair if for every f ∈ L2(R) we have: PSf = PΣf = f ⇒ f = 0. Definition 5.6. We say that (S, Σ) is a strong annihilating pair if there exists a constant c > 0 depending on S, Σ such that for all f ∈ L2(R) we have: kfk2 ≤ c(cid:16)k(I − PS)fk2 + k(I − PΣ)fk2(cid:17) . We want to recall some known facts ([HJ94], and [Len72]) about (strong) annihi- lating pairs. Proposition 5.7. The following assertions are equivalents: 1. (S, Σ) is an annihilating pair 2. 1 + i /∈ W (PS + iPΣ) 3. Ran(PS) ∩ Ran(PΣ) = {0}. Proposition 5.8. The following assertions are equivalents: a. (S, Σ) is a strong annihilating pair b. 1 + i /∈ W (PS + iPΣ) c. Ran(PS) ∩ Ran(PΣ) = {0} and cos(PS, PΣ) < 1 d. kPSPΣk < 1 e. r(PS PΣ) < 1 f. 1 /∈ σ(PSPΣ). The following proposition is a new characterization of annihilating pairs. Proposition 5.9. The following assertions are equivalent to the assertions of Propo- sition 5.7: 1. (S, Σ) is an annihilating pair 4. 1 /∈ W (PSPΣ). 22 Proof. We have that 1 ∈ W (PSPΣ) if and only if there exist h ∈ H such that khk = 1 ad hPSPΣh, hi = 1. This is equivalent to the existence of some h ∈ H such that kPSPΣhk = khk = 1. This last assertion is equivalent to the negation of (3) in Proposition 5.7. Proposition 5.10. The following assertions are equivalent to the assertions of Propo- sition 5.8: a. (S, Σ) is a strong annihilating pair g. 1 /∈ W (PSPΣ) h. ω(PSPΣ) < 1 i. for all α ∈ [0, π j. there exists α ∈ [0, π k. there exists θ < π 3 ], ω(Re(exp(−iα)PSPΣ)) < cos(α) 3 ] such that ω(Re(exp(−iα)PSPΣ)) < cos(α) 6 such that W (PSPΣ) ⊂ {z ∈ C,arg(1 − z) ≤ θ} \ {1}. Proof. "f ⇔ g′′. By Theorem 1.2, 1 ∈ W (PSPΣ) if and only if E (1) ⊂ W (PSPΣ), if and only if 1 ∈ σ(PSPΣ). "e ⇔ h′′. This is a direct consequence of Proposition 1.4. "f ⇒ i′′. This is a consequence of Corollary 4.3. "i ⇒ j′′ This is trivial. "j ⇒ f′′ This is a consequence of Corollary 4.3. "c ⇔ k′′ This consequence of Lemma 1.5, and of the previous Proposition. Acknowledgment I would like to thank Catalin Badea for several discussions and for his help to improve this paper, and Gustavo Corach for pointing out to me the content of Remark 3.6. I would also like to thank the referee for the careful reading of the manuscript and helpful comments. References [BDH09] Heinz H. Bauschke, Frank Deutsch, and Hein Hundal. Characterizing ar- bitrarily slow convergence in the method of alternating projections. Int. Trans. Oper. Res. 16, no. 4, 413 -- 425., 2009. [BGM] Catalin Badea, Sophie Grivaux, and Vladimir Müller. The rate of con- vergence in the method of alternating projections. Algebra i Analiz 23 (2011), no. 3, 1 -- 30; translation in St. Petersburg Math. J. 23 (2012), no. 3, 413 -- 434. [BGM10] Catalin Badea, Sophie Grivaux, and Vladimir Müller. A generalization of the Friederichs angle and the method of alternating projections. C. R. Math. Acad. Sci. Paris 348, no. 1-2, 53 -- 56., 2010. [BL10] Catalin Badea and Yuri Lyubich. Geometric, spectral and asymptotic prop- erties of averaged products of projections in Banach spaces. Studia Math. 201, no. 1, 21 -- 35., 2010. [BS10] A. Bottcher and I.M. Spitkovsky. A gentle guide to the basics of two pro- jections theory. Linear Algebra Appl. 432, no. 6, 1412 -- 1459., 2010. [CM11] G. Corach and A. Maestripieri. Products of orthogonal projections and polar decomposition. Linear Algebra Appl. 434, no. 6, 1594 -- 1609., 2011. 23 [Coh07] Guy Cohen. Iterates of a product of conditional expectation operators. J. Funct. Anal. 242, no. 2, 658 -- 668., 2007. [Cro07] Michel Crouzeix. Numerical Range and functional calculus in Hilbert space. J. Funct. Anal. 244, no. 2, 668 -- 690., 2007. [Cro08] Michel Crouzeix. A functional calculus based on the numerical range: ap- plications. Linear Multilinear Algebra 56, no. 1-2, 81 -- 103., 2008. [DD99] Bernard Delyon and François Delyon. Generalization of von Neumann's spectral sets and integral representation of operators. Bull. Soc. Math. France 127, no. 1, 25 -- 41., 1999. [Deu01] Frank Deutsch. Best Approximation in Inner Product Spaces. Springer- Verlag, New York, 2001. [DH10a] Frank Deutsch and Hein Hundal. Slow convergence of sequences of linear operators I, arbitrarily slow convergence. J. Approx. Theory 162, no. 9, 1701 -- 1716., 2010. [DH10b] Frank Deutsch and Hein Hundal. Slow convergence of sequences of linear operators II, arbitrarily slow convergence. J. Approx. Theory 162, no. 9, 1717 -- 1738., 2010. [Gal04] A. Galántai. Projectors and projection methods. Kluwer Academic Publish- ers, Boston, MA, 2004. [Gal08] A. Galántai. Subspaces, angles and pairs of orthogonal projections. Linear Multilinear Algebra 56, no. 3, 227 -- 260., 2008. [GR97] Karl E. Gustafson and Duggirala K.M. Rao. Numerical Range. Springer, 1997. [Hal69] [HJ94] [Kit03] Paul R Halmos. Two Subspaces. Trans. Amer. Math. Soc.,144, 381 -- 389., 1969. Victor Havin and Burglind Joricke. The Uncertainty Principle in Harmonic Analysis. Springer-Verlag, 1994. Fuad Kittaneh. A numerical radius inequality and an estimate for the numerical radius of the Froebenius companion. Studia Math. 158, no. 1, 11 -- 17., 2003. [KW88] S. Kayalar and H.L. Weinert. Error bounds for the method of alternating projections. Math. Control Signals Systems 1, no. 1, 43 -- 59., 1988. [Len72] Andrew Lenard. The numerical range of a pair of projection. J. Functional Analysis 10 (1972), 410 -- 423., 1972. [Lum61] G. Lumer. Semi inner product spaces. Trans. Amer. Math. Soc. 100 1961 29 -- 43., 1961. [Nee99] Manuela Nees. Products of orthogonal projections as Carleman operators. Integral Equations Operator Theory 35, no. 1, 85 -- 92., 1999. [NN87] Stuart Nelson and Michael Neumann. Generalisations of the projection method with applications to SOR theory for Hermitian positive semi definite linear system. Numer. Math. 51, no. 2, 123 -- 141., 1987. [Roc70] R.T. Rockfellar. Convex Analysis. Princeton University Press, 1970. [RSN90] Frigyes Riesz and Béla Sz.-Nagy. Functional analysis. Dover Books on Advanced Mathematics. Dover Publications Inc., New York, 1990. Trans- lated from the second French edition by Leo F. Boron, Reprint of the 1955 original. 24 [SS10] Valeria Simoncini and Daniel B. Szyld. On the field of values of oblique projections. Linear Algebra Appl. 433 (2010), no. 4, 810 -- 818., 2010. [TUZ03] Hideo Takemoto, Atsushi Uchiyama, and Laszlo Zsido. The σ-convexity of all bounded convex sets in Rn and Cn. Nihonkai Math. J., 14(1):61 -- 64, 2003. 25
1810.01058
1
1810
2018-10-02T04:17:23
Reducing Subspaces of de Branges-Rovnyak Spaces
[ "math.FA" ]
For $b\in H^\infty_1$, the closed unit ball of $H^\infty$, the de Branges-Rovnyak spaces $\mathcal{H}(b)$ is a Hilbert space contractively contained in the Hardy space $H^2$ that is invariant by the backward shift operator $S^*$. We consider the reducing subspaces of the operator $S^{*2}|_{\mathcal{H}(b)}$. When $b$ is an inner function, $S^{*2}|_{\mathcal{H}(b)}$ is a truncated Toepltiz operator and its reducibility was characterized by Douglas and Foias using model theory. We use another approach to extend their result to the case where $b$ is extreme. We prove that if $b$ is extreme but not inner, then $S^{*2}|_{\mathcal{H}(b)}$ is reducible if and only if $b$ is even or odd, and describe the structure of reducing subspaces.
math.FA
math
Reducing Subspaces of de Branges-Rovnyak Spaces Cheng Chu Abstract. For b ∈ H∞ 1 , the closed unit ball of H∞, the de Branges-Rovnyak spaces H(b) is a Hilbert space contractively contained in the Hardy space H 2 that is invariant by the backward shift operator S ∗. We consider the reducing subspaces of the operator S ∗2H(b). When b is an inner function, S ∗2H(b) is a truncated Toepltiz operator and its reducibility was characterized by Douglas and Foias using model theory. We use another approach to extend their result to the case where b is extreme. We prove that if b is extreme but not inner, then S ∗2H(b) is reducible if and only if b is even or odd, and describe the structure of reducing subspaces. 1. Introduction Let D denote the unit disk. Let L2 denote the Lebesgue space of square in- tegrable functions on the unit circle T. The Hardy space H 2 is the subspace of analytic functions on D whose Taylor coefficients are square summable. Then it can also be identified with the subspace of L2 of functions whose negative Fourier coefficients vanish. The space of bounded analytic functions on the unit disk is denoted by H ∞. The Toeplitz operator on the Hardy space H 2 with symbol f in L∞(D) is defined by Tf (h) = P (f h), for h ∈ H 2(D). Here P be the orthogonal projections from L2 to H 2. The unilateral shift operator on H 2 is S = Tz. Let A be a bounded operator on a Hilbert space H. We define the range space M(A) = AH, and endow it with the inner product hAf, AgiM(A) = hf, giH , f, g ∈ H ⊖ KerA. M(A) has a Hilbert space structure that makes A a coisometry on H. Let b be a function in H ∞ 1 , the closed unit ball of H ∞. The de Branges-Rovnyak space H(b) is defined to be the space (I − TbT¯b)1/2H 2. 2010 Mathematics Subject Classification. Primary 47B32. 1 2 CHU We also define the space H(¯b) in the same way as H(b), but with the roles of b and ¯b interchanged, i.e. H(¯b) = (I − T¯bTb)1/2H 2. The spaces H(b) and H(¯b) are also called sub-Hardy Hilbert spaces (the terminology comes from the title of Sarason's book [10]). The space H(b) was introduced by de Branges and Rovnyak [2]. Sarason and several others made essential contributions to the theory [10]. A recent two-volume monograph [4], [5] presents most of the main developments in this area. There are two special cases for H(b) spaces. If b∞ < 1, then H(b) is just a re-normed version of H 2. If b is an inner function, then H(b) = H 2 ⊖ bH 2 is a closed subspace of H 2, the so-called model space (see [6] for a brief survey). Let T be a bounded linear operator on a Hilbert space H. A closed subspace M of H is called a reducing subspace of T if T M ⊂ M and T ∗M ⊂ M . If T has a proper reducing subspace, T is called reducible. The reducing subspaces of shift operators or multiplication operators have been studied in various function spaces: for weighted unilateral shift operators of finite multiplicity, see [11]; for multiplication operators induced by finite Blaschke products on the Bergman space, see [14], [8] and the references therein. Our motivation is the study of reducing subspaces of truncated Toeplitz op- erators on the model space. For an inner function θ and ϕ ∈ L2, the truncated Toeplitz operator Aθ ϕ with symbol ϕ is defined by Aθ ϕf = Pθ(ϕf ), It is known that Aθ for f on the dense subset H(θ) ∩ H ∞ of H(θ). Here Pθ is the orthogonal projection from H 2 to H(θ). [7]). A function f ∈ L2 is called even if f (z) = f (−z), for every z ∈ D, and f is called odd if f (z) = −f (−z), for every z ∈ D. The operator Aθ z is called the compressed shift operator. The reducibility of Aθ z2 is characterized by Douglas and Foias [3] using model theory for contractions [13] as the following. z is always irreducible (see e.g. Theorem 1.1. The operator Aθ z2 is reducible if and only if either θ is even or there exists µ ∈ D such that where p is even. θ(z) = p(z) z + µ 1 + ¯µz , Recently, Li, Yang and Lu found a different proof of Theorem 1.1 and extended it to the case where the symbol of a truncated Toeplitz operator is a Blaschke product of order 2 or 3 [9]. The theory of H(b) spaces is pervaded by a fundamental dichotomy, when b is an extreme point of H ∞ 1 and when it is not. The nonextreme case includes b∞ < 1 and the extreme case includes b is an inner function. Roughly speaking, when b is nonextreme, H(b) behaves similar to H 2, while in the extreme case, H(b) REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 3 is more closely related to the model space. For example, the polynomials belong to H(b) if and only if b is non-extreme (see [10, Chapter IV, V]). Notice that (Aθ z2 )∗ = S ∗2H(θ). Thus, in view of Theorem 1.1, it is natural to consider reducing subspaces of S ∗2H(b) when b is extreme. The main purpose of this paper is to characterize the reducibility of S ∗2H(b) on H(b) in the extreme case and describe the reducing subspaces (Theorem 4.1). We also show that Xb is irreducible for every b. 2. Background on de Branges-Rovnyak Spaces In this section, we present some basic theory of de Branges-Rovnyak spaces and the results we shall use later. The relation between H(b) and H(¯b) can be found in [10, II-4]. Here we use h , ib to denote the inner product of H(b). Theorem 2.1. A function f belongs to H(b) if and only if T¯bf belongs to H(¯b). If f1, f2 ∈ H(b), then hf1, f2ib = hf1, f2i2 + hT¯bf1, T¯bf2i¯b. Let b ∈ H ∞ 1 . Let ρ = 1 − b2 on T and let H 2(ρ) be the closure of polynomials 2π ) (we will keep using these notations in the remaining of this in L2(ρ) = L2(T, ρ dθ paper). The Cauchy transform Kρ is the mapping from L2(ρ) to H 2 defined by Kρf = P (ρf ). In the theory of H(b) spaces, H(¯b) is often more amenable than H(b) because of a representation theorem for H(¯b) [10, III-2]. Theorem 2.2. The operator Kρ is an isometry from H 2(ρ) to H(¯b). The operator on H 2(ρ) of multiplication by the independent variable will be denoted by Zρ. We have the intertwining relation [10, III-3] (2.1) KρZ ∗ ρ = S ∗Kρ. The space H(b) is invariant under S ∗ = T¯z [10, II-7], and the restriction of S ∗ is a contraction. We use Xb to denote S ∗H(b). This operator can serve as a model for a large class of Hilbert space contractions [2], [1]. The following identity shows the difference between Xb and S ∗ [10, II-9]. Theorem 2.3. Let b ∈ H ∞ 1 . For every f ∈ H(b), X ∗ b f = Sf − hf, S ∗bibb. If x and y are in a Hilbert space H, we shall use x ⊗ y to be the following rank one operator on H It is obvious that (x ⊗ y)(f ) = hf, yiH · x, f ∈ H. and if A, B are bounded linear operators on H, then A(x ⊗ y)B = (Ax) ⊗ (B∗y). (x ⊗ y)∗ = y ⊗ x, 4 CHU It could be misleading to write the identity in Theorem 2.3 as X ∗ b = S − b ⊗ S ∗b because b may not be in H(b). But it is known that S ∗b ∈ H(b) [10, II-8], and we have (2.2) I − XbX ∗ b = (S ∗b) ⊗ (S ∗b). The space H(b) is a reproducing kernel Hilbert space with kernel function: When b is extreme, we have the following identity (see e.g. [5, Theorem 25.11]). kb w(z) = 1 − b(w)b(z) 1 − ¯wz . Lemma 2.1. Let b be an extreme point in H ∞ 0 ⊗ kb 0. b Xb = kb I − X ∗ 1 . Then For an inner function θ, S ∗θ is a cyclic vector of (Aθ z)∗. A similar result holds for extreme functions (see e.g. [5, Section 26.6]). Theorem 2.4. If b is extreme, then H(b) = Span{S ∗nb : n > 0}. 3. An Equivalent Condition for the Reducibility In this section we first prove that Xb is irreducible for every b. The idea in the proof will be used to study X 2 b . Theorem 3.1. Let b ∈ H ∞ 1 . Then Xb is not reducible. Proof. Suppose Xb is reducible. Then H(b) = M1 ⊕b M2, where M1, M2 are nontrivial reducing subspaces of Xb. Note that for every f ∈ M1, g ∈ M2, (I − XbX ∗ b )f ⊥ (I − XbX ∗ b )g in H(b). By Lemma 2.2, Then one of the two range spaces dim((I − XbX ∗ b )H(b)) 6 1. (I − XbX ∗ must be 0. WLOG, we may assume b )M1, (I − XbX ∗ b )M2 i.e. for every f ∈ M1, (I − XbX ∗ b )M1 = 0, 0 = (I − XbX ∗ b )f = hf, S ∗bibS ∗b. Thus f is orthogonal to S ∗b in H(b) and then S ∗b ∈ M2. Since M2 is invariant under S ∗, we have Span{S ∗nb : n > 0} ⊂ M2. REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 5 If b is extreme, it follows from Theorem 2.4 that M2 = H(b), which is a con- tradiction. If b is nonextreme, then polynomials are dense in H(b). We see from Theorem 2.3 that M1 is invariant under both S and S ∗. Pick a nonzero function h ∈ M1, then ∞ h(z) = for some k > 0 with hk 6= 0. Thus hjzj, Xj=k 1 hk (I − Sk+1S ∗k+1)h = zk ∈ M1, which implies that M1 contain all the polynomials. So M1 = H(b), which is a contradiction. (cid:3) For the extreme case, we have the following equivalent condition for the re- ducibility of X 2 b . Theorem 3.2. Let b be an extreme point in H ∞ 1 . Then X 2 b is reducible if and only if there exist complex numbers α, β, αβ 6= 1, such that for every n, m > 0, (3.1) S ∗2m(S ∗b + αS ∗2b) ⊥ S ∗2n(βS ∗b + S ∗2b) in H(b). In this case the reducing subspaces of X 2 b are given by H(b) = M1 ⊕b M2, where (3.2) and (3.3) M1 = Span{S ∗2n(S ∗b + αS ∗2b) : n > 0} M2 = Span{S ∗2n(βS ∗b + S ∗2b) : n > 0}. Proof. Suppose (3.1) holds, then take M1, M2 as in (3.2), (3.3). It is clear b (or S ∗2) and are orthogonal in H(b). By that M1, M2 are invariant under X 2 Theorem 2.4, we have Thus H(b) = Span{M1, M2}. H(b) = M1 ⊕b M2, and X 2 b is reducible. Next we assume X 2 b is reducible. Then where M1, M2 are nontrivial reducing subspaces of X 2 M1, g ∈ M2, b . Note that for every f ∈ H(b) = M1 ⊕b M2, (I − X 2 b X ∗2 b )f ∈ M1, (I − X 2 b X ∗2 b )g ∈ M2. Then (I − X 2 b X ∗2 b )f ⊥ (I − X 2 b X ∗2 b )g 6 CHU in H(b). Using (2.2), we have b X ∗2 I − X 2 (3.4) b = I − Xb(XbX ∗ b )X ∗ b = I − Xb(I − XbX ∗ b )X ∗ b = I − XbX ∗ = S ∗b ⊗ S ∗b + S ∗2b ⊗ S ∗2b. b − Xb(S ∗b ⊗ S ∗b)X ∗ b Note that when b is extreme, S ∗b and S ∗2b are linearly independent. Thus dim(I − X 2 b )H(b)) = 2. Suppose one of the two range spaces b X ∗2 (I − X 2 b X ∗2 b )M1, (I − X 2 b X ∗2 b )M2 is zero, say (I − X 2 b X ∗2 b )M1 = 0. By (3.4), we see that every function in M1 is orthogonal to S ∗b and S ∗2b in H(b), which implies S ∗b, S ∗2b are in M2. Since M2 is invariant for X 2 b , using Theorem 2.4 we see that This is a contradiction. Therefore, we must have H(b) = Span{S ∗nb : n > 0} ⊂ M2. dim(I − X 2 b X ∗2 b )M1 = dim(I − X 2 b X ∗2 b )M2 = 1. This means, WLOG, there exist complex numbers α, β such that b )M1 = Span{S ∗b + αS ∗2b} ⊂ M1, (I − X 2 b X ∗2 (I − X 2 b X ∗2 b )M1 = Span{βS ∗b + S ∗2b} ⊂ M2. Since M1, M2 are invariant under X 2 b , we have Span{S ∗2n(S ∗b + αS ∗2b) : n > 0} ⊂ M1, Span{S ∗2n(βS ∗b + S ∗2b) : n > 0} ⊂ M2. Using Theorem 2.4, we obtain H(b) = M1 ∪ M2, and thus (3.2), (3.3) hold. The relation (3.1) follows from M1 ⊥b M2. Note that αβ 6= 1; otherwise M1 = M2 = 0. (cid:3) 4. Main Results In this section, we analyze the condition (3.1) and characterize the reducibility b when b is extreme but not inner. of X 2 Lemma 4.1. Let b be an extreme point in H ∞ 1 . Then for every n > 1, n−1 I − X ∗n b X n b = Xj=0 (X ∗j b kb 0) ⊗ (X ∗j b kb 0). REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 7 Proof. This proof is by induction on n. For n = 1, the equality is exactly the one in Lemma 2.1. Assume that the equality holds for some n > 2. Then, using once again Lemma 2.1 and the induction hypothesis, we have X ∗n b X n b = X ∗ b (X ∗n−1 b X n−1 b )Xb n−2 =X ∗(I − =X ∗X − n−2 Xj=0 Xj=0 (X ∗j b kb 0) ⊗ (X ∗j b kb 0))Xb X ∗ b (X ∗j b kb 0) ⊗ (X ∗j b kb 0)Xb =I − kb 0 ⊗ kb 0 − n−1 n−2 (X ∗(j+1) b kb 0) ⊗ (X ∗(j+1) b kb 0) Xj=0 (X ∗j b kb 0) ⊗ (X ∗j b kb 0). =I − Xj=0 Lemma 4.2. Let b be an extreme point in H ∞ 1 and let f, g ∈ H(b). Then (cid:3) (4.1) hX 2m b f, X 2n b gib = 0, for every m, n > 0 if and only if the following hold (1) for every k > 0, (4.2) hT¯bf, T¯bX 2k b gi¯b = hT¯bg, T¯bX 2k b f i¯b = 0. (2) for every m, n > 0, hS ∗2mf, S ∗2ngi2 = 0, there exist functions F, G ∈ H 2 and complex numbers a0, b0, a1, b1 i.e. such that (4.3) f (z) = F (z2)(a0 + a1z), g(z) = G(z2)(b0 + b1z) and Proof. Let a0 ¯b0 + a1 ¯b1 = 0. f (z) = ∞ Xk=0 fkzk and g(z) = gkzk. ∞ Xk=0 Suppose (4.1) holds. Then for m 6 n, we have (4.4) 0 = hX 2m b f, X 2n b gib = hX ∗2m b X 2m b f, X 2n−2m b gib. 8 CHU By Lemma 4.1, we have (I − X ∗2m b X 2m b )f = f − X ∗2m b X 2m b f 2m−1 = = = = Then Xj=0 Xj=0 Xj=0 Xj=0 2m−1 2m−1 X ∗2m b X 2m b f = f − This together with (4.4) implies hf, X ∗j b kb 0ib · (X ∗j b kb 0) 2m−1 hX j b f, kb 0ib · (X ∗j b kb 0) (S ∗jf )(0) · (X ∗j b kb 0) fj · X ∗j b kb 0. 2m−1 Xj=0 fj · X ∗j b kb 0. 0 = hf − 2m−1 Xj=0 fj · X ∗j b kb 0, X 2n−2m b gib 2m−1 = hf, X 2n−2m b gib − h = hf, X 2n−2m b gib − = hf, X 2n−2m b gib − = hf, X 2n−2m b gib − (4.5) = hf, X 2n−2m b gib − fj · X ∗j b kb 0, X 2n−2m b fj · hX ∗j b kb 0, X 2n−2m b gib gib fj · hkb 0, X 2n−2m+j b gib fj · hX 2n−2m+j b g, kb 0ib fj · g2n−2m+j. 2m−1 2m−1 2m−1 Xj=0 Xj=0 Xj=0 Xj=0 Xj=0 2m−1 Replacing n, m in (4.5) by n + 1, m + 1 respectively, we have (4.6) 0 = hf, X 2n−2m b gib − 2m+1 Xj=0 fj · g2n−2m+j. Subtracting (4.6) by (4.5) implies (4.7) f2mg2n + f2m+1g2n+1 = 0, REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 9 for m 6 n. A similar argument shows that (4.7) also holds for n 6 m. Thus we have for every n, m > 0, the two vectors (f2m, f2m+1), (g2n, g2n+1) are orthogonal in C2. It is easy to check f, g must have the form (4.3). In particular, we have (4.8) hf, X 2k b gi2 = hg, X 2k b f i2 = 0, It follows from (4.5) and (4.7) that hf, X 2k b gib = hg, X 2k b f ib = 0, for every k > 0. for every k > 0. This together with (4.8) and Theorem 2.1 give (4.2). The sufficiency follows easily from the calculation in (4.5). Remark 4.1. When b is an inner function, H(¯b) is trivial and then (4.2) is b is more restrictive automatically satisfied. One may expect the reducibility of X 2 if b is not inner. We shall see it is true in the remaining of this section. When b is extreme, the following Lemma will be used to calculate the inner products in (4.2). Lemma 4.3. Let b be an extreme point in H ∞ 1 . Let ρ = 1 − b2 on T. Then for every m, n > 1, (cid:3) hT¯bS ∗mb, T¯bS ∗nbi¯b =  − hzn−m, b2i2, m < n, − hb2, zm−ni2, m > n, 1 − b2 m = n. 2, Proof. Suppose m 6 n. Using the intertwining relation (2.1), we can easily get Thus KρZ ∗n ρ = S ∗nKρ. KρZ ∗n ρ 1 =S ∗nKρ1 = S ∗nP (ρ) = S ∗nP (1 − b2) = − S ∗nP (b2) = −S ∗nT¯bb = −T¯bS ∗nb. By Theorem 2.2, we have hT¯bS ∗mb, T¯bS ∗nbi¯b = hKρZ ∗m ρ 1, KρZ ∗n ρ 1i¯b = hZ ∗m ρ 1, Z ∗n ρ 1iL2(ρ). If b is extreme, then H 2(ρ) = L2(ρ) [12], which implies Zρ is a unitary operator. Then hT¯bS ∗mb, T¯bS ∗nbi¯b =hZ ∗m ρ 1, Z ∗n ρ 1iL2(ρ) = hZ n−m ρ =hzn−m, 1i2 − hzn−m, b2i2 =( − hzn−m, b2i2, m < n, 1 − b2 2, m = n. 1, 1iL2(ρ) = hzn−m, 1iL2(ρ) (cid:3) 10 CHU We also need the following three elementary results. Lemma 4.4. Let b ∈ H ∞ 1 . Then Proof. Let Then lim n→∞ hzn, b2i2 = 0. b(z) = bkzk. ∞ Xk=0 ∞ hzn, b2i2 = hznb, bi2 = bkbn+k ∞ 1 2 ( bn+k2) 1 ∞ Xk=0 Xk=0 ∞ 6 ( Xk=0 bk2) Since b2 6 1, we have 2 = b2( bn+k2) 1 2 . Xk=0 lim n→∞ hzn, b2i2 6 ( lim n→∞ bn+k2) 1 2 = 0. ∞ Xk=0 (cid:3) Lemma 4.5. Let b ∈ H ∞. Then b2 is even if and only if b is even or odd. Proof. Let b(z) = b0(z) + zb1(z), where b0, b1 are even functions. Then b2 is even if and only if b0(z) + zb1(z)2 = b0(z) − zb1(z)2, which is equivalent to b0zb1 ≡ 0. Then the conclusion follows easily. (cid:3) Lemma 4.6. Let α, β ∈ C with αβ 6= 0 or 1. Let {an}∞ n=0 be a sequence of complex numbers but not the zero sequence. Suppose lim n→∞ an = 0 and for every n > 1, the following conditions hold. (4.9) (4.10) Then we have either a2n+1 + (α + ¯β)a2n + α ¯βa2n−1 = 0, a2n+1 + ( 1 ¯α + 1 β )a2n + 1 ¯αβ a2n−1 = 0. β = − ¯α, and a2n−1 = 0, for every n > 1 or α = β = 1. REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 11 Proof. Subtracting (4.10) from (4.9), we have (4.11) (α + ¯β − 1 ¯α − 1 β )a2n + (α ¯β − 1 ¯αβ )a2n−1 = 0. Since {an}∞ n=0 is nonzero, we have the following four cases. Case I: α + ¯β − 1 ¯α − 1 β = α ¯β − 1 ¯αβ = 0. Then we have αβ = 1 and 0 =α + ¯β − 1 ¯α − = 1 ¯α (1 − α2) + 1 β 1 β = 1 ¯α (1 − α2) + 1 β (1 − β2) (1 − 1 α2 ) = 1 − α2 ¯α (1 − 1 αβ ). Thus α = β = 1. Case II: α + ¯β = 1 ¯α + 1 β and a2n−1 = 0, for every n > 1. Then (4.9) implies β = − ¯α. we have Case III: α ¯β = 1 ¯αβ and a2n = 0, for every n > 1. Then αβ = 1 and by (4.9), a2n+1 = αβ · a2n−1 = a2n−1. Since an tends to 0, we have a2n−1 = 0 and thus {an}∞ which contradicts the assumption. Case IV: α + ¯β − 1 β 6= 0 and α ¯β − 1 ¯αβ 6= 0. Then by (4.11), n=0 is the zero sequence, ¯α − 1 ¯αβ − α ¯β 1 a2n = α + ¯β − 1 ¯α − 1 β a2n−1 = 1 − αβ2 βα2 + ¯αβ2 − β − ¯α a2n−1. Put this in (4.9), we have a2n+1 = −(α + ¯β)a2n − α ¯βa2n−1 = −(cid:16)(α + ¯β) 1 − αβ2 βα2 + ¯αβ2 − β − ¯α = αβ2 + ¯βα2 − α − ¯β βα2 + ¯αβ2 − β − ¯α a2n−1. + α ¯β(cid:17)a2n−1 Thus a2n+1 = a2n−1 and, similar to Case III, a2n−1 = 0, for every n > 1. From (4.9), (4.10), we see that either a2n = 0, for every n > 1 or α + ¯β = 1 β = 0. They are both excluded by the assumptions. (cid:3) ¯α + 1 Now we are ready to prove the main Theorem. Theorem 4.1. Let b be an extreme point in H ∞ b is reducible if and only if b is even or odd. 1 . If b is not an inner function, If b is even, the reducing then X 2 subspaces of X 2 b are M = Span{(S ∗2nb)(z + α) : n > 1} M ⊥ = Span{(S ∗2nb)(− ¯αz + 1) : n > 1}, with for all α ∈ C. 12 CHU If b is odd, the reducing subspaces of X 2 b are with M = Span{S ∗2n−1b : n > 1} M ⊥ = Span{S ∗2nb : n > 1}. Proof. Necessity. We assume X 2 b is reducible and b is not inner. Let ∞ b(z) = bkzk. Xk=0 By Theorem 3.2, there exists α, β ∈ C such that αβ 6= 1 and (3.1) holds. Let f = S ∗b + αS ∗2b, g = βS ∗b + S ∗2b. Then f, g are in H(b), and using Lemma 4.2, we have b gi¯b = hT¯bg, T¯bX 2n for every n > 0. If n > 1, using Lemma 4.3, we have hT¯bf, T¯bX 2n b f i¯b = 0, 0 =hT¯bf, T¯bX 2n b gi¯b =hT¯bS ∗b + αT¯bS ∗2b, βT¯b(S ∗)2n+1b + T¯b(S ∗)2n+2bi¯b = ¯βhT¯bS ∗b, T¯b(S ∗)2n+1bi¯b + hT¯bS ∗b, T¯b(S ∗)2n+2bi¯b + α ¯βhT¯bS ∗2b, T¯b(S ∗)2n+1i¯b + αhT¯bS ∗2b, T¯b(S ∗)2n+2bi¯b = − ¯βhz2n, b2i2 − hz2n+1, b2i2 − α ¯βhz2n−1, b2i2 − αhz2n, b2i2, which can be simplified to (4.12) hz2n+1, b2i2 + (α + ¯β)hz2n, b2i2 + α ¯βhz2n−1, b2i2 = 0. Similarly, hT¯bg, T¯bX 2n b f i¯b = 0 implies (4.13) ¯αβhz2n+1, b2i2 + (¯α + β)hz2n, b2i2 + hz2n−1, b2i2 = 0. If α = β = 0, then (4.13) implies for every n > 1, 0 = hz2n−1, b2i2 = hb2, ¯z2n−1i2. This means b is even or odd by Lemma 4.5. If α = 0 and β 6= 0, using (4.12), (4.13), we have and Thus hz2n+1, b2i2 + ¯βhz2n, b2i2 = 0, βhz2n, b2i2 + hz2n−1, b2i2 = 0. hz2n+1, b2i2 = ¯β β hz2n−1, b2i2 = hz2n−1, b2i2. REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 13 By Lemma 4.4, we see that for every n > 1, hz2n−1, b2i2 = 0. Thus (4.12) shows that hzn, b2i2 = 0, which implies b is inner. A similar argument works for the case when β = 0 and α 6= 0. Next, suppose αβ 6= 0. Rewrite (4.13) as (4.14) hz2n+1, b2i2 + ( 1 ¯α + 1 β )hz2n, b2i2 + 1 ¯αβ hz2n−1, b2i2 = 0. Consider the sequence {hzn, b2i2}∞ n=1. If it is the zero sequence, then b is an inner function. Otherwise by Lemma 4.4 and (4.12), (4.14), it satisfies the assumptions in Lemma 4.6. Then we have the following two cases: Case I: β = − ¯α, hz2n−1, b2i2 = 0, for every n > 1. Case II: α = β = 1. By condition (2) in Lemma 4.2, we have for every n > 0, hS ∗2nf, S ∗2ngi2 = 0. Then 0 =hS ∗2nf, S ∗2ngi2 = hS ∗2n+1b + αS ∗2n+2b, βS ∗2n+1b + S ∗2n+2bi2 = ¯βS ∗2n+1b2 2 + αS ∗2n+2b2 2 + hS ∗2n+1b, S ∗2n+2bi2 + α ¯βhS ∗2n+2b, S ∗2n+1bi2. For simplicity, let cn = hS ∗2n+1b, S ∗2n+2bi2. Since we obtain (4.15) S ∗2n+2b2 2 = S ∗2n+1b2 2 − b2n+12, ( ¯β + α)S ∗2n+1b2 2 − αb2n+12 + cn + α ¯β ¯cn = 0. In Case I, b2 is even, and Lemma 4.5 implies that b is even or odd. Thus cn = 0 and (4.15) becomes αb2n+12 = 0, which implies b2n+1 = 0 and b is even. In Case II, taking conjugate on (4.15), we get (4.16) (β + ¯α)S ∗2n+1b2 2 − ¯αb2n+12 + ¯αβcn + ¯cn = 0. Multiplying (4.15) by ¯αβ and using α = β = 1, we have (4.17) (¯α + β)S ∗2n+1b2 2 − βb2n+12 + ¯αβcn + ¯cn = 0. By (4.16), (4.17), we have (¯α − β)b2n+12 = 0. Note that ¯α 6= β because αβ 6= 1. We see that b2n+1 = 0, which means b is even. Using (4.12), we see that if b is not inner, then β = − ¯α. Sufficiency. Let and M1 = Span{S ∗2n(S ∗b + αS ∗2b) : n > 0} M2 = Span{S ∗2n(− ¯αS ∗b + S ∗2b) : n > 0}. We show that M1, M2 are reducing subspaces of X 2 b for appropriate choices of α. By Theorem 3.2 and Lemma 4.2, we need to verify (4.2) and (4.3) when β = − ¯α. 14 CHU Note that hz2n−1, b2i2 = 0, for every n > 1. whenever b is even or odd. For (4.2), if n > 1, (4.2) follows from (4.12), (4.13) and the above relation. When n = 0, using Lemma 4.3, we have hT¯b(S ∗b + αS ∗2b), T¯b(− ¯αS ∗b + S ∗2b)i¯b = − αT¯bS ∗b2 = − α(1 − b2 ¯b + αT¯bS ∗2b2 2) + α(1 − b2 ¯b + hT¯bS ∗b, T¯bS ∗2bi¯b − α2hT¯bS ∗2b, T¯bS ∗bi¯b 2) − hz, b2i2 + α2h¯z, b2i2 = 0. If b is odd and α = 0, it is obvious that (4.3) holds. If b is even, then S ∗2b is also even and S ∗b = zS ∗2b. We can write and Thus (4.3) is satisfied. S ∗b + αS ∗2b = (S ∗2b)(z + α), − ¯αS ∗b + S ∗2b = (S ∗2b)(− ¯αz + 1). (cid:3) References [1] L. de Branges and J. Rovnyak, Appendix on square summable power series, Perturbation Theory and its Applications in Quantum Mechanics, John Wiley and Sons, New York, 1966. , Square summable power series, Holt, Rinehart, and Winston, New York, 1966. [2] [3] R.G. Douglas and C. Foias, On the structure of the square of a C0(1) operator, Modern operator theory and applications, Oper. Theory Adv. Appl., vol. 170, Birkhauser, Basel, 2007. [4] E. Fricain and J. Mashreghi, The theory of H(b) spaces, New Mathematical Monographs, 20, vol. 1, Cambridge University Press, Cambridge, 2016. [5] , The theory of H(b) spaces, New Mathematical Monographs, 21, vol. 2, Cambridge University Press, Cambridge, 2016. [6] S. Garcia and W. Ross, Model spaces: a survey, Contemp. Math., vol. 638, Amer. Math. Soc., Providence, RI, 2015. [7] S.R. Garcia, W.T. Ross, and W.R. Wogen, C ∗-algebras generated by truncated Toeplitz oper- ators, Concrete operators, spectral theory, operators in harmonic analysis and approximation, Oper. Theory Adv. Appl., vol. 236, Birkhauser, Basel, 2014. [8] K. Guo and H. Huang, Multiplication Operators on the Bergman Space, Lecture Notes in Mathematics, vol. 2154, Springer, 2015. [9] Yufei Li, Yixin Yang, and Yufeng Lu, Reducibility for a class of truncated Toeplitz operators. preprint. [10] D. Sarason, Sub-Hardy Hilbert spaces in the unit disk, University of Arkansas Lecture Notes, Wiley, New York, 1994. [11] M. Stessin and K. Zhu, Reducing subspaces of weighted shift operators, Proc. Amer. Math. Soc. 130 (2002), no. 9, 2631 -- 2639. [12] G. Szego, Beitrage zur Theorie der Toeplitzschen Formen, Math. Z. 6 (1920), 167 -- 202. [13] B. Szokefalvi-Nagy and C. Foia¸s, Harmonic analysis of operators on Hilbert space, North Holland, Amsterdam, 1970. REDUCING SUBSPACES OF DE BRANGES-ROVNYAK SPACES 15 [14] K. Zhu, Reducing subspaces for a class of multiplication operators, J. London Math. Soc. 62 (2000), no. 2, 553 -- 568. Department of Mathematics, Vanderbilt University, Nashville, Tennessee, USA E-mail address: [email protected]
1709.00585
1
1709
2017-09-02T14:16:53
Generalized frames and controlled operators in Hilbert space
[ "math.FA" ]
Controlled frames and g-frames were considered recently as generalizations of frames in Hilbert spaces. In this paper we generalize some of the known results in frame theory to controlled g-frames. We obtain some new properties of controlled g-frames and obtain new controlled g-frames by considering controlled g-frames for its components. And we obtain some new resolutions of the identity. Furthermore, we study the stabilities of controlled g-frames under small perturbations.
math.FA
math
GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE DONGWEI LI AND JINSONG LENG Abstract. Controlled frames and g-frames were considered recently as generalizations of frames in Hilbert spaces. In this paper we generalize some of the known results in frame theory to con- trolled g-frames. We obtain some new properties of controlled g-frames and obtain new controlled g-frames by considering controlled g-frames for its components. And we obtain some new reso- lutions of the identity. Furthermore, we study the stabilities of controlled g-frames under small perturbations. 1. Introduction Frames were first introduced in 1952 by Duffin and Schaeffer [8] to study some problems in non- harmonic Fourier series, and were widely studied from 1986 since the great work by Daubechies et al. [7]. To date, frame theory has broad applications in pure mathematics, for instance, Kadison- Singer problem and statistics, as well as in applied mathematics, computer science and engineering applications. We refer to [4, 11, 12, 13, 14, 17] for an introduction to frame theory and its applica- tions. In 2006, Sun [18] introduced the concept of g-frame. G-frames are generalized frames, which include ordinary frames, bounded invertible linear operators, fusion frames, as well as many recent generalizations of frames. For more details see [9, 19]. G-frames and g-Riesz bases in Hilbert spaces have some properties similar to those of frames, but not all the properties are similar [18]. Controlled frames for spherical wavelets were introduced in [5] and were reintroduced recently to improve the numerical efficiency of iterative algorithms [2]. The role of controller operators is like the role of preconditions matrices or operators in linear algebra. Controlled g-frames were introduce by Khosravi et al. [16]. Now, many excellent results of con- trolled frames have been achieved and applied successfully [2, 3], which properties of the controlled frames may be extended to the g-frames? It is a tempting subject because of the complexity of the structure of g-frames compared with conventional frames. In this paper, we give some new properties of controlled g-frames and construct new controlled g-frames from a given controlled g-frame, and we generalize some of known results in g-frames to controlled g-frames in Section 2. In section 3 we obtain some new resolutions of the identity with controlled g-frames, and in Section 4 we study the stability of controlled g-frames under small perturbations. Throughout this paper, H and K are two separable Hilbert spaces and {H i : i ∈ I} is a sequence of subspaces of K , where I is a subset of Z . L(H ,H i) is the collection of all bounded linear operators from H into H i, and GL(H ) denotes the set of all bounded linear operators which It is easy to see that if T, U ∈ GL(H ), then T ∗, T and T U are also in have bounded inverse. 2000 Mathematics Subject Classification. 42C15, 46C05. Key words and phrases. g-frame, controlled operators, controlled g-frame, perturbation. 1 2 DONGWEI LI AND JINSONG LENG GL(H ). Let GL+(H ) be the set of positive operators in GL(H ). Also IH denotes the identity operator on H . Note that for any sequence {H i : i ∈ I} of Hilbert spaces, we can always find a large Hilbert space K such that for all i ∈ I, H i ⊂ K (for example K = ⊕i∈IH i). Definition 1.1. A sequence Λ = {Λi ∈ L(H ,H i) : i ∈ I} is called a generalized frame, or simply a g-frame, for H with respect to {H i : i ∈ I} if there exist constants 0 < A ≤ B < ∞ such that Akfk2 ≤Xi∈I kΛifk2 ≤ B kfk2 , for all f ∈ H . (1) The numbers A and B are called g-frame bounds. We call Λ a tight g-frame if A = B and Parseval g-frame if A = B = 1. If the second inequality in (1) holds, the sequence is called g-Bessel sequence. Λ = {Λi ∈ L(H ,H i) : i ∈ I} is called a g-frame sequence, if it is a g-frame for span{Λ∗ For each sequence {H i}i∈I , we define the space (Pi∈I ⊕H i)ℓ2 by ⊕H i)ℓ2 =(cid:8){fi}i∈I : fi ∈ H i, i ∈ I and Xi∈I kfik2 < +∞(cid:9) (Xi∈I i (H )}i∈I . with the inner product defined by h{fi},{gi}i =Xi∈I hfi, gii . Definition 1.2. Let Λ = {Λi ∈ L(H ,H i) : i ∈ I} be a g-frame for H . Then the synthesis operator for Λ = {Λi ∈ L(H ,H i) : i ∈ I} is the operator ΘΛ : (Xi∈I ⊕H i)ℓ2 −→ H defined by The adjoint Θ∗ Λ of the synthesis operator is called analysis operator which is given by ΘΛ({fi}i∈I ) =Xi∈I Λ∗ i (fi). By composing ΘΛ and Θ∗ Θ∗ Λ : H −→ (Xi∈I ⊕H i)ℓ2 , T ∗(f ) = {Λif}i∈I. Λ, we obtain the g-frame operator SΛ : H −→ H , SΛf = ΘΛΘ∗ Λ∗ i Λif. Λf =Xi∈I It is easy to see that g-frame operator is a bounded, positive and invertible operator. 2. Controlled g-frames and constructing new controlled g-frames Controlled g-frames with two controlled operators were studied in [15, 16]. Next, we give the definition of controlled g-frames. GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 3 Definition 2.1. Let T, U ∈ GL(H ). The family Λ = {Λi ∈ L(H ,H i) : i ∈ I} is called a (T, U )- controlled g-frame for H , if Λ is a g-Bessel sequence and there exist constants 0 < A ≤ B < ∞ such that (2) hΛiT f, ΛiU fi ≤ Bkfk2, ∀f ∈ H . Akfk2 ≤Xi∈I A and B are called the lower and upper controlled frame bounds, respectively. If U = IH , we call Λ = {Λi} a T -controlled g-frame for H with bounds A and B. If the second part of the above inequality holds, it is called (T, U )-controlled g-Bessel sequence with bound B. Let Λ = {Λi ∈ L(H ,H i) : i ∈ I} be a (T, U )-controlled g-frame for H , then the (T, U )-controlled g-frame operator is defined by ST ΛU : H −→ H , ST ΛU f =Xi∈I U ∗Λ∗ i ΛiT f, ∀f ∈ H . It follows from the definition that for a g-frame, this operator is positive and invertible and also ST ΛU = U ∗SΛT . AIH ≤ ST ΛU ≤ BIH , For convenience we state the following lemma. Lemma 2.2. [2] Let T : H −→ H be a linear operator. Then the following conditions are equiva- lent: (1) There exist m > 0 and M < ∞, such that mIH ≤ T ≤ M IH . (2) T is positive and there exist m > 0 and M < ∞, such that mkfk2 ≤ kT 1/2fk2 ≤ Mkfk2. (3) T ∈ GL+(H ). Proposition 2.3. Let T, U ∈ GL+(H ) and Λ = {Λi ∈ L(H ,H i) : i ∈ I} be a family of operator. Then the following statements hold. (1) If {Λi : i ∈ I} is a (T, U )-controlled g-frame for H , then {Λi : i ∈ I} is a g-frame for H . (2) If {Λi : i ∈ I} is a g-frame for H and T, U ∈ GL+(H ), which commute with each other and commute with SΛ, then {Λi : i ∈ I} is a (T, U )-controlled g-frame for H . Proof. 1. For f ∈ H , since the operator SΛ(f ) = (U ∗)−1ST ΛU T −1(f ) =Xi∈I Λ∗ i Λif is well defined, we show that it is a bounded and invertible operator. It is also a positive linear operator on H because Hence, kΛifk2. hSΛf, fi =Xi∈I T ΛU U ∗k ≤ kTkkS−1 T ΛUkkU ∗k ≤ 1 AkTkkU ∗k, kS−1 Λ k = kT S−1 which A is the lower frame bound of (T, U )-controlled g-frame {Λi : i ∈ I}. So SΛ ∈ GL+(H ). Therefore, by Lemma 2.2, we have CIH ≤ SΛ ≤ DIH for some 0 < C ≤ D < ∞. So the result holds. 4 DONGWEI LI AND JINSONG LENG 2. Let {Λi : i ∈ I} be a g-frame with bounds C, D and m, m′ > 0, M, M ′ < ∞ so that mIH ≤ T ≤ M IH , m′IH ≤ U ∗ ≤ M ′IH . By Lemma 2.2, then we have because T commutes with SΛ. Again U ∗ commutes with SΛT and then mCIH ≤ SΛT ≤ M DIH mm′CIH ≤ ST ΛU ≤ M M ′DIH . This completes the proof. (cid:3) Corollary 2.4. Let T, U ∈ GL(H ) and {Λi : i ∈ I} be a g-frame with frame operator SΛ. If U ∗SΛT is positive, then {Λi : i ∈ I} is a (T, U )-controlled g-frame for H . The Theorem 2.8 of [1] leads to the following result. Proposition 2.5. Let T, U ∈ GL(H ) and {Λi : i ∈ I} be a (T, U )-controlled g-frame for H with lower and upper bounds A and B, respectively. Let {Γi : i ∈ I} be a g-complete family of bounded operator. If there exists a number 0 < R < A such that 0 ≤Xi∈I hU ∗(Λ∗ i Λi − Γ∗ i Γi)T f, fi ≤ Rkfk2, ∀f ∈ H , then {Γi : i ∈ I} is also a (T, U )-controlled g-frame for H . Proof. Let {Λi : i ∈ I} be a (T, U )-controlled g-frame for H with frame bounds A and B, for any f ∈ H , we have hU ∗Λ∗ i ΛiT f, fi ≤ Bkfk2. Akfk2 ≤Xi∈I hU ∗(Γ∗ Hence Xi∈I hU ∗Γ∗ i ΓiT f, fi =Xi∈I i Γi − Λ∗ i Λi)T f, fi +Xi∈I ≤ Rkfk2 + Bkfk2 = (R + B)kfk2. hU ∗Λ∗ i ΛiT f, fi On the other hand hU ∗Γ∗ Xi∈I hU ∗Λ∗ hU ∗Λ∗ i ΓiT f, fi =Xi∈I ≥Xi∈I ≥ Akfk − Rkfk2 = (A − R)kfk2 > 0. i ΛiT f, fi +Xi∈I i ΛiT f, fi −Xi∈I hU ∗(Γ∗ hU ∗(Γ∗ i Γi − Λ∗ i Γi − Λ∗ i Λi)T f, fi i Λi)T f, fi So we have the result. (cid:3) Proposition 2.6. Let T, U ∈ GL(H ) and {Λi : i ∈ I} be a (T, U )-controlled g-frame for H . Let {Γi : i ∈ I} be a g-complete family of bounded operator. Suppose that Φ : H −→ H defined by Φ(f ) =Xi∈I U ∗(Γ∗ i Γi − Λ∗ i Λi)T f, ∀f ∈ H , is a positive and compact operator. Then {Γi : i ∈ I} is a (T, U )-controlled g-frame for H . GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 5 Proof. Let {Λi : i ∈ I} be a (T, U )-controlled g-frame for H . By Proposition 2.3, then it is a g-frame for H with bounds A and B. On the other hand, since Φ is a positive compact operator, U −1ΦT −1 is also a positive compact operator. Hence (U ∗)−1ΦT −1f =Xi∈I Γ∗ i Γif − Λ∗ i Λif, ∀f ∈ H . Let Ψ = (U ∗)−1ΦT −1 and Θ : H −→ H be an operator defined by Θ = SΛ + Ψ. A simple computation shows that Ψ is bounded and self-adjoint and Θ is bounded, linear, self- adjoint and for any f ∈ H . Let f be an arbitrary element of H , we have kΘfk = kSΛf + Ψfk ≤ kSΛfk + kΨfk ≤ (B + kΨk)kfk. Γ∗ i Γif, Θf =Xi∈I Therefore, kΓifk2 = hΘf, fi ≤ (B + kΨk)kfk2. Since Ψ is a compact operator, ΨS−1 Λ is also a compact operator on H . By Theorem 2.8 of [1], Θ has closed range. Now we show that Θ is injective. Let g be an element of H such that Θg = 0, then Xi∈I Hence Γig = 0 for each i ∈ I. Since {Γi ∈ L(H ,H i) : i ∈ I} is g-complete, we have g = 0. Furthermore, we have kΓigk2 = hΘg, gi = 0. Xi∈I Hence Θ is onto and therefore invertible on H . Similar to the proof of Theorem 2.8 of [1], we have Range(Θ) = (N (Θ∗))⊥ = N (Θ)⊥ = H . Xi∈I kΓigk2 ≥ (B + kΨk)−1kΘ−1k−2kfk2. Then {Γi : i ∈ I} is a g-frame for H . Since Φ = U ∗SΓT − U ∗SΛT , U ∗SΓT = Φ + U ∗SΛT . It is easy to see that U ∗SΓT is a bounded positive operator. Hence, we have {Γi : i ∈ I} is a (T, U )-controlled g-frame for H . (cid:3) Our next result is a generalization of Theorem 3.3 of [6]. Theorem 2.7. Let T, U ∈ GL(H ) and {Λi ∈ L(H ,H i) : i ∈ I} be a family of bounded operators. Let {Γij ∈ L(H i,H ij ) : j ∈ Ji} be a (T, U )-controlled g-frame for each H i with bounds Ci and Di, which 0 < C ≤ Ci ≤ Di ≤ D < ∞. Then the following conditions are equivalent. (1) {Λi ∈ L(H ,H i) : i ∈ I} is a (T, U )-controlled g-frame for H . (2) {ΓijΛi ∈ L(H i,H ij ) : i ∈ I, j ∈ Ji} is a (T, U )-controlled g-frame for H . Proof. 1→2. Let {Λi ∈ L(H ,H i) : i ∈ I} be a (T, U )-controlled g-frame with bounds (A, B) for H . Then for all f ∈ H we have Xi∈I Xj∈Ji hΓij ΛiT f, ΓijΛiU fi =Xi∈I Xj∈Ji(cid:10)Γ∗ ij ΓijΛiT f, ΛiU f(cid:11) 6 DONGWEI LI AND JINSONG LENG Di hΛiT f, ΛiU fi ≤ DBkfk2. ≤Xi∈I Also we have Xi∈I Xj∈Ji hΓij ΛiT f, ΓijΛiU fi =Xi∈I Xj∈Ji(cid:10)Γ∗ ij ΓijΛiT f, ΛiU f(cid:11) Ci hΛiT f, ΛiU fi ≥ CAkfk2. ≥Xi∈I 2→1. Let {ΓijΛi ∈ L(H i,H ij ) : i ∈ I, j ∈ Ji} be a (T, U )-controlled g-frame with bounds A, B for H . Since Λif ∈ H i, we have Also Xi∈I hΛiT f, ΛiU fi ≤Xi∈I Xi∈I hΛiT f, ΛiU fi ≥Xi∈I 1 Ci Xj∈Ji 1 Di Xj∈Ji hΓij ΛiT f, ΓijΛiU fi ≤ B C kfk2. hΓijΛiT f, ΓijΛiU fi ≥ A Dkfk2. (cid:3) The next result is a characterization for (T, U )-controlled g-frames. Theorem 2.8. Let T, U ∈ GL(H ) and {Λi ∈ L(H ,H i) : i ∈ I} be a family of bounded operators. Suppose that {eij : j ∈ Ji} is an orthonormal basis for H i for each i ∈ I. Then {Λi : i ∈ I} is a (T, U )-controlled g-frame for H if and only if {T ∗uij : i ∈ I, j ∈ Ji} is a U ∗(T ∗)−1-controlled frame for H , where uij = Λ∗ Proof. Let {eij : j ∈ Ji} be an orthonormal basis for H i for each i ∈ I. For any f ∈ H , since Λif ∈ H i, we have i eij . Λi(T f ) = Xj∈Ji Λi(U f ) = Xj∈Ji hΛi(T f ), eiji eij = Xj∈Ji hΛi(U f ), eiji eij = Xj∈Ji hf, T ∗Λ∗ i eiji eij. hf, U ∗Λ∗ i eiji eij. Also, Hence, hΛiT f, ΛiU fi = Xj∈Ji hf, T ∗Λ∗ i eijihU ∗Λ∗eij, fi . i eij, fij = T ∗uij and Ω = U ∗(T ∗)−1, then hΛiT f, ΛiU fi ≤ Bkfk2 Now, if we take uij = Λ∗ is equivalent to So we have the result. Akfk2 ≤Xi∈I Akfk ≤Xi∈I Xj∈Ji hf, fijihΩfij , fi ≤ Bkfk2. (cid:3) GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 7 Note that {uij : i ∈ I, j ∈ Ji} is the sequence induced by {Λi : i ∈ I} with respect to {eij : j ∈ Ji}. By the above result, finding suitable operator T and U such that {Λi : i ∈ I} forms a (T, U )- controlled fusion frame for H with optimal bounds, is equivalent to finding suitable operators T and U such that {T ∗uij : i ∈ I, j ∈ Ji} is a U ∗(T ∗)−1-controlled frame for H with optimal frame bounds. Let H and K be two Hilbert spaces. We recall that H ⊕ K = {(f, g) : f ∈ H , g ∈ K }, is a Hilbert space with pointwise operations and inner product h(f, g), (f ′, g′)i := hf, f ′iH + hg, g′iK , ∀f, f ′ ∈ H , g, g′ ∈ K . Also if Λ ∈ L(H , V ) and Γ ∈ L(K , W ), then for all f ∈ H , g ∈ K we define Λ ⊕ Γ ∈ L(H ⊕ K , V ⊕ W ), by (Λ ⊕ Γ)(T f, U g) := (ΛT f, ΓU g), where V, W are Hilbert spaces and T ∈ GL(H ), U ∈ GL(K ). Theorem 2.9. Let T ∈ GL(H ), U ∈ GL(H ). Let {Λi ∈ L(H , Vi) : i ∈ I} and {Γi ∈ L(K , Wi) : i ∈ I} be (T, T )-controlled g-frame with bounds (A, B) and (U, U )-controlled g-frame with bounds (C, D), respectively. Then {Λi ⊕ Γi ∈ L(H ⊕ K , Vi ⊕ Wi) : i ∈ I} is a (T, U )-controlled g-frame with bounds (min{A, C}, max{B, D}). Proof. Let (f, g) be an arbitrary element of H ⊕ K . Then we have Xi∈I h(ΛiT f, ΓiU g), (ΛiT f, ΓiU g)i h(Λi ⊕ Γi)(T f, U g), (Λi ⊕ Γi)(T f, U g)i k(Λi ⊕ Γi)(T f, U g)k2 =Xi∈I =Xi∈I =Xi∈I hΛiT f, Λifi + hΓiU g, ΓiU gi =Xi∈I kΛiT fk2 +Xi∈I ≤ Bkfk2 + Dkgk2 ≤ max{B, D}(kfk2 + kgk2) = max{B, D}k(f, g)k2. kΓiU fk2 Similarly we have So we have the result. min{A, C}(kfk2 + kgk2) ≤Xi∈I k(Λi ⊕ Γi)(T f, U g)k2. (cid:3) 3. Resolutions of the identity In this section, we will find new resolution of the identity. {Λi ∈ L(H ,H i) : i ∈ I} be a (T, U )-controlled g-frame, then we have In fact, let T, U ∈ GL(H ) and U ∗Λ∗ i ΛiT S−1 T ΛU f, ∀f ∈ H . f =Xi∈I S−1 T ΛU U ∗Λ∗ i ΛiT f =Xi∈I By choosing suitable controlled operators we may obtain more suitable approximations. Now we will give a new resolution of the identity by using two controlled operators. 8 DONGWEI LI AND JINSONG LENG Definition 3.1. Let T, U ∈ GL(H ) and let {Λi ∈ L(H ,H i) : i ∈ I} and {Γi ∈ L(H ,H i) : i ∈ I} be (T, T )-controlled and (U, U )-controlled g-Bessel sequence, respectively. We define a (T, U )- controlled g-frame operator for this pair of controlled g-Bessel sequence as follows: ST ΓΛU (f ) =Xi∈I U ∗Γ∗ i ΛiT (f ), ∀f ∈ H . As mentioned before, {Λi ∈ L(H ,H i) : i ∈ I} and {Γi ∈ L(H ,H i) : i ∈ I} are also two g-Bessel sequence. So by [10] the g-frame operator SΓΛ(f ) =Pi∈I Γ∗ i Λi(f ) for this pair of g-Bessel sequence is well defined and bounded. Since ST ΓΛU = U ∗SΓΛT , ST ΓΛU is a well defined and bounded operator. Lemma 3.2. Let T, U ∈ GL(H ) and let {Λi : i ∈ I} and {Γi : i ∈ I} be (T, T )-controlled and (U, U )-controlled g-Bessel sequence with bounds BT and BU , respectively. If ST ΓΛU is bounded below, then {Λi : i ∈ I} and {Γi : i ∈ I} are (T, T )-controlled and (U, U )-controlled g-frames, respectively. Proof. Suppose that there exists a number λ > 0 such that for all f ∈ H λkfk ≤ kST ΓΛUk, then we have λkfk ≤ kST ΓΛUk = sup U ∗Γ∗ i ΛiT f, g >(cid:12)(cid:12) = sup g∈H ,kgk=1(cid:12)(cid:12) <Xi∈I kgk=1(cid:12)(cid:12) <Xi∈I kgk=1(cid:0)Xi∈I ≤ sup ≤pBU(cid:0)Xi∈I ΛiT f, ΓiU g >(cid:12)(cid:12) kΛiT fk2(cid:1)1/2(cid:0)Xi∈I kΛiT fk2(cid:1)1/2 . kΓiU gk2(cid:1)1/2 Hence, On the other hand, since λ2 BU kfk2 ≤Xi∈I kΛiT fk2. S∗ T ΓΛU = (U ∗SΓΛT )∗ = T ∗S∗ ΓΛU = T ∗SΛΓU = SU ΛΓT , we can say that SU ΛΓT is also bounded below. So by the above result {Γi : i ∈ I} is a (U, U )- controlled g-frame. (cid:3) Theorem 3.3. Let T ∈ GL(H ) and let Λ = {Λi ∈ L(H ,H i) : i ∈ I} be a (T, T )-controlled g-Bessel sequence. Then the following conditions are equivalent. (1) Λ is a (T, T )-controlled g-frame for H . (2) There exists an operator U ∈ GL(H ) and a (U, U )-controlled g-Bessel sequence Γ = {Γi ∈ L(H ,H i) : i ∈ I} such that SU ΓΛT ≥ mIH on H , for some m > 0. GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 9 Proof. 1→2. Let Λ be a (T, T )-controlled g-frame with lower and upper g-frame bounds AT and BT , respectively. Then we take U = T , Γi = Λi, for all i ∈ I. Hence we have hST ΛΛT f, fi =*Xi∈I T ∗Λ∗ i ΛiT f, f+ =Xi∈I hΛiT f, ΛiT fi ≥ ATkfk2 for all f ∈ H . Moreover, ATkfk2 ≤ kS T ΛΛTk2 ≤ BTkfk2. 1/2 By Lemma 2.2, ST ΛΛT ∈ GL+(H ). 2→1. Suppose that there exists an operator U ∈ GL(H ) and a (U, U )-controlled g-Bessel sequence Γ = {Γi ∈ L(H ,H i) : i ∈ I} with Bessel bound BU . Also let m > 0 be a constant such that for all f ∈ H . Then we have hSU ΓΛT f, fi ≥ mkfk2 mkfk2 ≤ hSU ΓΛT f, fi =Xi∈I hΛiT f, ΓiU fi kΓiU fk2(cid:1)1/2 kΛiT fk2(cid:1)1/2(cid:0)Xi∈I ≤(cid:0)Xi∈I ≤pBUkfk(cid:0)Xi∈I kΛiT fk2(cid:1)1/2 , by the Cauchy-Schwartz inequality. Hence, m2 BU kfk2 ≤Xi∈I kΛiT fk2 ≤ BTkfk2. So Λ is a (T, T )-controlled g-frame for H . Theorem 3.4. Let T, U ∈ GL(H ) and let {Λi ∈ L(H ,H i) : i ∈ I} be a (T, T )-controlled g-frame with bounds (A, B) for H . Let the family {Γi ∈ L(H ,H i) : i ∈ I} be a (U, U )-controlled g-Bessel sequence. Suppose that there exists a number 0 ≤ λ ≤ A such that (cid:3) k(ST ΓΛU − ST ΛT )fk ≤ λkfk, ∀f ∈ H . Then ST ΓΛU is invertible and also {Γi ∈ L(H ,H i) : i ∈ I} is a (U, U )-controlled g-frame for H . Proof. Let f ∈ H be an arbitrary element of H , then we have kST ΓΛU fk = kST ΓΛU f − ST ΛT f + ST ΛT fk ≥ kST ΛT fk − kST ΓΛU f − ST ΛT fk ≥ (A − λ)kfk. So ST ΓΛU is bounded below and therefore one-to-one with closed range. On the other hand, since kSU ΓΛT − ST ΛTk = k(ST ΓΛU − ST ΛT )∗k ≤ λ, by the above result SU ΓΛT is also bounded below (A − λ) and therefore one-to-one with closed range. Hence both ST ΓΛU and SU ΓΛT are invertible. And, (A − λ)kfk ≤ kSU ΓΛTk = sup g∈H ,kgk=1(cid:12)(cid:12) <Xi∈I T ∗Λ∗ i ΓiU f, g >(cid:12)(cid:12) 10 DONGWEI LI AND JINSONG LENG = sup kgk=1(cid:12)(cid:12) <Xi∈I kgk=1(cid:0)Xi ≤ sup √B(cid:0)Xi kfk2 ≤Xi∈I ΓiU f, ΛiT g >(cid:12)(cid:12) kΓiU fk2(cid:1)1/2(cid:0)Xi kΓiU fk2(cid:1)1/2 kΓiU fk2. ≤ . (A − λ)2 B kΛiT gk2(cid:1)1/2 Hence, Therefore, {Γi ∈ L(H ,H i) : i ∈ I} is a (U, U )-controlled g-frame for H . (cid:3) Another version of these cases is as follows. Proposition 3.5. Let Λ and Γ be controlled g-Bessel sequences as mentioned in Definition 3.1. Suppose that there exists 0 < ε < 1 such that kf − ST ΓΛU fk ≤ εkfk, ∀f ∈ H . Then Λ and Γ are (T, T )-controlled and (U, U )-controlled g-frames, respectively. Furthermore, ST ΓΛU is invertible. Proof. Firstly therefore ST ΓΛU is invertible. Secondly, let f ban an arbitrary element of H . Then we have kIH − ST ΓΛUk ≤ ε < 1, kST ΓΛU fk ≥ kfk − kf − ST ΓΛU fk ≥ (1 − ε)kfk. Hence ST ΓΛU is bounded below. By Lemma 3.2, we know that Λ is a (T, T )-controlled g-frame. On the other hand, we have kIH − SU ΛΓT k = k(IH − ST ΓΛU )∗k ≤ ε. Hence similarly we can say that Γ is a (U, U )-controlled g-frame. (cid:3) With the hypotheses of Theorem 3.4 or Proposition 3.5, both ST ΓΛU and SU ΓΛT are invertible. Then the family is a resolution of the identity. Also we have new reconstruction formulas as follows T ΓΛU U ∗Γ∗ {S−1 i ΛiT}i∈I and S−1 T ΓΛU U ∗Γ∗ i ΛiT S−1 Γ∗ T ΓΛU f S−1 U ΛΓT T ∗Λ∗ T ∗Λ∗ i ΓiU S−1 U ΛΓT f. f =Xi∈I f =Xi∈I i ΛiT f =Xi∈I i ΓiU f =Xi∈I Suppose that kIH − ST ΓΛUk < 1, then as we mentioned in Proposition 3.5, ST ΓΛU is invertible and we have ∞ S−1 T ΓΛU = (IH − ST ΓΛU )n. Xn=0 GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 11 ∞ Xn=0 (IH − ST ΓΛU )nU ∗Γ∗ i ΛiT f =Xi∈I ∞ Xn=0 U ∗Γ∗ i ΛiT (IH − ST ΓΛU )nf. Then we have f =Xi∈I Furthermore, Therefore, kS−1 {(IH − ST ΓΛU )nU ∗Γ∗ T ΓΛUk ≤ (1 − kIH − ST ΓΛUk)−1. i ΛiT}i∈I,n∈Z + is a new resolution of the identity. 4. Stability under perturbations Perturbation of frames is an important and useful objects to construct new frames from a given one. In this section we give new definitions of perturbations of g-frames with respect to the operators T, U . Definition 4.1. Let T, U ∈ GL(H ) and let {Λi ∈ L(H ,H i) : i ∈ I} and {Γi ∈ L(H ,H i) : i ∈ I} be two g-complete family of bounded operator. Let 0 ≤ λ1, λ2 < 1 be real numbers and let C = {ci}i∈I be an arbitrary sequence of positive numbers such that kC k2 < ∞. We say that the family {Γi ∈ L(H ,H i) : i ∈ I} is a (λ1, λ2,C , T, U )-perturbation of {Λi ∈ L(H ,H i) : i ∈ I} if we have kΛiT f − ΓiU fk ≤ λ1kΛiT fk + λ2kΓiU fk + cikfk, ∀f ∈ H . We have the following important result. Theorem 4.2. Let {Λi ∈ L(H ,H i) : i ∈ I} be a g-frame for H with frame bounds A, B. Suppose that T, U ∈ GL(H ). Let {Γi ∈ L(H ,H i) : i ∈ I} be a (λ1, λ2,C , T, U )-perturbation of {Λi ∈ L(H ,H i) : i ∈ I}, in which (1 − λ1)√AkT −1k−1 > kC k2. Then {Γi ∈ L(H ,H i) : i ∈ I} is a g-frame for H with g-frame bounds (1 − λ1)√AkT −1k−1 − kC k2 1 + λ2 kUk−1!2 , (1 + λ1)√BkTk + kC k2 1 − λ2 kUk−1!2 Proof. Since {Λi ∈ L(H ,H i) : i ∈ I} be a g-frame for H with frame bounds A, B, for all f ∈ H , we have Then by triangular inequality we have 1 2 kΓiU fk2! Xi∈I 1 2 ≤ √BkTkfk. √A kT −1kkfk ≤Xi∈I (cid:0)kΛiT fk2(cid:1) ≤ Xi∈I ≤ Xi∈I ≤ (1 + λ1)Xi∈I (cid:0)kΛiT fk2(cid:1) 1 1 2 (kΛiT fk + kΛiT f − ΓiU fk)2! (kΛiT fk + λ1kΛiT fk + λ2kΓiU fk + cikfk)2! 1 2 2 + λ2Xi∈I (cid:0)kΓiU fk2(cid:1) 1 2 + kC k2kfk. DONGWEI LI AND JINSONG LENG 1 2 ≤ (1 + λ1)√BkTk kU fk kΓifk2 ≤ (1 + λ1)√BkTk + kC k2) 1 − λ2 12 Hence (1 − λ2)Xi∈I (cid:0)kΓiU fk2(cid:1) Since U f ∈ H , finally we have Now for the lower bound we have Xi∈I kΓiU fk2! Xi∈I kUk−1 + kC k2 kU fk kUk−1 . kUk−1!2 kfk2. 1 2 1 2 (kΛiT fk − kΛiT f − ΓiU fk)2! (kΛiT fk − λ1kΛiT fk − λ2kΓiU fk − cikfk)2! ≥ Xi∈I ≥ Xi∈I ≥ (1 − λ1)Xi∈I (cid:0)kΛiT fk2(cid:1) 1 2 − λ2Xi∈I (cid:0)kΓiU fk2(cid:1) 1 2 − kC k2kfk. 1 2 Hence which yields (1 + λ2)Xi∈I (cid:0)kΓiU fk2(cid:1) 1 2 ≥ (1 − λ1)√AkT −1k−1 kU fk kΓifk2 ≥ (1 − λ1)√AkT −1k−1 − kC k2 1 + λ2 Xi∈I kUk−1 − kC k2 kU fk kUk−1 , kUk−1!2 kfk2. Therefore, we get the results. (cid:3) The research is supported by the National Natural Science Foundation of China (Nos. 11271001 and 61370147). Acknowledgements References 1. MS Asgari and Amir Khosravi, Frames and bases of subspaces in Hilbert spaces, Journal of mathematical analysis and applications 308 (2005), no. 2, 541–553. 2. Peter Balazs, Jean-Pierre Antoine, and ANNA GRYBO´S, Weighted and controlled frames: Mutual relationship and first numerical properties, International journal of wavelets, multiresolution and information processing 8 (2010), no. 01, 109–132. 3. Peter Balazs, Dominik Bayer, and Asghar Rahimi, Multipliers for continuous frames in Hilbert spaces, Journal of Physics A: Mathematical and Theoretical 45 (2012), no. 24, 244023. 4. Bernhard G Bodmann and Vern I Paulsen, Frames, graphs and erasures, Linear Algebra and its Applications 404 (2005), 118–146. 5. Iva Bogdanova, Pierre Vandergheynst, Jean-Pierre Antoine, Laurent Jacques, and Marcella Morvidone, Stereo- graphic wavelet frames on the sphere, Applied and Computational Harmonic Analysis 19 (2005), no. 2, 223–252. 6. Peter G Casazza, Gitta Kutyniok, and Shidong Li, Fusion frames and distributed processing, Applied and computational harmonic analysis 25 (2008), no. 1, 114–132. 7. Ingrid Daubechies, A. Grossmann, and Y. Meyer, Painless nonorthogonal expansions, Journal of Mathematical Physics 27 (1986), no. 5. 8. Richard J Duffin and Albert C Schaeffer, A class of nonharmonic Fourier series, Transactions of the American Mathematical Society 72 (1952), no. 2, 341–366. GENERALIZED FRAMES AND CONTROLLED OPERATORS IN HILBERT SPACE 13 9. Xunxiang Guo, Characterizations of disjointness of g-frames and constructions of g-frames in Hilbert spaces, Complex Analysis and Operator Theory 8 (2014), no. 7, 1547–1563. 10. Amir Khosravi and Kamran Musazadeh, Fusion frames and g-frames, Journal of Mathematical Analysis and Applications 342 (2008), no. 2, 1068–1083. 11. Jelena Kovacevi´c and Amina Chebira, An introduction to frames, Foundations and Trends in Signal Processing 2 (2008), no. 1, 1–94. 12. Jinsong Leng and Deguang Han, Optimal dual frames for erasures II, Linear Algebra and its Applications 435 (2011), no. 6, 1464–1472. 13. Jinsong Leng, Deguang Han, and Tingzhu Huang, Optimal dual frames for communication coding with proba- bilistic erasures, Signal Processing, IEEE Transactions on 59 (2011), no. 11, 5380–5389. 14. Patricia Mariela Morillas, Optimal dual fusion frames for probabilistic erasures, Electronic Journal of Linear Algebra 32 (2017), 191–203. 15. Kamran Musazadeh and Hassan Khandani, Some results on controlled frames in Hilbert spaces, Acta Mathe- matica Scientia 36 (2016), no. 3, 655–665. 16. Asghar Rahimi and Abolhassan Fereydooni, Controlled g-frames and their g-multipliers in Hilbert spaces, Analele Universitatii" Ovidius" Constanta-Seria Matematica 21 (2013), no. 2, 223–236. 17. Qiyu Sun and Wai Shing Tang, Nonlinear frames and sparse reconstructions in Banach spaces, Journal of Fourier Analysis and Applications (2015), 1–35. 18. Wenchang Sun, G-frames and g-Riesz bases, Journal of Mathematical Analysis and Applications 322 (2006), no. 1, 437–452. 19. , Stability of g-frames, Journal of mathematical analysis and applications 326 (2007), no. 2, 858–868. School of Mathematical Sciences, University of Electronic Science and Technology of China, 611731, P. R. China E-mail address: [email protected] School of Mathematical Sciences, University of Electronic Science and Technology of China, 611731, P. R. China E-mail address: [email protected]
1103.1409
1
1103
2011-03-08T01:05:30
Maximally Monotone Linear Subspace Extensions of Monotone Subspaces: Explicit Constructions and Characterizations
[ "math.FA" ]
Monotone linear relations play important roles in variational inequality problems and quadratic optimizations. In this paper, we give explicit maximally monotone linear subspace extensions of a monotone linear relation in finite dimensional spaces. Examples are provided to illustrate our extensions. Our results generalize a recent result by Crouzeix and Anaya.
math.FA
math
Maximally Monotone Linear Subspace Extensions of Monotone Subspaces: Explicit Constructions and Characterizations Xianfu Wang ∗ and Liangjin Yao † Dedicated to Jonathan Borwein on the occasion of his 60th birthday. 1 1 0 2 r a M 8 ] . A F h t a m [ March 6, 2011 Abstract 1 v 9 0 4 1 . 3 0 1 1 : v i X r a Monotone linear relations play important roles in variational inequality problems and quadratic optimizations. In this paper, we give explicit maximally monotone linear subspace extensions of a monotone linear relation in finite dimensional spaces. Exam- ples are provided to illustrate our extensions. Our results generalize a recent result by Crouzeix and Anaya. 2000 Mathematics Subject Classification: Primary 47H05; Secondary 47B65, 47A06, 49N15 Key words and phrases: Adjoint of linear relation, linear relation, monotone operator, maximally monotone extensions, Minty parametrization. 1 Introduction Throughout this paper, we assume that Rn ( n ∈ N = {1, 2, 3, . . .}) is an Euclidean space with the inner product h·,·i, and induced Euclidean norm k · k. Let G : Rn ⇒ Rn be a ∗Mathematics, Irving K. Barber School, The University of British Columbia Okanagan, Kelowna, B.C. V1V 1V7, Canada. Email: [email protected]. †Mathematics, Irving K. Barber School, The University of British Columbia Okanagan, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected]. 1 set-valued operator from Rn to Rn, i.e., for every x ∈ Rn, Gx ⊆ Rn, and let gra G = (cid:8)(x, x∗) ∈ Rn × Rn x∗ ∈ Gx(cid:9) be the graph of G. Recall that G is monotone if (1) (cid:0)∀(x, x∗) ∈ gra G(cid:1)(cid:0)∀(y, y∗) ∈ gra G(cid:1) hx − y, x∗ − y∗i ≥ 0, and maximally monotone if G is monotone and G has no proper monotone extension (in the sense of graph inclusion). We say that G is a linear relation if gra G is a linear subspace. While linear relations have been extensively studies [9, 18, 6, 2, 3, 17], monotone operators are ubiquitous in convex optimization and variational analysis [1, 15, 6, 7]. The central object of this paper is to consider the linear relation G : Rn ⇒ Rn: (i) gra G = {(x, x∗) Ax + Bx∗ = 0}. (ii) A, B ∈ Rp×n. (iii) rank(A B) = p. Our main concern is to find explicit maximally monotone linear subspace extensions of G. Recently, finding constructive maximal monotone extensions instead of using Zorn's lemma has been a very active topic [5, 4, 12, 11, 10]. In [10], Crouzeix and Anaya gave an algorithm to find maximally monotone linear subspace extensions of G, but it is not clear what the maximally monotone extensions are analytically. In this paper, we provide some maximally monotone extensions of G with closed analytical forms. Along the way, we also give a new proof to Crouzeix and Anaya's characterizations on monotonicity and maximal monotonicity of G. Our key tool is the Brezis-Browder characterization of maximally monotone linear relations. The paper is organized as follows. In the remainder of this introductory section, we describe some central notions fundamental to our analysis. In Section 2, we collect some auxiliary results for future reference and for the reader's convenience. Section 3 provides explicit self-dual maximal monotone extensions by using subspaces on which AB⊺ + BA⊺ is negative semidefinite, and obtain a complete characterization of all maximal monotone extensions. Section 4 deals with Minty's parameterizations of monotone operator G. In Section 5, we get some explicit maximally monotone extensions with the same domain or the same range by utilizing normal cone operators. In Section 6, we illustrate our maximally monotone extensions by considering three examples. Our notations are standard. We use dom G =(cid:8)x ∈ Rn Gx 6= ∅(cid:9) for the domain of G, ran G = G(Rn) for the range of G and ker G = (cid:8)x ∈ Rn 0 ∈ Gx(cid:9) for the kernal of G. Given a subset C of Rn, span C is the span (the set of all finite linear combinations) of C. We set C⊥ = {x∗ ∈ Rn (∀c ∈ C)hx∗, ci = 0}. 2 Then the adjoint of G, denoted by G∗, is defined by gra G∗ = {(x, x∗) ∈ Rn × Rn (x∗,−x) ∈ (gra G)⊥}. The set Rn×p is the set of all the n×p matrices, for n, p ∈ N. Then rank(M) is the rank of the matrix M ∈ Rn×p. Let Id : Rn → Rn denote the identity mapping, i.e., Id x = x for x ∈ Rn. We also set PX : Rn × Rn → Rn : (x, x∗) 7→ x, and PX ∗ : Rn × Rn → Rn : (x, x∗) 7→ x∗. If X ,Y are subspaces of Rn, we let X + Y =(cid:8)x + y x ∈ X , y ∈ Y(cid:9). Counting multiplicities, let (2) (3) λ1, λ2,· · · , λk be all positive eigenvalues of (AB⊺ + BA⊺) and λk+1, λk+2,· · · , λp be nonpositive eigenvalues of (AB⊺ + BA⊺). Moreover, let vi be an eigenvector of eigenvalue λi of (AB⊺ + BA⊺) satisfying kvik = 1, and hvi, vji = 0 for 1 ≤ i 6= j ≤ q. It will be convenient to put 0 0 ... 0 λp Idλ = diag(λ1,· · · , λp) = V = [v1 v2 · · · vp] . 0 λ1 0 λ2 · · · · · · . . . 0 0 0 0 0 (4)  0 ... 0 0 λ3 0 0 ,  2 Auxiliary results on linear relations In this section, we collect some facts and preliminary results which will be used in sequel. We first provide a result about subspaces on which a linear operator from Rn → Rn, i.e, an n × n matrix, is monotone. For M ∈ Rn×n, define three subspaces of Rn, namely, the positive eigenspace, null eigenspace and negative eigenspace associated with M + M ⊺ by V+(M) = span(w1,· · · , ws : wi is an eigenvector of positive eigenvalue αi of M + M ⊺ hwi, wji = 0, ∀ i 6= j,kwik = 1, i, j = s + 1, . . . , l.) V0(M) = span(ws+1,· · · , wl : wi is an eigenvector of 0 eigenvalue of M + M ⊺ hwi, wji = 0 ∀ i 6= j,kwik = 1, i, j = 1, . . . , s. ) 3 V−(M) = span(wl+1,· · · , wn : wi is an eigenvector of negative eigenvalue αi of M + M ⊺ hwi, wji = 0 ∀ i 6= j,kwik = 1, i, j = l + 1, . . . , n. ) which is possible since a symmetric matrix always has a complete orthonormal set of eigen- vectors, [14, pages 547 -- 549]. Proposition 2.1 Let M be an n × n matrix. Then (i) M is strictly monotone on V+(M). Moreover, M + M ⊺ : V+(M) → V+(M) is a bijection. (ii) M is monotone on V+(M) + V0(M). (iii) −M is strictly monotone on V−(M). Moreover, −(M + M ⊺) : V−(M) → V−(M) is a bijection. (iv) −M is monotone on V−(M) + V0(M). (v) For every x ∈ V0(M), (M + M ⊺)x = 0 and hx, M xi = 0. In particular, the orthogonal decomposition holds: Rn = V+(M) ⊕ V0(M) ⊕ V−(M). Proof. {w1,· · · , ws} is a set of orthonormal vectors, they are linearly independent so that (i): Let x ∈ V+(M). Then x = Ps i=1 liwi for some (l1, . . . , ls) ∈ Rs. Since Note that αi > 0 when i = 1, . . . , s and hwi, wji = 0 for i 6= j. We have x 6= 0 ⇔ (l1,· · · , ls) 6= 0. 2hx, M xi = hx, (M + M ⊺)xi = h liwi, (M + M ⊺)( sXi=1 sXi=1 liαiwii = αil2 i > 0 sXi=1 liwi)i i=1 liwi, we have li(M + M ⊺)wi = sXi=1 αiliwi ∈ V+(M). if x 6= 0. = h sXi=1 liwi, sXi=1 For every x ∈ V+(M) with x =Ps sXi=1 (M + M ⊺)x = As αi > 0 for i = 1, . . . , s and {w1, . . . , ws} is an orthonormal basis of V+(M), we conclude that M + M ⊺ : V+(M) → V+(M) is a bijection. 4 αi ≥ 0 when i = 1, . . . , l and hwi, wji = 0 for i 6= j. We have (ii): Let x ∈ V+(M) + V0(M). Then x =Pl lXi=1 lXi=1 2hx, M xi = hx, (M + M ⊺)xi = h liαiwii = = h lXi=1 liwi, lXi=1 i=1 liwi for some (l1, . . . , ll) ∈ Rl. Note that liwi, (M + M ⊺)( lXi=1 liwi)i αil2 i ≥ 0. The proofs for (iii), (iv) are similar as (i), (ii). (v): For x ∈ V0(M), 2hx, M xi = hx, (M + M ⊺)xi = hx, 0i = 0. Corollary 2.2 Then following holds. is strictly monotone. gra T = {(B⊺u, A⊺u) u ∈ V+(BA⊺)} gra T = {(B⊺u, A⊺u) u ∈ V+(BA⊺) + V0(BA⊺)} is monotone. is strictly monotone. gra T = {(B⊺u,−A⊺u) u ∈ V−(BA⊺)} (i) (ii) (iii) (iv) gra T = {(B⊺u,−A⊺u) u ∈ V−(BA⊺) + V0(BA⊺))} is monotone. Proof. It follows from Proposition 2.1 and hB⊺u, A⊺ui = hu, BA⊺ui, ∀u ∈ Rn. Lemma 2.3 For every subspace S ⊆ Rp, the following hold. (5) dim{(B⊺u, A⊺u) u ∈ S} = dim S. (6) dim{(B⊺u,−A⊺u) u ∈ S} = dim S. 5 (cid:4) (cid:4) Proof. Let dim S = t and {u1, . . . , ut} be a basis of S. We claim that the set of vectors A⊺(cid:19) ui i = 1, . . . , t(cid:27) (cid:26)(cid:18)B⊺ is linearly independent. Indeed, because (A B) has full row rank p, tXi=1 A⊺(cid:19) ui =(cid:18)B⊺ li(cid:18)B⊺ A⊺(cid:19) tXi=1 Note that and is invertible, we have liui = 0 ⇔ li = 0 for i = 1, . . . , t. liui = 0 ⇔ tXi=1 A⊺u(cid:19) 0 − Id(cid:19)(cid:18)B⊺u −A⊺u(cid:19) =(cid:18)Id (cid:18) B⊺u 0 − Id(cid:19) (cid:18)Id 0 0 dim{(B⊺u, A⊺u) u ∈ S} = dim{(B⊺u,−A⊺u) u ∈ S} so (6) follows from (5). Alternatively, see [14, page 208, Exercise 4.49]. Fact 2.4 We have (AB⊺ + BA⊺)V = V Idλ . Proof. Let y = [y1 y2 · · · yp]⊺ ∈ Rp. Then we have pXi=1 (AB⊺ + BA⊺)V y =(cid:0)AB⊺ + BA⊺(cid:1)( yivi) = pXi=1 λiyivi = V Idλ y. (cid:4) (cid:4) Two key criteria concerning maximally monotone linear relations come as follows: Fact 2.5 (See [19, Proposition 4.2.9 ] or [3, Proposition 2.10].) Let T : Rn ⇒ Rn be a monotone linear relation. The following are equivalent: (i) T is maximally monotone. (ii) dim gra T = n. (iii) dom T = (T 0)⊥. 6 Fact 2.6 (Br´ezis-Browder) (See [8, Theorem 2], or [18] or [16].) Let T : Rn ⇒ Rn be a monotone linear relation. Then the following statements are equivalent. (i) T is maximally monotone. (ii) T ∗ is maximal monotone. (iii) T ∗ is monotone. Some basic properties of G are: Lemma 2.7 (i) gra G = ker(A B). (ii) G0 = ker B, G−1(0) = ker A. (iii) dom G = PX (ker(A B)) and ran G = PX ∗(ker(A B)). (iv) ran(G + Id) = PX ∗(ker(A − B B)) = PX(ker(A B − A)), and dom G = PX(ker(A − B B)), ran G = PX ∗(ker(A (B − A)). (v) dim G = 2n − p. Proof. (i), (ii), (iii) follow from definition of G. Since Ax + Bx∗ = 0 ⇔ (A − B)x + B(x + x∗) = 0 ⇔ A(x + x∗) + (B − A)x∗ = 0, (iv) holds. (v): We have 2n = dim ker(A B) + dim ran(cid:18)A⊺ B⊺(cid:19) = dim G + p. Hence dim G = 2n − p. The following result summarizes the monotonicities of G∗ and G. (cid:4) Lemma 2.8 The following hold. (i) gra G∗ = {(B⊺u,−A⊺u) u ∈ Rp}. (ii) G∗ is monotone ⇔ A⊺B + B⊺A is negative-semidefinite. 7 (iii) Assume G is monotone. Then n ≤ p. Moreover, G is maximally monotone if and only dim G = n = p. Proof. (i): By Lemma 2.7(i), we have (x, x∗) ∈ gra G∗ ⇔ (x∗,−x) ∈ gra G⊥ = ran(cid:18)A⊺ B⊺(cid:19) = {(A⊺u, B⊺u) u ∈ Rp}. Thus gra G∗ = {(B⊺u,−A⊺u) u ∈ Rp}. (ii): Since gra G∗ is a linear subspace, by (i), G∗ is monotone ⇔ hB⊺u,−A⊺ui ≥ 0, ⇔ hu, BA⊺ui ≤ 0, ⇔ (A⊺B + B⊺A) is negative semidefinite. ∀u ∈ Rp ⇔ hu, (A⊺B + B⊺A)ui ≤ 0, ∀u ∈ Rp ⇔ hu,−BA⊺ui ≥ 0, ∀u ∈ Rp ∀u ∈ Rp (iii): By Fact 2.5 and Lemma 2.7(v), 2n − p = dim gra G ≤ n ⇒ n ≤ p. By Fact 2.5 and Lemma 2.7(v) again, G is maximally monotone ⇔ 2n − p = dim gra G = n ⇔ dim gra G = p = n. (cid:4) 3 Explicit maximal monotone extensions of monotone linear relations In this section, we give explicit maximal monotone linear subspace extensions of G by using V+(AB⊺) or Vg. A characterization of all the maximally monotone extensions of G is also given. Proof. Let (y, y∗) ∈ Rn × Rn. Then we have graeG = {(x, x∗) N ⊺V ⊺Ax + N ⊺V ⊺Bx∗ = 0} grabG = {(B⊺u,−A⊺u) u ∈ ran V N)}. Lemma 3.1 Let N ∈ Rp×p and eG and bG be defined by Then (eG)∗ = bG. (y, y∗) ∈ gra(eG)∗ ⇔ (y∗,−y) ∈ (graeG)⊥ = (ker(cid:0)N ⊺V ⊺A N ⊺V ⊺B(cid:1)(cid:1)⊥ = ran(cid:18)A⊺V N B⊺V N(cid:19) Hence (eG)∗ = bG. ⇔ (y, y∗) ∈ grabG. 8 (cid:4) Lemma 3.2 Define eG and bG by where Vg is (p − k) × p matrix defined by graeG = {(x, x∗) VgAx + VgBx∗ = 0} grabG = {(B⊺u,−A⊺u) u ∈ V−(BA⊺) + V0(BA⊺)},  Vg = v⊺ k+1 v⊺ k+2 ... v⊺ p . Then (i) bG is monotone. (ii) (bG)∗ = eG. (iii) graeG = gra G +(cid:26)(cid:18)B⊺ Proof. (i): Apply Corollary 2.2(iv). (ii): Notations are as in (4). Let A⊺(cid:19) u u ∈ V+(BA⊺)(cid:27) . N = [0 0 · · · 0 ek+1 · · · ep] , (7) where ei = [0, 0 · · · 1, 0 · · · 0]⊺: the ith entry is 1 and the others are 0. =(cid:18) 0 Vg(cid:19) . N ⊺V ⊺ =(cid:18)(v1 · · · vk V ⊺ Then we have (8) ⇔ N ⊺V ⊺Ax + N ⊺V ⊺Bx∗ = 0, ∀(x, x∗) ∈ Rn × Rn. Hence graeG = {(x, x∗) N ⊺V ⊺Ax + N ⊺V ⊺Bx∗ = 0}. 9 Then we have VgAx + VgBx∗ = 0 ⇔(cid:18) 0 Id(cid:19)(cid:19)⊺ g )(cid:18)0 0 VgAx + VgBx∗(cid:19) = 0 0 Thus by Lemma 3.1, gra(eG)∗ = {(B⊺u,−A⊺u) u ∈ ran V N = ran(cid:0)0 V ⊺ Hence (bG)∗ = (eG)∗∗ = eG. (iii): Let J be defined by g(cid:1) = V−(BA⊺) + V0(BA⊺)} = grabG. gra J = gra G +(cid:26)(cid:18)B⊺ A⊺(cid:19) u u ∈ V+(BA⊺)(cid:27) . Then we have By Lemma 2.7(i), Then if and only if . (gra J)⊥ = (gra G)⊥ ∩(cid:26)(cid:18)B⊺ gra G⊥ =(cid:26)(cid:18)A⊺ B⊺(cid:19) w ∈(cid:26)(cid:18)B⊺ (cid:18)A⊺ A⊺(cid:19) u u ∈ V+(BA⊺)(cid:27)⊥ B⊺(cid:19) w w ∈ Rp(cid:27) A⊺(cid:19) u u ∈ V+(BA⊺)(cid:27)⊥ h(A⊺w, B⊺w), (B⊺u, A⊺u)i = 0 ∀ u ∈ V+(BA⊺), that is, (9) hA⊺w, B⊺ui + hB⊺w, A⊺ui = hw, (AB⊺ + BA⊺)ui = 0 ∀u ∈ V+(AB⊺). Because AB⊺ + BA⊺ : V+(AB⊺) 7→ V+(AB⊺) is onto by Proposition 2.1(i), we obtain that (9) holds if and only if w ∈ V−(AB⊺) + V0(AB⊺). Hence (gra J)⊥ = {(A⊺w, B⊺w) w ∈ V−(BA⊺) + V0(BA⊺)}, from which gra J∗ = grabG. Then by (i), graeG = gra(bG)∗ = gra J∗∗ = gra J. (cid:4) We are ready to apply Brezis-Browder Theorem, namely Fact 2.6, to improve Crouzeix- Anaya's characterizations of monotonicity and maximal monotonicity of G and provide a different proof. 10 (i) G is monotone; Theorem 3.3 Let bG,eG be defined in Lemma 3.2. The following are equivalent: (ii) eG is monotone; (iii) eG is maximally monotone; (iv) bG is maximally monotone; (v) dim V+(BA⊺) = p− n, equivalently, AB⊺ + BA⊺ has exactly p− n positive eigenvalues (counting multiplicity). Proof. (i)⇔(ii): Lemma 3.2(iii) and Corollary 2.2(i). Corollary 2.2(iv). It suffices to combine Lemma 3.2 and Fact 2.6. (ii)⇔(iii)⇔(iv): Note that eG = (cid:0)bG(cid:1)∗ and bG is always a monotone linear relation by "(i)⇒(v)": Assume that G is monotone. Then eG is monotone by Lemma 3.2(iii) and Corollary 2.2(i). By Lemma 3.2(ii), Corollary 2.2(iv) and Fact 2.6, bG is maximally monotone, so that dim(grabG) = p − k = n by Fact 2.5 and Lemma 2.3, thus k = p − n. Note that for "(v)⇒(i)": Assume that k = p − n. Then dim(grabG) = p − k = n by Lemma 2.3, so that bG is maximally monotone by Fact 2.5(i)(ii). By Lemma 3.2(ii) and Fact 2.6, eG is monotone, each eigenvalue of a symmetric matrix, its geometric multiplicity is the same as its algebraic multiplicity [14, page 512]. which implies that G is monotone. (cid:4) Corollary 3.4 Assume that G is monotone. Then is a maximally monotone extension of G, where = {(x, x∗) VgAx + VgBx∗ = 0} A⊺(cid:19) u u ∈ V+(BA⊺i(cid:27) graeG = gra G +(cid:26)(cid:18)B⊺  Vg = v⊺ p−n+1 v⊺ p−n+2 ... v⊺ p . Proof. Combine Theorem 3.3 and Lemma 3.2(iii) directly. (cid:4) A remark is in order to compare our extension with the one by Crouzeix and Anaya. 11 Remark 3.5 (i). Crouzeix-Anaya [10] defines the union of monotone extension of G as S = gra G +(cid:26)(cid:18)B⊺ A⊺(cid:19) u u ∈ K(cid:27) , where K = {u ∈ Rn hu, (AB⊺ + BA⊺)ui ≥ 0}. Although this is the set monotonically related to G, it is not monotone in general as long as (AB⊺ + BA⊺) has both positive eigenvalues and negative eigenvalues. Indeed, let (α1, u1) and (α2, u2) be eigen-pairs of (AB⊺ + BA⊺) with α1 > 0 and α2 < 0. We have hu1, (AB⊺ + BA⊺)u1i = α1ku1k2 > 0, hu2, (AB⊺ + BA⊺)u2i = α2ku2k2 < 0. Choose ǫ > 0 sufficiently small so that hu1 + ǫu2, (AB⊺ + BA⊺)(u1 + ǫu2)i > 0. Then However, has A⊺(cid:19) u1,(cid:18)B⊺ (cid:18)B⊺ A⊺(cid:19) (u1 + ǫu2) −(cid:18)B⊺ (cid:18)B⊺ A⊺(cid:19) (u1 + ǫu2) ∈ S. A⊺(cid:19) u1 = ǫ(cid:18)B⊺ A⊺(cid:19) u2 hǫB⊺u2, ǫA⊺u2i = ǫ2hu2, BA⊺u2i = ǫ2hu2, (AB⊺ + BA⊺)u2i 2 < 0. Therefore S is not monotone. By using V+(BA⊺) ⊆ K, we have obtained a maximally monotone extension of G . (ii). Crouzeix and Anaya [10] find the maximal monotone linear subspace extension of G algorithmically by using u ∈ fGk \ Gk. Computationally, it is not completely clear to us how to find such an u. The following result extends the characterization of maximally monotone linear relations given by Crouzeix-Anaya [10]. Theorem 3.6 Let bG,eG be defined in Lemma 3.2. The following are equivalent: (i) G is maximally monotone; (ii) p = n and G is monotone; (iii) p = n and AB⊺ + BA⊺ is negative semidefinite. (iv) p = n and bG is maximally monotone. 12 Proof. (i)⇒(ii): Apply Lemma 2.8(iii). (ii)⇒(iii): Apply directly Theorem 3.3(i)(v). (iii)⇒(i): Assume that p = n and (AB⊺ + BA⊺) is negative semidefinite. Then k = 0 and eG = G. It follows that dim(grabG) = p − k = n by Lemma 2.3, so that bG is maximally monotone by Corollary 2.2(iv) and Fact 2.5(i)(ii). Since(cid:0)bG(cid:1)∗ = eG by Lemma 3.2(ii), Fact 2.6 gives that eG = G is maximally monotone. and dim(grabG) = p−k = n−0 = n. Hence (iv) holds by Corollary 2.2(iv) and Fact 2.5(i)(ii). (iv)⇒(iii): Assume that bG is maximally monotone and p = n. We have dim(grabG) = (iii)⇒(iv): Assume that p = n and (AB⊺ + BA⊺) is negative semidefinite. We have k = 0 p − k = n − k = n so that k = 0. Hence (AB⊺ + BA⊺) is negative semidefinite. We end this section with a characterization of all the maximally monotone linear subspace (cid:4) extensions of G. Theorem 3.7 Let G be monotone. Then eG is a maximally monotone extension of G if and only if there exists N ∈ Rp×p with rank of n such that N ⊺ Idλ N is negative semidefinite and (10) graeG = {(x, x∗) N ⊺V ⊺Ax + N ⊺V ⊺Bx∗ = 0}. Proof. "⇒": By Lemma 2.8(i), we have (11) gra G∗ = {(B⊺u,−A⊺u) u ∈ Rp}. Thus by (11), there exists a subspace F of Rp such that Since gra G ⊆ graeG and thus gra(eG)∗ is a subspace of gra G∗. gra(eG)∗ = {(B⊺u,−A⊺u) u ∈ F}. By Fact 2.6, Fact 2.5 and Lemma 2.3, we have (12) (13) dim F = n. Thus, there exists N ∈ Rp×p with rank n such that ran V N = F and (14) gra(eG)∗ = {(B⊺V N y,−A⊺V N y) y ∈ Rp}. As eG is maximal monotone, (eG)∗ is maximal monotone by Fact 2.6, so N ⊺V ⊺(BA⊺ + AB⊺)V N is negative semidefinite. 13 Using Fact 2.4, we have (15) N ⊺ Idλ N = N ⊺V ⊺V Idλ N = N ⊺V ⊺(AB⊺ + BA⊺)V N which is negative semidefinite. (10) follows from (14) by Lemma 3.1. "⇐": By Lemma 3.1, we have (16) gra(eG)∗ = {(B⊺V N u,−A⊺V N u) u ∈ Rp}. semidefinite by Fact 2.4 and the assumption. As rank(V N) = n, it follows from (16) and Observe that (eG)∗ is monotone because N ⊺V ⊺(AB⊺ + BA⊺)V N = N ⊺ Idλ N is negative Lemma 2.3 that dim gra(eG)∗ = n. Therefore (eG)∗ is maximally monotone by Fact 2.5. Applying Fact 2.6 for T = (eG)∗ yields that eG = (eG)∗∗ is maximally monotone. From the above proof, we see that to find a maximal monotone extension extension of G one essentially need to find subspace F ⊆ Rp such that dim F = n and AB⊺ + BA⊺ is negative semidefinite on F . If F = ran M and M ∈ Rp×p with rank M = n, one can let N = V ⊺M. The maximal monotone linear subspace extension of G is (cid:4) eG = {(x, x∗) M ⊺Ax + M ⊺Bx∗ = 0}. vp−n+1 · · · vp(cid:1). In Corollary 3.4, one can choose M =(cid:0) 0 0 · · · 0 } {z Corollary 3.8 Let G be monotone. Then eG is a maximally monotone extension of G if and only if there exists M ∈ Rp×p with rank of n such that M ⊺(AB⊺ + BA⊺)M is negative semidefinite and n (17) graeG = {(x, x∗) M ⊺Ax + M ⊺Bx∗ = 0}. Note that G may have different representations in terms of A, B. The maximal monotone extension of eG given in Theorem 3.7 and Corollary 3.4 relies on A, B matrices and N. This might leads different maximal monotone extensions, see Section 6. 4 Minty parameterizations Although G is set-valued in general, when G is monotone it has a beautiful Minty parametrization in terms of A, B, which is what we are going to show in this section. 14 Lemma 4.1 The linear relation G is monotone if and only if (18) (19) kyk2 − ky∗k2 ≥ 0, whenever (A + B)y + (B − A)y∗ = 0. Consequently, if G is monotone then the p × n matrix B − A must have full column rank, namely n. Proof. Define It is easy to see that G is monotone if and only if Id P =(cid:18) 0 h(x, x∗), P(cid:18) x Id 0(cid:19) . x∗(cid:19)i ≥ 0, √2(cid:18)Id − Id Id (cid:19) y∗(cid:19) . x∗(cid:19) = Q(cid:18) y (cid:18) x Q = 1 Id whenever Ax + Bx∗ = 0. Define the orthogonal matrix and put Then G is monotone if and only if (20) (21) kyk2 − ky∗k2 ≥ 0, whenever (A + B)y + (B − A)y∗ = 0. If (B − A) does not have full column rank, then there exists y∗ 6= 0 such that (B − A)y∗ = 0. Then (0, y∗) satisfies (21) but (20) fails. Therefore, B − A has to be full column rank. (cid:4) Theorem 4.2 (Minty parametrization) Assume that G is a monotone operator. Then (x, x∗) ∈ gra G if and only if (22) x = (23) x∗ = 1 2 1 2 [Id +(B − A)†(B + A)]y [Id−(B − A)†(B + A)]y for y = x + x∗ ∈ ran(Id +G). Here the Moore-Penrose inverse (B − A)† = [(B − A)⊺(B − A)]−1(B − A)⊺. In particular, when G is maximally monotone, we have gra G = {((B − A)−1By,−(B − A)−1Ay) y ∈ Rn}. 15 Proof. As (B − A) is full column rank, (B − A)⊺(B − A) is invertible. It follows from (19) that (B − A)⊺(A + B)y + (B − A)⊺(B − A)y∗ = 0 so that y∗ = −((B − A)⊺(B − A))−1(B − A)⊺(A + B)y = −(B − A)†(A + B)y. Then x = x∗ = 1 √2 1 √2 (y − y∗) = (y + y∗) = 1 √2 1 √2 [Id +(B − A)†(B + A)]y [Id−(B − A)†(B + A)]y where y = x+x∗ √2 with (x, x∗) ∈ gra G. Since ran(Id +G) is a subspace, we have x = x∗ = 1 2 1 2 [Id +(B − A)†(B + A)]y [Id−(B − A)†(B + A)]y with y = x + x∗ ∈ ran(Id +G). If G is maximally monotone, then p = n by Theorem 3.6 and hence B − A is invertible, thus (B − A)† = (B − A)−1. Moreover, ran(G + Id) = Rn. Then (22) and (23) transpire to (24) x = (25) x∗ = 1 2 1 2 (B − A)−1[B − A + (B + A)]y = (B − A)−1By (B − A)−1[(B − A) − (B + A)]y = −(B − A)−1Ay for y ∈ Rn. Remark 4.3 See Lemma 2.7 for ran(G + Id). Note that as G is a monotone linear relation, the mapping (cid:4) is bijective and linear from ran(G + Id) to gra G, therefore dim(ran(G + Id)) = dim(gra G). z 7→ ((G + Id)−1, Id−(G + Id)−1)(z) Corollary 4.4 Let G be a monotone operator. Then eG define in Corollary 3.4, the maxi- mally monotone extension of G, has its Minty parametrization given by graeG = {((VgB − VgA)−1VgBy,−(VgB − VgA)−1VgAy) y ∈ Rn} where Vg is given as in Corollary 3.4. Proof. Since rank(Vg) = n and rank(A B) = p, by Lemma 2.3(5), rank(VgA VgB) = n. Apply Corollary 3.4 and Theorem 4.2 directly. (cid:4) 16 Corollary 4.5 When G is maximally monotone, dom G = (B − A)−1(ran B), ran G = (B − A)−1(ran A). Recall that T : Rn → Rn is firmly nonexpansive if kT x − T yk2 ≤ hT x − T y, x − yi ∀ x, y ∈ dom T. In terms of matrices Corollary 4.6 Suppose that p = n, AB⊺ + BA⊺ is negative semidefinite. Then (B − A)−1B and −(B − A)−1A are firmly nonexpansive. Proof. By Theorem 3.6, G is maximal monotone. Theorem 4.2 gives that (B − A)−1B = (Id +G)−1, −(B − A)−1A = (Id +G−1)−1. Being resolvent of monotone operators G, G−1, they are firmly nonexpansive, see [1, 13] or [4, Fact 2.5]. (cid:4) 5 Maximally monotone extensions with the same do- main or the same range The purpose of this section is to find maximal monotone linear subspace extensions of G which keep either dom G or ran G unchanged. For a closed convex set S ⊆ Rn, let NS denote its normal cone mapping. Proposition 5.1 Assume that T : Rn ⇒ Rn is a monotone linear relation. Then (i) T1 = T + Ndom T , i.e., x 7→ T1x =(T x + (dom T )⊥ ∅ if x ∈ dom T otherwise is maximally monotone. In particular, dom T1 = dom T . (ii) T2 = (T −1 + Nran T )−1 is a maximally monotone extension of T and ran T2 = ran T . 17 (i): Since 0 ∈ T 0 ⊆ (dom T )⊥ by[2, Proposition 2.2(i)], we have T10 = T 0 + Proof. (dom T )⊥ = (dom T )⊥ so that dom T1 = dom T = (T10)⊥. Hence T1 is maximally mono- tone by Fact 2.5. (ii): Apply (i) to T −1 to see that T −1 + Nran T is a maximally monotone extension of T −1 with dom(T −1 + Nran T ) = ran T . Therefore, T2 is a maximally monotone extension of T with ran T2 = ran T . (cid:4) Since gra G = {(x, x∗) Ax + Bx∗ = 0} we can use Gaussian elimination to reduce (A B) to row echelon form. Then back substitution to solve basic variables in terms of the free variables, see [14, page 61]. Row-echelon form gives where y ∈ R2n−p and with C, D being n × (2n − p) matrices. Therefore, x∗(cid:19) = h1y1 + · · · + h2n−py2n−p =(cid:18)C D(cid:19) y (cid:18) x D(cid:19) = (h1, . . . , h2n−p) (cid:18)C gra G =(cid:26)(cid:18)Cy Dy(cid:19) +(cid:18) Dy(cid:19) +(cid:18)(ran D)⊥ Dy(cid:19) y ∈ R2n−p(cid:27) . (ran C)⊥(cid:19) y ∈ R2n−p(cid:27) . (cid:19) y ∈ R2n−p(cid:27) . gra E1 =(cid:26)(cid:18)Cy gra E2 =(cid:26)(cid:18)Cy 0 0 (26) Define (27) (28) Theorem 5.2 (i) E1 is a maximally monotone extension of G with dom E1 = dom G. Moreover, (29) gra E1 = ran(cid:18)C D(cid:19) +(cid:18) 0 (ran C)⊥(cid:19) = ran(cid:18)C D(cid:19) +(cid:18) 0 ker C ⊺(cid:19) . (ii) E2 is a maximally monotone extension of G with ran E2 = ran G. Moreover, (30) gra E2 = ran(cid:18)C D(cid:19) +(cid:18)(ran D)⊥ 0 (cid:19) = ran(cid:18)C 0 (cid:19) . D(cid:19) +(cid:18)ker D⊺ 18 Proof. (i): Note that dom G = ran C. The maximal monotonicity follows from Proposi- tion 5.1. (29) follows from (27) and that (ran C)⊥ = ker C ⊺ [14, page 405]. (ii): Apply (i) to G−1, i.e., (31) gra G−1 =(cid:26)(cid:18)Dy Cy(cid:19) y ∈ R2n−p(cid:27) and followed by taking the set-valued inverse. (cid:4) Apparently, both extensions E1, E2 rely on gra G, dom G, ran G, not on the A, B. In this sense, E1, E2 are intrinsic maximal monotone linear subspace extensions. Remark 5.3 Theorem 5.2 is much easier to use than Corollary 3.8 when G is written in the form of (26). Indeed, it is not hard to check that (32) (33) gra(E∗1) = {(B⊺u,−A⊺u) B⊺u ∈ dom G, u ∈ Rp}. gra(E∗2 ) = {B⊺u,−A⊺u) A⊺u ∈ ran G, u ∈ Rp}. According to Fact 2.6, E∗i is maximal monotone and dim E∗i = n. This implies that dim{u ∈ Rp B⊺u ∈ dom G} = n, dim{u ∈ Rp A⊺u ∈ ran G} = n. Let Mi ∈ Rp×p with rank M = n and (34) {u ∈ Rp B⊺u ∈ dom G} = ran M1, (35) Corollary 3.8 shows that {u ∈ Rp A⊺u ∈ ran G} = ran M2. gra Ei = {(x, x∗) M ⊺ i Ax + M ⊺ i Bx∗ = 0}. However, finding Mi from (34) and (35) may not be easy as it seems. 6 Examples In the final section, we illustrate our maximally monotone extensions by considering three examples. In particular, they show that maximal monotone extensions eG rely on the rep- resentation of G in terms of A, B and choices of N we shall use. However, the maximal monotone extensions Ei are intrinsic, only depending on gra G. 19 Example 6.1 Consider where C is a n × n symmetric, positive definite matrix. Clearly, gra G =(cid:26)(cid:18) x x∗(cid:19) (cid:18)Id C(cid:19) x∗ = 0, x, x∗ ∈ Rn(cid:27) 0(cid:19) x +(cid:18) 0 0(cid:19)(cid:27) . gra G =(cid:26)(cid:18)0 We have (i) For every α ∈ [−1, 1] ,eGα defined by graeGα =({(0, Rn)} , (cid:8)(x, 1+α is a maximally monotone linear extension of G. 1−α C−1x) x ∈ Rn(cid:9) , if α = 1; otherwise (ii) E1 = eG1 and E2 = eG−1. Proof. (i): To find eGα, we need eigenvectors of 0(cid:19) (0 C ⊺) +(cid:18) 0 A =(cid:18)Id C(cid:19) (Id 0) =(cid:18) 0 C C 0(cid:19) Counting multiplicity, the positive definite matrix C has eigen-pairs (λi, wi) (i = 1, . . . , n) such that λi > 0,kwik = 1 and hwi, wji = 0 for i 6= j. As such, the matrix A has 2n eigen-pairs, namely and with i = 1, . . . , n. Put W = [w1 · · · wn] and write (λi,(cid:18)wi wi(cid:19)) −wi(cid:19)) (−λi,(cid:18) wi V =(cid:18)W W W −W(cid:19) . Then W ⊺CW = D = diag(λ1, λ2,· · · , λn) 20 In Theorem 3.7, take We have rank Nα = n, N ⊺ α Idλ Nα =(cid:18)0 being negative semidefinite, and 0 Id (cid:19) . Nα =(cid:18)0 α Id 0 (α2 − 1)W ⊺CW(cid:19) =(cid:18)0 V Nα =(cid:18)0 (1 + α)W 0 (α − 1)W(cid:19) . 0 0 0 (α2 − 1)D(cid:19) Then by Theorem 3.7, we have an maximally monotone linear extension eGα given by (1 + α)W ⊺x + (α − 1)W ⊺Cx∗(cid:19) = 0(cid:27) 0 graeGα =(cid:26)(x, x∗) (cid:18) =({(0, Rn)} , (cid:8)(x, 1+α = {(x, x∗) (1 + α)x + (α − 1)Cx∗ = 0} if α = 1; otherwise 1−α C−1x) x ∈ Rn(cid:9) , , Hence we get the result as desire. (ii): It is immediate from Theorem 5.2 and (i). (cid:4) Example 6.2 Consider gra G = x∗(cid:19)  (cid:18) x −1 0 0 0 0 −1 x2(cid:19) + (cid:18)x1 eG1 = 1 0 2−√2! , 0 −1+√2 are the maximally monotone extensions of G. . 1 0 0 1 0 1 x∗2(cid:19) = 0, xi, x∗i ∈ R (cid:18)x∗1 10(cid:19) eG2 =(cid:18)1 2 5 √2 0 Then (i) (ii) E1(x1, 0) = (x1, R) ∀x1 ∈ R. 21 (iii) Proof. We have The matrix has positive eigenvalue −1 + √2 with eigenvector 0 1 u = 1 − √2 Then by Corollary 3.4, graeG1 =  x1 ∈ R +   2 − √2 −1 + √2  Therefore, x1 0 x1 0 0 0 −1 + √2 0 0 Idλ = We have (36) Take (37) is monotone. Since dim G = 1, G is not maximally monotone by Fact 2.5. E2(x1, y) = (x1, 0) ∀x1, y ∈ R. x1 0 x1 0 −2 0 0 0 −1 0 0 −1 −2  x1 ∈ R gra G =  AB⊺ + BA⊺ =  A⊺(cid:19) u =  so that (cid:18)B⊺  x2 x2 ∈ R =  2−√2 ! . eG1 = 1  , V =  . N = 0 −1 1 2 −1 0 1 1 0 −1 − √2 0 −1+√2 0 0 −2 0 0 0 22 0 2 − √2 −1 + √2 0  . x1 (2 − √2)x2 (−1 + √2)x2 x1  x1, x2 ∈ R . 0 0 − 1 −1+√2 − 1 −1−√2 1 1 1 0 0  . We have rank N = 2 and (38) being negative semidefinite. 0 N ⊺ Idλ N = 0 0 0 −7 − 3√2 1 + √2 −4 1 + √2 0  , By Theorem 3.7, with V, N given in (36) and (37), we use the NullSpace command in Maple to solve and get (V N)⊺Ax + (V N)⊺Bx∗ = 0, graeG2 = span −2√2 5√2 0 1 1 0 1 0    .   2 5 √2 10(cid:19). x1 0 x1 R  x1 ∈ R E1(x1, 0) = (x1, R) ∀x1 ∈ R.  x1 ∈ R gra E2 =  0 0 0 R  =  +  x1 ∈ R x1 R x1 0 . E2(x1, y) = (x1, 0) ∀x1, y ∈ R. 0 0 On the other hand, Thus eG2 =(cid:18)1 −2√2 5√2 (cid:19)−1 =(cid:18)1 gra E1 =  x1 0 x1 0 gives And gives In [5], the authors use autoconjugates to find maximally monotone extensions of monotone In general, it is not clear whether the maximally monotone extensions of a operators. linear relation is still a linear relation. As both monotone operators in Examples 6.2 and (cid:4) 23 are two maximally monotone linear extensions of G. (39) Then Moreover, 1 5 1 7 0 2 1 1 2 0 3 1  , B = 2(−1+√201)! , −107+7√201 2(−1+√201) 2(−1+√201) 2(−1+√201) − −21+√201 − −23+3√201  x2 x1, x2 ∈ R A = eG1 = −117+17√201 gra E1 =  x1 +  Idλ = Proof. We have rank(A B) = 3 and 13 + √201 −1 1 −5 1 0 −6 0 0 0 1 1 (40) 0 0 0 0 4 − − 29 20 +  , thus (A B) = eG2 = 33 gra E2 = V = 1(cid:19) .  , −1 1 1 20 , 13 − √201 Vg =(cid:18) 0 1−√201 1 1 1 5 2 0 1 7 3 1 0 2  . 13 √201 4 − 6 √201 30 − 9 20 + √201 6 √201 30 ! . −1 1 −5 1   x1 + 1 5 0 0 20 1+√201 1 1 0 −1 1 20 1−√201 1 1 .  x2 x1, x2 ∈ R  , Examples 6.1 are subset of {(x, x) x ∈ Rn}, [5, Example 5.10] shows that the maximally monotone extension obtained by autoconjugate must be Id, which are different from the ones given here. Example 6.3 Set gra G = {(x, x∗) Ax + Bx∗ = 0} where and (41) Clearly, here p = 3, n = 2 and AB⊺ +BA⊺ has exactly p−n = 3−2 = 1 positive eigenvalue. By Theorem 3.3(i)(v), G is monotone. Since AB⊺ + BA⊺ is not negative semidefinite, by Theorem 3.6(i)(iii), G is not maximally monotone. With Vg given in (41) and A, B in (39), use the NullSpace command in maple to solve VgAx + VgBx∗ = 0 and obtain eG1 defined by  − −21+√201 2(−1+√201) −23+3√201 2(−1+√201) graeG1 = span  1 0 24 , −−107+7√201 2(−1+√201) −117+17√201 2(−1+√201) 0 1 .   eG1 = − −21+√201 By Corollary 3.4, eG1 is a maximally monotone linear subspace extension of G. Then 2(−1+√201)! . = −117+17√201 −107+7√201 2(−1+√201) 2(−1+√201) 2(−1+√201) − −21+√201 − −23+3√201  . 0 0 1 5 0 1 0 0 0 1 Let N be defined by 2(−1+√201) !−1 2(−1+√201) −−107+7√201 2(−1+√201) −117+17√201 −23+3√201 2(−1+√201) N = N ⊺ Idλ N = 0 0 0 −6 0 0 Then rank N = 2 and  . 338−24√201 (42) 0 0 25 is negative semidefinite. With N in (42), A, B in (39) and V in (40), use the NullSpace command in maple to solve (V N)⊺Ax+ (V N)⊺Bx∗ = 0. By Theorem 3.7, we get a maximally monotone linear extension of G, eG2, defined by eG2 = − 9 20 + 29 20 − √201 30 − 13 4 + √201 33 4 − 30 √201 6 √201 6 !−1 = 33 4 − − 29 20 + 13 √201 4 − 6 √201 30 − 9 20 + √201 6 √201 30 ! . To find E1 and E2, using the LinearSolve command in Maple, we get gra G = ran(cid:18)C D(cid:19), where It follows from Theorem 5.2 that −1 1 −5 1 C =(cid:18)−1 1 (cid:19) , D =(cid:18)−5 1 (cid:19) . gra E1 =  x1 +  gra E2 =  x1 +   x2 x1, x2 ∈ R  x2 x1, x2 ∈ R −1 1 −5 1 1 5 0 0 0 0 1 1 25 , . (cid:4) Acknowledgments The authors thank Dr. Heinz Bauschke for bring their attentions of [10] and many valuable discussions. Xianfu Wang was partially supported by the Natural Sciences and Engineering Research Council of Canada. References [1] H.H. Bauschke and P.L. Combettes, Convex Analysis and Monotone Operator Theory in Hilbert Spaces, Springer-Verlag, 2011. [2] H.H. Bauschke, X. Wang and L. Yao, Monotone linear relations: maximality and Fitz- patrick functions, Journal of Convex Analysis, vol. 16, pp. 673 -- 686, 2009. [3] H.H. Bauschke, X. Wang, and L. Yao, On Borwein-Wiersma Decompositions of mono- tone linear relations, SIAM Journal on Optimization, vol. 20, pp. 2636 -- 2652, 2010. [4] H.H. Bauschke and X. Wang, Firmly nonexpansive and Kirszbraun-Valentine exten- sions: a constructive approach via monotone operator theory, Nonlinear analysis and optimization I. Nonlinear analysis, Contemp. Math., 513, Amer. Math. Soc., Provi- dence, RI, pp. 55-64, 2010. [5] H.H. Bauschke and X. Wang, The kernel average for two convex functions and its applications to the extension and representation of monotone operators, Transactions of the American Mathematical Society, vol. 36, pp. 5947-5965, 2009. [6] J.M. Borwein and A.S. Lewis, Convex Analysis and Nonlinear Optimization, Theory and Examples. Second edition, CMS Books in Mathematics/Ouvrages de Mathmatiques de la SMC, 3 Springer, New York, 2006. [7] J.M. Borwein and J.D. Vanderwerff, Convex Functions, Cambridge University Press, 2010. [8] H . Br´ezis, F.E. Browder, Linear maximal monotone operators and singular nonlinear integral equations of Hammerstein type, in Nonlinear analysis (collection of papers in honor of Erich H. Rothe), Academic Press, pp. 31 -- 42, 1978. [9] R. Cross, Multivalued Linear Operators, Marcel Dekker, 1998. [10] J.P. Crouzeix and E.O. Anaya, Monotone and maximal monotone affine subspaces, Operations Research Letters, vol. 38, pp. 139 -- 142, 2010. 26 [11] J.P. Crouzeix and E.O. Anaya, Maximality is nothing but continuity, Journal of Convex Analysis, vol. 17, pp. 521-534, 2010. [12] J.P. Crouzeix, E. O. Anaya and W. S. Sandoval, A construction of a maximal monotone extension of a monotone map, ESAIM: Proceedings 20, pp. 93-104, 2007. [13] J. Eckstein and D.P. Bertsekas, On the Douglas-Rachford splitting method and the proximal point algorithm for maximal monotone operators, Mathematical Programming, vol. 55, pp. 293 -- 318, 1992. [14] C. Meyer, Matrix Analysis and Applied Linear Algebra, Society for Industrial and Ap- plied Mathematics (SIAM), Philadelphia, PA, 2000. [15] R.T. Rockafellar and R.J-B Wets, Variational Analysis, corrected 3rd printing, Springer- Verlag, 2009. [16] S. Simons, A Br´ezis-Browder theorem for SSDB spaces; http://arxiv.org/abs/1004.4251v3, September 2010. [17] M.D. Voisei and C. Zalinescu, Linear monotone subspaces of locally convex spaces, Set-Valued and Variational Analysis, 18, pp. 29 -- 55, 2010. [18] L. Yao, Decompositions and Representations of Monotone Operators with Linear Graphs, M. Sc. Thesis, The University of British Columbia Okanagan, December 2007. [19] L. Yao, The Br´ezis-Browder Theorem revisited and properties of Fitzpatrick functions of order n, to appear Fixed Point Theory for Inverse Problems in Science and Engineering; http://arxiv.org/abs/0905.4056v1, May 2009. 27
1203.0793
2
1203
2013-09-11T14:45:00
A geometric technique to generate lower estimates for the constants in the Bohnenblust--Hille inequalities
[ "math.FA" ]
The Bohnenblust--Hille (polynomial and multilinear) inequalities were proved in 1931 in order to solve Bohr's absolute convergence problem on Dirichlet series. Since then these inequalities have found applications in various fields of analysis and analytic number theory. The control of the constants involved is crucial for applications, as it became evident in a recent outstanding paper of Defant, Frerick, Ortega-Cerd\'{a}, Ouna\"{\i}es and Seip published in 2011. The present work is devoted to obtain lower estimates for the constants appearing in the Bohnenblust--Hille polynomial inequality and some of its variants. The technique that we introduce for this task is a combination of the Krein--Milman Theorem with a description of the geometry of the unit ball of polynomial spaces on $\ell^2_\infty$.
math.FA
math
A GEOMETRIC TECHNIQUE TO GENERATE LOWER ESTIMATES FOR THE CONSTANTS IN THE BOHNENBLUST -- HILLE INEQUALITIES G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA Abstract. The Bohnenblust -- Hille (polynomial and multilinear) inequalities were proved in 1931 in order to solve Bohr's absolute convergence problem on Dirichlet series. Since then these inequalities have found applications in various fields of analysis and analytic number theory. The control of the constants involved is crucial for applications, as it became evident in a recent outstanding paper of Defant, Frerick, Ortega-Cerd´a, Ounaıes and Seip published in 2011. The present work is devoted to obtain lower estimates for the constants appearing in the Bohnenblust -- Hille polynomial inequality and some of its variants. The technique that we introduce for this task is a combination of the Krein -- Milman Theorem with a description of the geometry of the unit ball of polynomial spaces on ℓ2 ∞. 1. Preliminaries and background In 1913 H. Bohr proved that the maximal width T of the vertical strip in which a Dirichlet series ann−s converges uniformly but not absolutely is always less or equal than 1/2. Since then, the determination of the precise value of T remained a central problem in the study of Dirichlet series. Almost 20 years later, in 1931, H.F. Bohnenblust and E. Hille [3] showed that in fact T = 1/2. The technique used for this task was based on a puzzling generalization of Littlewood's 4/3 inequality to the framework of m-linear forms and homogeneous polynomials. The Bohnenblust -- Hille inequality for homogeneous polynomials [3] asserts that if P : ℓN ∞ → C is a m-homogeneous polynomial, then there is a constant DC,m so that aαzα, P (z) = Pα=m ∞ Pn=1 (1.1) Pα=maα m+1 2m 2m m+1! ≤ DC,m kPk . The control of the estimates DC,m, besides its challenging nature, plays a decisive role in the theory: for instance, with adequate estimates for DC,m in hands, Defant, Frerick, Ortega-Cerd´a, Ounaıes and Seip [8] were able to solve several important questions related to Dirichlet series. In particular they obtained a definitive generalization of a result of Boas and Khavinson [2], showing that the n-dimension Bohr radius Kn satisfies Kn ≍r log n n . The main result of [8] asserts that there is a C > 1 such that DC,m ≤ Cm for all m, i.e., the Bohnenblust -- Hille inequality for homogeneous polynomials is hypercontractive. More precisely it was shown that and, for example, one can take C = 2 and it is simple to verify that DC,m ≤ 2m. DC,m ≤(cid:18)1 + 1 m − 1(cid:19)m−1 √m(cid:16)√2(cid:17)m−1 Key words and phrases. Absolutely summing operators, Bohnenblust -- Hille Theorem, Krein -- Milman Theorem. D. Pellegrino was supported by Supported by CNPq Grant 301237/2009-3, INCT-Matem´atica and CAPES- NF. G.A. Munoz-Fern´andez and J. B. Seoane-Sep´ulveda were supported by the Spanish Ministry of Science and Innovation, grant MTM2009-07848. 2010 Mathematics Subject Classification: 46G25, 47L22, 47H60. 1 2 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA It is worth mentioning that for small values of m, however, there are better estimates for DC,m due to Queff´elec [26, Th. III-1]; for instance DC,2 ≤ 1.7431. In view of the pivotal role played by the constants involved in the Bohnenblust -- Hille inequality, a natural step forward is to try to obtain sharp constants and for this reason the search for lower estimates for the constants gains special importance. Moreover it is interesting to mention that, historically, the upper estimates obtained for the Bohnenblust -- Hille inequalities have shown to be quite far from sharpness (see [8, 25] for details). Just to illustrate this fact, in the multilinear Bohnenblust -- Hille inequality (complex case) the original upper estimate for the constant when m = 10 is 80.28 but now we know that this constant is not grater than 2.3. The multilinear version of Bohnenblust -- Hille inequality is also an important subject of investi- gation in modern Functional Analysis and, as mentioned in [12], "it had and has deep applications in various fields of analysis, as for example in operator theory in Banach spaces, Fourier and harmonic analysis, complex analysis in finitely and infinitely many variables, and analytic number theory". For recent developments and related results we refer to [6, 9 -- 11]. Everything begins with Littlewood's famous 4/3 theorem which asserts that for K = R or C, ∞ 3  for every continuous bilinear form A on c0 × c0, with CK,2 = √2.  Xi,j=1  A(ei, ej) 3 4 4 ≤ CK,2 kAk It is well-known that the power 4/3 is optimal (see [18]). For real scalars it also can be shown that the constant √2 is optimal (see [16]). For complex scalars, however, there are several estimates for CC,2; below KG stands for the complex Grothendieck's constant, and it is well-known that 1.338 ≤ KG ≤ 1.405 (see [14]): ([7, Theorem 34.11] or [27, Theorem 11.11]), • CC,2 ≤(cid:0)KG√2(cid:1)1/2 • CC,2 ≤ KG ([21, Corollary 2, p. 280]), • CC,2 ≤ 2√π ≈ 1.128 ([13, 26]). The optimal value for CC,2 seems unknown. In 1931 Bohnenblust and Hille [3] observed the connection between Littlewood's 4/3 theorem and the so called Bohr's absolute convergence prob- lem for Dirichlet series, which had been open for over 15 years. So, they generalized Littlewood's result to multilinear mappings, homogeneous polynomials and answered Bohr's problem. Although the work of Bohnenblust and Hille is focused on complex scalars, it is well-known that the result also holds for real scalars: If A is a continuous n-linear form on c0 ×···× c0, then there is a constant CK,n (depending only on n and K) such that ∞  Xi1,...,im=1  A(ei1 , ..., ein ) n+1 2n ≤ CK,n kAk . 2n n+1  The estimates for CK,n were improved along the decades (see [5, 19, 26]). From recent works (see [16, 25]) we know that, for real scalars, CR,2 = √2 ≈ 1.414 1.587 ≤ CR,3 ≤ 1.782 1.681 ≤ CR,4 ≤ 2 1.741 ≤ CR,5 ≤ 2.298 1.811 ≤ CR,6 ≤ 2.520 and, for the complex case, A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 3 √π(cid:19) ≈ 1.128 CC,2 ≤(cid:18) 2 CC,3 ≤ 1.273 CC,4 ≤ 1.437 CC,5 ≤ 1.621 CC,10 ≤ 2.292 CC,15 ≤ 2.805. m , so the precise value for CR,m with "big m" is The lower bounds for CR,m obtained in [16] are 2 quite uncertain. Very recently, it was shown that for both real and complex scalars the asymptotic behavior of the best values for CK,n is optimal [15]. m−1 The (complex and real) Bohnenblust -- Hille inequality can be re-written in the context of multiple summing multilinear operators. Let X1, . . . , Xm and Y be Banach spaces over K = R or C, and X′ be the topological dual of X. By L(X1, . . . , Xm; Y ) we denote the Banach space of all continuous m-linear mappings from X1 × ··· × Xm to Y with the usual sup norm. For x1, ..., xn in X, let k(xj )n j=1kw,1 := sup{k(ϕ(xj))n j=1k1 : ϕ ∈ X′,kϕk ≤ 1}. If 1 ≤ p < ∞, an m-linear mapping U ∈ L(X1, . . . , Xm; Y ) is multiple (p; 1)-summing (denoted Π(p;1)(X1, . . . , Xm; Y )) if there exists a constant UK,m ≥ 0 such that (1.2) U (x(1) j1 , . . . , x(m) jm ≤ UK,m   N Xj1,...,jm=1(cid:13)(cid:13)(cid:13) 1 p p )(cid:13)(cid:13)(cid:13)  m Yk=1(cid:13)(cid:13)(cid:13) (x(k) j )N j=1(cid:13)(cid:13)(cid:13)w,1 for every N ∈ N and any x(k) jk ∈ Xk, jk = 1, . . . , N , k = 1, . . . , m. The infimum of the constants satisfying (1.2) is denoted by kUkπ(p;1). For m = 1 we recover the well-known concept of absolutely (p; 1)-summing operators (see, e.g. [7, 14]). The Bohnenblust -- Hille inequality can be re-written in the context of multiple summing multi- linear operators in the following sense: every continuous m-linear form U : X1 × ··· × Xm → K is multiple ( 2m m+1 ; 1)-summing. Moreover (1.3) kUkπ( 2m m+1 ;1) ≤ CK,m kUk . For details we refer to [12] and references therein. From now on if P : X → Y is a m-homogeneous polynomial then ∨P denotes the (unique) sym- metric m-linear map (also called the polar of P ) associated to P . Recall that an m-homogeneous polynomial P : X → Y is multiple (p; 1)-summing (denoted P(p;1)(mX; Y )) if there exists a con- stant PK,m ≥ 0 such that jk ∈ X, jk = 1, . . . , N , k = 1, . . . , m. The infimum of the constants 1 p ≤ PK,m N jm (1.4)   ∨P (x(1) j1 , . . . , x(m) Xj1,...,jm=1(cid:13)(cid:13)(cid:13)(cid:13) p )(cid:13)(cid:13)(cid:13)(cid:13)  for every N ∈ N and any x(k) satisfying (1.4) is denoted by kPkπ(p;1). Note that kPkπ(p;1) =(cid:13)(cid:13)(cid:13)(cid:13) ∨P(cid:13)(cid:13)(cid:13)(cid:13)π( 2m If P ∈ P(mX; K) then ∨P ∈ L(mX; K) = Π( 2m m+1 ;1) =(cid:13)(cid:13)(cid:13)(cid:13) kPkπ( 2m m+1 ;1) (1.3) . ∨P(cid:13)(cid:13)(cid:13)(cid:13)π(p;1) ∨P(cid:13)(cid:13)(cid:13)(cid:13) ≤ CK,m(cid:13)(cid:13)(cid:13)(cid:13) m+1 ;1) (mX; K) and m Yk=1(cid:13)(cid:13)(cid:13) (x(k) j )N j=1(cid:13)(cid:13)(cid:13)w,1 mm m! ≤ CK,m kPk . 4 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA So, since CK,m does not depend on X and P we conclude that there are constants LK,m (which does not depend on X and P ) such that N  Xj1,...,jm=1(cid:13)(cid:13)(cid:13)(cid:13)  ∨P (x(1) j1 , . . . , x(m) jm ≤ LK,m kPk m Yk=1(cid:13)(cid:13)(cid:13) (x(k) j . )N j=1(cid:13)(cid:13)(cid:13)w,1 Note that if X = ℓN ∞, and x(j) = ej for every j = 1, ..., N , since 1 p (x(j))N p )(cid:13)(cid:13)(cid:13)(cid:13)  (cid:13)(cid:13)(cid:13) j=1(cid:13)(cid:13)(cid:13)w,1 m+1 ∨P (ej1 , . . . , ejm)(cid:13)(cid:13)(cid:13)(cid:13)  2m = 1, N  Xj1,...,jm=1(cid:13)(cid:13)(cid:13)(cid:13)  m+1 2m ≤ LK,m kPk we have (1.5) for every N ∈ N, which can be regarded as a kind of polynomial Bohnenblust -- Hille inequality. values for CK,m and the optimal values of LK,m. Since (1.5) is confined to the symmetric case, there is no obvious relation between the optimal For m = 2 it is well-known that CR,2 = √2. For m > 2 the precise values of CR,m are not known. Since we have LR,m ≤ mm m! CR,m, LR,2 ≤ 2.828 LR,3 ≤ 8.018 LR,4 ≤ 21.333 The main goal of this paper is to introduce a technique that helps to find nontrivial lower bounds for the constants involved in the Bohnenblust -- Hille inequalities. Our approach is shown to be effective for the cases of LR,m and DR,m. In the complex case we succeed in obtaining a lower bound for DC,2. More precisely, as a consequence of our estimates we show that if DR,m > 0 is such that for all m-homogeneous polynomial P : ℓN then ≤ DR,m kPk , Pα=maα m+1 2m 2m m+1! ∞ → R, P (x) = Pα=m DR,m ≥ (1.495)m . aαxα, Regarding to LR,m, we show, for instance, that 1.770 ≤ LR,2 1.453 ≤ LR,3 2.371 ≤ LR,4 3.272 ≤ LR,8 5.390 ≤ LR,16 In the complex case we show that DC,2 ≥ 1.1066. So, combining this information with the best known upper estimate known for DC,2 we conclude that The techniques used in this paper in order to obtain good estimates for the constants LK,n and 1.1066 ≤ DC,2 ≤ 1.7431. DK,n are based on the following result: A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 5 Theorem 1.1 (consequence of Krein -- Milman Theorem). If C is a convex body in a Banach space and f : C → R is a convex function that attains its maximum, then there is an extreme point e ∈ C so that f (e) = max{f (x) : x ∈ C}. This consequence of the Krein -- Milman Theorem ([20]) provides good lower estimates on the constants LK,n when it is combined with a description of the geometry of the unit ball of a polynomial space on ℓm ∞. The problem of finding the extreme points of the unit ball of a polynomial space has been largely studied in the past few years. In particular, the following results will be particularly useful for our purpose. Theorem 1.2 (Choi & Kim [4]). The extreme points of the unit ball of P(2ℓ2 ∞) are the polynomials of the form ±x2, ±y2, ±(tx2 − ty2 ± 2pt(1 − t)xy), with t ∈ [1/2, 1]. Theorem 1.3 (G´amez-Merino, Munoz-Fern´andez, S´anchez, Seoane-Sep´ulveda [17]). If P(2(cid:3)) denotes the space P(2R2) endowed with the sup norm over the unit interval (cid:3) = [0, 1]2 and B(cid:3) is its unit ball, then the extreme points of B(cid:3) are ±(tx2 − y2 + 2√1 − txy) and ± (−x2 + ty2 + 2√1 − txy) with t ∈ [0, 1] or ±(x2 + y2 − xy), ±(x2 + y2 − 3xy), ±x2, ±y2. Note that Theorem 1.3 is a kind on non-symmetric version of Theorem 1.2 and will be specially important when we are estimating the constants for m ≥ 4. 2. Estimates for LR,m In order to deal with polynomials and their polars we will introduce some notation and a few basic results. If α = (α1, . . . , αn) ∈ N∗ then we define α := α1 + ··· + αn and (cid:18)m α(cid:19) := m! , α1!··· αn! for α = m ∈ N∗. Also, xα stands for the monomial xα1 for x = (x1, . . . , xn) ∈ Kn. Having all this in mind, a straightforward consequence of the multinomial formula yields the following relationship between the coefficients of a homogeneous polynomial and the polar of the polynomial. 1 ··· xαn n Lemma 2.1. If P is a homogeneous polynomial of degree n on Kn given by P (x1, . . . , xn) = Xα=m aαxα, and L is the polar of P , then aα L(eα1 1 , . . . , eαn , n ) = where {e1, . . . , en} is the canonical basis of Kn and eαk Definition 2.2. Let us call d the dimension of the space of all m-homogeneous polynomials on Rn. For every m, n ∈ N, we define Φm,n : Rd → R as follows: Take a ∈ Rd and consider the the m-homogeneous polynomial Pa(x) = Pα=m aαxα whose coefficients are the coordinates of a. In order to avoid redundancies, assume that a = (aα) where the coordinates are arranged according to the lexicographic order of the α's. Then if La is the polar of Pa we define stands for ek repeated αk times. (cid:0)m α(cid:1) k Φm,n(a) :=" Xi1+···+im=m La(ei1 , . . . , eim ) m+1 2m . 2m m+1# 6 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA Remark 2.3. Notice that Lemma 2.1 allows us to write Φm,n as (2.1) Φm,n(a) = 1 , . . . , eαn n ) α=m (cid:18)m α(cid:19)La(eα1  Xα=(α1 ,··· ,αn) m+1 = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) α(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  Xα=m(cid:18)m  (cid:0)m α(cid:1) aα 2m m+1 2m . m+1 2m 2m m+1  with the natural isomorphism Also Φm,n is, essentially, the composition of the norm in ℓd 2m m+1 between Ls(mRn) and P(mRn). Therefore Φm,n is convex and by virtue of Krein -- Milman Theorem LR,m ≥ LR,m (ℓn ∞) := sup{Φm,n(a) : a ∈ B P(mℓn ∞)) is the set of extreme points of B P(mℓn where ext(B P(mℓn the geometry of B ∞). Observe that even in the case where ∞) is not know, the mapping Φm,n provides a lower bound for LR,m, namely ∞)} = sup{Φm,n(a) : a ∈ ext(B P(mℓn P(mℓn ∞))}, (2.2) for all a ∈ Rd. LR,m ≥ Φm,n(a) kPak , In the following we will try to use the fact that the extreme points of B characterized for some choices of m and n (see for instance Theorem 1.2). ∞) have been P(mℓn 2.1. Case m = 2. We begin by illustrating that even sharp information for lower estimates for CR,2 may be useless for evaluating lower estimates for LR,2. For instance, if m = 2 in the multilinear Bohnenblust -- Hille inequality (in fact, Littlewood's 4/3 inequality) the best constant is CR,2 = √2 and this estimate is achieved (see [16]) when we use the bilinear form T2 : ℓ2 ∞ → R given by ∞ × ℓ2 T2(x, y) = x1y1 + x1y2 + x2y1 − x2y2. Note that T2 is symmetric and the polynomial associated to T2 is P2 : ℓ2 ∞ → R given by P2(x) = x2 1 + 2x1x2 − x2 2. Since kP2k = kT2k = 2, the constant LR,2 that appears for this choice of P2 is again √2, which is far from being a good lower estimate, as we shall see in the next result, that gives the exact value for the constant LR,2(ℓ2 ∞). Theorem 2.4. LR,2 ≥ 1.7700. More precisely, LR,2(cid:0)ℓ2 ∞(cid:1) = sup((cid:20)2t 4 3 + 2(cid:16)pt(1 − t)(cid:17) 3 4 4 3(cid:21) : t ∈ [1/2, 1]) ≈ 1.7700 and the supremum is attained at t0 ≈ 0.9147. Proof. Observe that for polynomials in P(2ℓ2 a = (a, b, c) we have ∞) of the form Pa(x, y) = ax2 + by2 + cxy with (2.3) Using the Krein -- Milman approach Now, by Theorem 1.2, ext(B 4 4 3 + b Φ2,2(a, b, c) =(cid:20)a 3 + 2(cid:16) c 2(cid:17) ∞(cid:1) = sup{Φ22(a) : a ∈ ext(B ∞)) consists of the polynomials LR,2(cid:0)ℓ2 P(2ℓ2 3 4 . 4 3(cid:21) P(2ℓ2 ∞))}. ±(1, 0, 0), ±(0, 1, 0) and ± (t,−t,±2pt(1 − t)), A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 7 with t ∈ [1/2, 1]. Since the contribution of ±(1, 0, 0) and ±(0, 1, 0) to the supremum is irrelevant, we end up with LR,2(cid:0)ℓ2 ∞(cid:1) = sup{Φ2,2(±(t,−t,±2pt(1 − t))) : t ∈ [1/2, 1]} = sup((cid:20)2t : t ∈ [1/2, 1]) . 3 + 2(cid:16)pt(1 − t)(cid:17) 3(cid:21) 3 4 4 4 The problem of maximizing explicitly this function is a hard one and the final result is far from being good looking. The interested reader can obtain an explicit solution in radical form using a variety of symbolic calculus packages, such as Mathematica, Matlab or Maple. A 4-digit approximation yields LR,2 ≥ LR,2(cid:0)ℓ2 where the maximum is attained at t0 ≈ 0.9147. Remark 2.5. A very good approximation of LR,2(cid:0)ℓ2 i.e., a = (1,−1, 1). It is easy to check that kPak = 5/4. Hence, using (2.2) we have ∞(cid:1) ≈ 1.7700, ∞(cid:1) can be obtained considering the polynomial Pa(x, y) = x2 − y2 + xy, (cid:3) LR,2(cid:0)ℓ2 ∞(cid:1) ≥ Φ2,2(1,−1, 1) kPak = 4 5 · 2 + 2(cid:18) 1 2(cid:19)4/3!3/4 ≈ 1.728. 2.2. Case m = 4. In this section we calculate the exact value of LR,4 in a subspace of Ls(4ℓ2 ∞). Observe that the value of LR,4 in a subspace is, obviously, a lower bound for LR,4. Theorem 2.6. If E = {ax4 + by4 + cx2y2 : a, b, c ∈ R} and ∨E is the space of polars of elements in E endowed with the sup norm over the unit ball of ℓ2 Moreover, equality is attained in the Bohnenblust-Hille inequality in ∨E for the polars of the poly- nomials P (x, y) = ±(x4 − y4 + 3xy). Proof. We just need to calculate the maximum of Φ4,2 over E, which is trivially isometric to the space P(2(cid:3)) (see Theorem 1.3 for the definition of P(2(cid:3))). If Φ = Φ4,2P(2(cid:3)), then Φ is obviously convex and we have LR,4 ≥ LR,4(cid:0)ℓ2 ∞(cid:1) = sup{Φ4,2(a) : a ∈ B P(2(cid:3)))}, ≥ sup{Φ(a) : a ∈ B = sup{Φ(a) : a ∈ ext(B P(2(cid:3))} P(4ℓ2 ∞)} where the last equality is due to the Krein -- Milman Theorem. Now by (2.1) we have Φ(a, b, c) =(cid:20)a 8 5 + b 5 8 . 8 5 + 6(cid:16) c 6(cid:17) 8 5(cid:21) Using Theorem 1.3 we obtain sup{Φ(a) : a ∈B P(2(cid:3))} 8 5 + 6(cid:18)√1 − t = max "1 + t 3 (cid:19)  5# ="2 + 6(cid:18) 1 2(cid:19) 5 8 . 8 5 8 8 5# ,"2 + 6(cid:18) 1 6(cid:19) 8 5# 5 8 5 8 ,"2 + 6(cid:18) 1 2(cid:19) 8 5# : t ∈ [0, 1]  In particular 8 5 8 ∞, then 5# ≈ 2.371. LR,4( ∨E) ="2 + 6(cid:18) 1 2(cid:19) LR,4 ≥ LR,4(ℓ2 ∞) ≥ LR,4( ∨E) ≈ 2.371. 8 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA Observe that the maximum is attained at the polynomials P (x, y) = ±(x4 − y4 + 3xy). Hence we have proved that 8 5 8 5# ≈ 2.371, LR,4 ≥"2 + 6(cid:18) 1 2(cid:19) moreover, a better (bigger) lower estimate for LR,4 cannot be obtained by considering polynomials of the form ax4 + by4 + cx2y2 with a, b, c ∈ R. (cid:3) 2.3. Higher values of m. The previous sections allow us to obtain lower estimates for LR,m for arbitrary large m's. In this section we consider polynomials of the form P2k(x, y) = (ax2 + by2 + cxy)k. In the following, if h ∈ Z, ⌊h⌋ denotes the biggest integer H so that H ≤ h. Proposition 2.7. If P2k(x, y) = (ax2 + by2 + cxy)k, then P2k(x, y) =P2k j=0 Ajxj y2k−j with (2.4) Aj = for j = 0, . . . , 2k. ⌊ j 2⌋ Xℓ=0 k!aℓbk−j+ℓcj−2ℓ ℓ!(j − 2ℓ)!(k − j + ℓ)! , Proof. Using the multinomial formula: P2k(x, y) = (ax2 + by2 + cxy)k = Xα1+α2+α3=k α1,α2,α3≥0 k! α1!α2!α3! aα1 bα2cα3 x2α1+α3y2α2+α3. Therefore, xj y2n−j = x2α1+α3 y2α2+α3 for j = 1, . . . , 2k implies that ( 2α1 + α3 = j, 2α2 + α3 = 2k − j, which, together with the fact that α1 + α2 + α3 = k and α1, α2, α3 ≥ 0 yield ( α3 = j − 2α1, α2 = k − j + α1, with α1 = 0, . . . ,(cid:4) j Corollary 2.8. If k ∈ N then 2(cid:5). As a result of the previous comments, the coefficient Aj is given by (2.4). (cid:3) (2.5) where Aj = 2k+1 4k , 4k 2k Aj LR,2k ≥  2k+1  j (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0(cid:18)2k (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:0)2k j(cid:1) k!(−1)k−j+ℓtk−j+2ℓ (2pt0(1 − t0))j−2ℓ ℓ!(j − 2ℓ)!(k − j + ℓ)! 0 2⌋ ⌊ j Xℓ=0 , for j = 0, . . . , 2k and t0 is as in Theorem 2.4. Proof. If P2k(x, y) = (ax2 + by2 + cxy)k, using (2.1), (2.2) and Proposition 2.7 we arrive at LR,2k ≥ 1 kP2kk  2k j (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0(cid:18)2k Aj (cid:0)2k j(cid:1) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 4k 2k+1  2k+1 4k , with Aj as in (2.4). Then the corollary follows by considering the polynomial which has norm 1. P2k(x, y) = (t0x2 − t0y2 + 2pt0(1 − t0)xy)2k, (cid:3) A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 9 Hence (2.5) provides a systematic formula to obtain a lower bound for LR,m for even m's. Observe that for k = 2 we have LR,4 ≥"2t 16 5 0 + 6(cid:18) 2t0 − 3t2 3 0 8 5 (cid:19) + 8(t0pt0(1 − t0)) 5 8 8 5# ≈ 2.1595, which is a slightly worse constant than the one obtained in Section 2.2. Actually, the estimates (2.5) can be improved for multiples of 4. Indeed, we just need to consider the polynomials Q4k(x, y) = (ax4 + by4 + cx2y2)k, with k ∈ N. Using exactly the same procedure described in this section LR,4k ≥ Aj 1 2k kQ4kk 2j(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0(cid:18)4k  8k 4k+1  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:0)4k 2j(cid:1) where the Aj's, with j = 1, . . . , 2k are the same as in (2.4). Now, putting a = 1, b = 1 and c = −3, i.e., considering powers of the extreme polynomial that appeared in Section 2.2, we would have that kQ4kk = 1 for all k ∈ N, which proves the following: Theorem 2.9. If k ∈ N then 4k+1 8k , (2.6) where (2.7) for j = 0, . . . , 2k. 4k+1 8k , LR,4k ≥  2⌋ ⌊ j Xℓ=0 Bj = 2k 2j(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0(cid:18)4k 8k 4k+1  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Bj (cid:0)4k 2j(cid:1) k!(−3)j−2ℓ ℓ!(j − 2ℓ)!(k − j + ℓ)! , As an example, let us apply (2.6) and (2.7) to obtain estimates for LR,8 and LR,12. The poly- nomials are Then Q8(x, y) = x8 − 6x6y2 + 11x4y4 − 6x2y6 + y8, Q12(x, y) = x12 − 9x10y2 + 30x8y4 − 45x6y6 + 30x4y8 − 9x2y10 + y12. LR,8 ≥ LR,12 ≥ 2 + 2(cid:18)8 2 + 2(cid:18)12 ! 2(cid:19) 6 (cid:0)8 2(cid:1) 2(cid:19) 9 (cid:0)12 2(cid:1) For higher degrees see Table 1. 9 16 16 9 ! 16 9  ! 4(cid:19) 11 +(cid:18)8  (cid:0)8 4(cid:1) ! 4(cid:19) 30 + 2(cid:18)12 (cid:0)12 4(cid:1) 24 13 LR,16 ≥ 5.390975019 k = 4 LR,20 ≥ 6.787708182 k = 5 LR,24 ≥ 8.511696468 k = 6 LR,36 ≥ 16.65124974 k = 9 k = 10 LR,40 ≥ 20.81051033 ≈ 4.2441. ≈ 3.2725. 13 24 24 24 13 ! +(cid:18)12 6(cid:19) 45 (cid:0)12 6(cid:1) 13  k = 40 LR,160 ≥ 16805.46318 LR,200 ≥ 1.5654 × 105 k = 50 LR,240 ≥ 1.4581 × 106 k = 60 LR,360 ≥ 1.1781 × 109 k = 90 k = 100 LR,400 ≥ 1.0972 × 1010 Table 1. LR,4k for some values of k In order to clarify what the asymptotic growth of the sequence (LR,4k)k∈N is, a simple calculation of the quotients of the estimates obtained in Table 1 for higher values of k indicates that the ratio of the estimates on LR,4(k+1) and LR,4k seem to tend to 5 4 . 10 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA 3. Estimates for DR,m ∞) has dimension d, then DR,m(ℓn and P(mℓn (smallest) equivalence constant between the spaces ℓd First observe that if P(mℓn the polynomial Pa(x) = Pα=m aαxα ∈ P(mℓn : Pa ∈ P(mℓn ∞) = sup(kak 2m kPak ∞) is nothing but the optimal ∞). In other words, if we identify ∞) with the vector a in Rd of all its coefficients, ∞)) = supnkak 2m ∞)o , : Pa ∈ B DR,m(ℓn P(mℓn (3.1) then m+1 2m m+1 m+1 where k · kp denotes the ℓp norm. By convexity of k · kp we also have (3.2) P(mℓn : Pa ∈ ext(B DR,m(ℓn m+1 ∞) = supnkak 2m As an easy consequence of Theorem 1.2 and (3.2) we have: ∞))o . Theorem 3.1. DR,2 ≥ DR,2(ℓ2 ∞) = sup((cid:20)2t 4 3 +(cid:16)2pt(1 − t)(cid:17) 3 4 4 3(cid:21) : t ∈ [1/2, 1]) ≈ 1.8374. The above supremum can be given explicitly in radical form using a symbolic calculus package, however the result is too lengthy to be shown. An excellent approximation can be obtain though in very simple terms considering the polynomial P ∈ P(2ℓ2 P (x, y) = x2 − y2 + xy. ∞) defined by Since kPk = 5/4, from (3.1) it follows that DR,2 ≥ DR,2(ℓ2 ∞) ≥ (3)3/4 5/4 ≈ 1.823. 3.1. The case m = 3. Let us define P3 : ℓ6 We have kP3k = 2 × 5 P3(x) = (x1 + x2)(cid:0)x2 4 . Also ∞ → R by 3 + x3x4 − x2 5 + x5x6 − x2 6(cid:1) . 4(cid:1) + (x1 − x2)(cid:0)x2 ≤ DR,3 kP3k . 4 6 Pα=3aα 6 4! Therefore Proposition 3.2. DR,3 ≥ (4 × 3)4/6 2 × 5 4 ≈ 2.096. 3.2. The case m = 4. Acting as in Section 2.2, we can prove that the maximum value of kak 8 kPak where Pa ranges over the subspace of P(4ℓ2 ∞) given by 5 {ax4 + by4 + cx2y2 : a, b, c ∈ R}, is attained for the polynomial Q4(x, y) = x4 + y4 − 3x2y2. Hence, by (3.1), we have: Theorem 3.3. If E = {ax4 + by4 + cx2y2 : a, b, c ∈ R} is endowed with the sup norm over the unit ball of ℓ2 ∞, then In particular DR,4(E) = k(1, 1,−3)k 8 5 DR,4 ≥ DR,4(ℓ2 =(cid:16)2 + (3)8/5(cid:17)5/8 ∞) ≥ DR,4(E) ≈ 3. 610. ≈ 3. 610. Moreover, equality is attained in the polynomial Bohnenblust-Hille inequality in E for the polyno- mials P (x, y) = ±(x4 − y4 + 3xy). A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 11 3.3. Higher values of m. We consider again the polynomials for all k ∈ N. Notice that kQ4kk = 1 for all k ∈ N. Therefore, using (3.1) together with the formula for the coefficients of the Q4k given by (2.7), we can obtain estimates for DR,4k with k arbitrary (see Table 2). In fact we have: Q4k(x, y) =(cid:0)x4 + y4 − 3x2y2(cid:1)k , Theorem 3.4. If k ∈ N then where for j = 0, . . . , 2k. 2k DR,4k ≥ Xj=0  Bj = ⌊ j 2⌋ Xℓ=0 4k+1 8k , 8k Bj 4k+1  k!(−3)j−2ℓ ℓ!(j − 2ℓ)!(k − j + ℓ)! , m = 8 DR,8 ≥ 14.86998167 m = 80 DR,80 ≥ 3.0496 × 1013 m = 12 DR,12 ≥ 66.39260961 m = 120 DR,120 ≥ 2.6821 × 1020 m = 16 DR,16 ≥ 306.6665737 m = 160 DR,160 ≥ 2.4320 × 1027 m = 20 DR,20 ≥ 1442.799763 m = 200 DR,200 ≥ 2.2443 × 1034 m = 24 DR,24 ≥ 6866.770014 m = 240 DR,240 ≥ 2.0924 × 1041 m = 28 DR,28 ≥ 32940.16505 m = 280 DR,280 ≥ 1.9649 × 1048 m = 32 DR,32 ≥ 1.5892 × 105 m = 320 DR,320 ≥ 1.8549 × 1055 m = 36 DR,36 ≥ 7.7009 × 105 m = 360 DR,360 ≥ 1.7582 × 1062 m = 40 DR,40 ≥ 3.7444 × 106 m = 400 DR,400 ≥ 1.6718 × 1069 Table 2. Estimates for DR,m for some values of m. Obtaining more constants, we also get the following representation on the form Cm of these lower bounds: DR,8 ≥ (1.40132479)8 m = 5600 DR,5600 ≥ (1.49475760)5600 m = 8 m = 200 DR,200 ≥ (1.48509930)200 m = 6400 DR,6400 ≥ (1.49482368)6400 m = 7200 DR,7200 ≥ (1.49487590)7200 m = 800 DR,800 ≥ (1, 49212548)800 m = 1600 DR,1600 ≥ (1, 49357368)1600 m = 8000 DR,8000 ≥ (1.49491825)8000 m = 3200 DR,3200 ≥ (1.49437981)3200 m = 8800 DR,8800 ≥ (1.49495333)8800 m = 4000 DR,4000 ≥ (1.49455267)4000 m = 9600 DR,9600 ≥ (1.49498289)9600 m = 4800 DR,4800 ≥ (1.49467111)4800 m = 12000 DR,12000 ≥ (1.49504910)12000 Table 3. Estimates for DR,m in the form DR,m ≥ Cm. Let P2 : ℓ2 4. A lower estimate for DC,2 ∞ (C) → C be a 2-homogeneous polynomial given by 2 + cz1z2. P2(z1, z2) = az2 1 + bz2 with a, b, c ∈ R. The following result can be obtained from a standard application of the Maximum Modulus Principle together with [1, eq. (3.1)]. 12 G.A. MU NOZ-FERN ´ANDEZ, D. PELLEGRINO, J. RAMOS CAMPOS AND J.B. SEOANE-SEP ´ULVEDA Proposition 4.1. If P2 : ℓ2 then ∞ (C) → C is defined by P2(z1, z2) = az2 1 + bz2 2 + cz1z2 with a, b, c ∈ R, kP2k =(a + b + c (a + b)q1 + c2 4ab if ab ≥ 0 or c(a + b) > 4ab, otherwise. So, for these polynomials P2 and ab < 0 and c(a + b) ≤ 4ab, the Bohnenblust -- Hille inequality is and thus (cid:16) 3√a4 + 3√b4 + 3√c4(cid:17) 3 4 ≤ DC,2 (a + b)s1 + c2 4 ab So, we must find real scalars a, b, c so that ab < 0, c(a + b) ≤ 4ab and . 3 4 DC,2 ≥ (cid:16) 3√a4 + 3√b4 + 3√c4(cid:17) (a + b)q1 + c2 f2(a, b, c) = (cid:16) 3√a4 + 3√b4 + 3√c4(cid:17) (a + b)q1 + c2 4ab 3 4 4ab is as big as possible. A straightforward examination shows that for all a, b, c and, on the other hand, f2(a, b, c) < 1.1067 f2(1,−1, 352 203 125 000 ) ≈ 1.1066. Combining the previous result and the known fact that DC,2 ≤ 1.7431 we have the following result: Theorem 4.2. 1.1066 ≤ DC,2 ≤ 1.7431. 5. Final remarks In the real case we were able to deal with the case m ≥ 2 even in the absence of information on the geometry of the unit ball of P(mℓn ∞). However, in the complex case the technique seemed less effective for m ≥ 3. For obtaining lower estimates for DC,m, with m ≥ 3 and sharper estimates for DR,m, we believe that some effort should be made to get more information on the geometry of the unit ball of P(mℓn geometry of the unit ball of complex and real polynomial spaces on ℓn ∞. We do hope that the present work may serve as a motivation for future works investigating the ∞) for higher values of m, n. References [1] R. M. Aron and M. Klimek, Supremum norms for quadratic polynomials, Arch. Math. (Basel) 76 (2001), no. 1, 73 -- 80. [2] H. P. Boas and D. Khavinson, Bohr's power series theorem in several variables, Proc. Amer. Math. Soc. 125 (1997), no. 10, 2975 -- 2979. [3] H. F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. (2) 32 (1931), no. 3, 600 -- 622. [4] Y. S. Choi and S. G. Kim, The unit ball of P(2l2 [5] A. M. Davie, Quotient algebras of uniform algebras, J. London Math. Soc. (2) 7 (1973), 31 -- 40. [6] A. Defant, J. C. D´ıaz, D. Garc´ıa, and M. Maestre, Unconditional basis and Gordon-Lewis constants for spaces 2), Arch. Math. (Basel) 71 (1998), no. 6, 472 -- 480. of polynomials, J. Funct. Anal. 181 (2001), no. 1, 119 -- 145. [7] A. Defant and K. Floret, Tensor norms and operator ideals, North-Holland Mathematics Studies, vol. 176, North-Holland Publishing Co., Amsterdam, 1993. [8] A. Defant, L. Frerick, J. Ortega-Cerd`a, M. Ounaıes, and K. Seip, The Bohnenblust-Hille inequality for homo- geneous polynomials is hypercontractive, Ann. of Math. (2) 174 (2011), no. 1, 485 -- 497. [9] A. Defant, D. Garc´ıa, M. Maestre, and D. P´erez-Garc´ıa, Bohr's strip for vector valued Dirichlet series, Math. Ann. 342 (2008), no. 3, 533 -- 555. A GEOMETRIC TECHNIQUE FOR THE BOHNENBLUST -- HILLE INEQUALITIES 13 [10] A. Defant, D. Garc´ıa, M. Maestre, and P. Sevilla-Peris, Bohr's strips for Dirichlet series in Banach spaces. part 2, Funct. Approx. Comment. Math. 44 (2011), no. part 2, 165 -- 189. [11] A. Defant, M. Maestre, and U. Schwarting, Bohr radii for vector valued holomorphic functions, Preprint. [12] A. Defant, D. Popa, and U. Schwarting, Coordinatewise multiple summing operators in Banach spaces, J. Funct. Anal. 259 (2010), no. 1, 220 -- 242. [13] A. Defant and P. Sevilla-Peris, A new multilinear insight on Littlewood's 4/3-inequality, J. Funct. Anal. 256 (2009), no. 5, 1642 -- 1664. [14] J. Diestel, H. Jarchow, and A. Tonge, Absolutely summing operators, Cambridge Studies in Advanced Mathe- matics, vol. 43, Cambridge University Press, Cambridge, 1995. [15] D. Diniz, G. A. Munoz-Fern´andez, D. Pellegrino, and J. B. Seoane-Sep´ulveda, The asymptotic growth of the constants in the Bohnenblust-Hille inequality is optimal, arXiv:1108.1550v2 [math.FA]. [16] , Lower bounds for the constants in the Bohnenblust -- Hille inequality: the case of real scalars, arXiv:1111.3253v2 [math.FA]. [17] J. L. G´amez-Merino, G. A. Munoz-Fern´andez, V. S´anchez, and J. B. Seoane-Sep´ulveda, Inequalities for poly- nomials on the unit square via the Krein -- Milman Theorem, Preprint. [18] D. J. H. Garling, Inequalities: a journey into linear analysis, Cambridge University Press, Cambridge, 2007. [19] S. Kaijser, Some results in the metric theory of tensor products, Studia Math. 63 (1978), no. 2, 157 -- 170. MR511301 (80c:46069) [20] M. Krein and D. Milman, On extreme points of regular convex sets, Studia Math. 9 (1940), 133 -- 138. [21] J. Lindenstrauss and A. Pe lczy´nski, Absolutely summing operators in L p-spaces and their applications, Studia Math. 29 (1968), 275 -- 326. [22] J. E. Littlewood, On bounded bilinear forms in an infinite number of variables, Q. J. Math. 1 (1930), 164 -- 174. [23] G. A. Munoz-Fern´andez, D. Pellegrino, and J. B. Seoane-Sep´ulveda, Estimates for the asymptotic behavior of the constants in the Bohnenblust -- Hille inequality, Linear Multilinear Algebra, In Press. [24] G. A. Munoz-Fern´andez and J. B. Seoane-Sep´ulveda, Geometry of Banach spaces of trinomials, J. Math. Anal. Appl. 340 (2008), no. 2, 1069 -- 1087, DOI 10.1016/j.jmaa.2007.09.010. MR2390911 (2008m:46014) [25] D. Pellegrino and J. B. Seoane-Sep´ulveda, New upper bounds for the constants in the Bohnenblust-Hille in- equality, J. Math. Anal. Appl. 386 (2012), no. 1, 300 -- 307. [26] H. Queff´elec, H. Bohr's vision of ordinary Dirichlet series: old and new results, J. Anal. 3 (1995), 43 -- 60. [27] N. Tomczak-Jaegermann, Banach-Mazur distances and finite-dimensional operator ideals, Pitman Monographs and Surveys in Pure and Applied Mathematics, vol. 38, Longman Scientific & Technical, Harlow, 1989. Departamento de An´alisis Matem´atico, Facultad de Ciencias Matem´aticas, Plaza de Ciencias 3, Universidad Complutense de Madrid, Madrid, 28040, Spain. E-mail address: gustavo [email protected] Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900 - Joao Pessoa, Brazil. E-mail address: [email protected] and [email protected] Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900 - Joao Pessoa, Brazil. E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Ciencias Matem´aticas, Plaza de Ciencias 3, Universidad Complutense de Madrid, Madrid, 28040, Spain. E-mail address: [email protected]
1807.08214
1
1807
2018-07-21T23:57:03
On further refinements for Young inequalities
[ "math.FA" ]
In this paper, sharp results on operator Young's inequality are obtained. We first obtain sharp multiplicative refinements and reverses for the operator Young's inequality. Secondly, we give an additive result, which improves a well-known inequality due to Tominaga. We also provide some estimates for $A{{\sharp}_{v}}B-A{{\nabla }_{v}}B$ in which $v\notin \left[ 0,1 \right]$.
math.FA
math
ON FURTHER REFINEMENTS FOR YOUNG INEQUALITIES SHIGERU FURUICHI AND HAMID REZA MORADI Abstract. In this paper, sharp results on operator Young's inequality are obtained. We first obtain sharp multiplicative refinements and reverses for the operator Young's inequality. Secondly, we give an additive result, which improves a well-known inequality due to Tominaga. We also provide some estimates for A♯vB − A∇vB in which v /∈ [0, 1]. 8 1 0 2 l u J 1 2 ] . A F h t a m [ 1 v 4 1 2 8 0 . 7 0 8 1 : v i X r a 1. Introduction This note lies in the scope of operator inequalities. We assume that the reader is familiar with the continuous functional calculus and Kubo-Ando theory [6]. It is to be understood throughout the paper that the capital letters present bounded linear operators acting on a Hilbert space H. A is positive (written A ≥ 0) in case hAx, xi ≥ 0 for all x ∈ H also an operator A is said to be strictly positive(denoted by A > 0) if A is positive and invertible. If A and B are self-adjoint, we write B ≥ A in case B − A ≥ 0. As usual, by I we denote the identity operator. The weighted arithmetic mean ∇v, geometric mean ♯v, and harmonic mean !v, for v ∈ [0, 1] and a, b > 0, are defined as follows: a∇vb = (1 − v) a + vb, a♯vb = a1−vbv, If v = 1 2, we denote the arithmetic, geometric, and harmonic means, respectively, by ∇, ♯ and !, for the simplicity. Like the scalar cases, the operator arithmetic mean, the operator geometric a!vb =(cid:8)(1 − v)a−1 + vb−1(cid:9)−1 . mean, and the operator harmonic mean for A, B > 0 are defined as follows: A∇vB = (1 − v) A + vB, A♯vB = A 2 BA− 1 A 1 2(cid:16)A− 1 2(cid:17)v 1 2 , A!vB =(cid:8)(1 − v)A−1 + vB−1(cid:9)−1 . The celebrated arithmetic-geometric-harmonic-mean inequalities for scalars assert that if a, b > 0, then (1.1) a!vb ≤ a♯vb ≤ a∇vb. 2010 Mathematics Subject Classification. Primary 47A63, Secondary 26D07, 47A60. Key words and phrases. Operator inequality, Young inequality, arithmetic -- geometric mean inequality, posi- tive operator. 1 2 S. Furuichi & H.R. Moradi Generalization of the inequalities (1.1) to operators can be seen as follows: If A, B > 0, then A!vB ≤ A♯vB ≤ A∇vB. The last inequality above is called the operator Young's inequality. During the past years, several refinements and reverses were given for Young's inequality, see for example [4, 5, 7]. Zuo et al. showed in [9, Theorem 7] that the following inequality holds: (1.2) K(h, 2)rA♯vB ≤ A∇vB, r = min {v, 1 − v} , K (h, 2) = (h + 1)2 4h , h = M m whenever 0 < m′I ≤ B ≤ mI < MI ≤ A ≤ M ′I or 0 < m′I ≤ A ≤ mI < MI ≤ B ≤ M ′I. As the authors mentioned in [9], the inequality (1.2) improves the following refinement of Young's inequality involving Specht's ratio S (t) = t (t > 0, t 6= 1) (see [2, Theorem 2]), t−1 1 1 t−1 e log t Another improvement of Young's inequality, is shown in [1, Corollary 1]: S (hr) A♯vB ≤ A∇vB. A∇vB ≤ exp(cid:20) v (1 − v) 2 (h − 1)2(cid:21) A♯vB. We remark that there is no relationship between the constants K(h, 2)r and exph v(1−v) in general. 2 (h − 1)2i In [3, 5] we proved some sharp multiplicative reverses of Young's inequality. In this brief note, as the continuation of our previous works, we establish sharp bounds for the arithmetic, geometric and harmonic mean inequalities. Moreover, we shall show some additive-type refine- ments and reverses of Young's inequality. We will formulate our new results in a more general setting, namely the sandwich assumption sA ≤ B ≤ tA (0 < s ≤ t). Additionally, we provide some estimates for A♯vB − A∇vB in which v /∈ [0, 1]. 2. Main Results In our previous work [3], we gave new sharp inequalities for reverse Young inequalities. In this section, we firstly give new sharp inequalities for Young inequalities, as limited cases in the first inequalities both (i) and (ii) of the following theorem. Theorem 2.1. Let A, B > 0 such that sA ≤ B ≤ tA for some scalars 0 < s ≤ t and let fv(x) ≡ (1−v)+vx for x > 0, and v ∈ [0, 1]. xv (i) If t ≤ 1, then fv(t)A♯vB ≤ A∇vB ≤ fv(s)A♯vB. (ii) If s ≥ 1, then fv(s)A♯vB ≤ A∇vB ≤ fv(t)A♯vB. On inequalities of Young type 3 Proof. Since f ′ monotone increasing for x ≥ 1. v(x) = v(1 − v)(x − 1)x−v−1, fv(x) is monotone decreasing for 0 < x ≤ 1 and (i) For the case 0 < s ≤ x ≤ t ≤ 1, we have fv(t) ≤ fv(x) ≤ fv(s), which implies fv(t)A♯vB ≤ A∇vB ≤ fv(s)A♯vB by the standard functional calculus. (ii) For the case 1 ≤ s ≤ x ≤ t, we have fv(s) ≤ fv(x) ≤ fv(t) which implies fv(s)A♯vB ≤ A∇vB ≤ fv(t)A♯vB by the standard functional calculus. (cid:3) Remark 2.1. It is worth emphasizing that each assertion in Theorem 2.1, implies the other one. For instance, assume that the assertion (ii) holds, i.e., (2.1) fv (s) ≤ fv (x) ≤ fv (t) , 1 ≤ s ≤ x ≤ t. Let t ≤ 1, then 1 ≤ 1 t ≤ 1 x ≤ 1 s . Hence (2.1) ensures that fv(cid:18)1 t(cid:19) ≤ fv(cid:18) 1 x(cid:19) ≤ fv(cid:18) 1 s(cid:19) . (1 − v) t + v t1−v ≤ ≤ (1 − v) x + v x1−v (1 − v) + vx xv ≤ ≤ (1 − v) s + v s1−v . (1 − v) + vs sv So Now, by replacing v by 1 − v we get (1 − v) + vt tv which means fv (t) ≤ fv (x) ≤ fv (s) , 0 < s ≤ x ≤ t ≤ 1. Corollary 2.1. Let A, B > 0, m, m′, M, M ′ > 0, and v ∈ [0, 1]. (i) If 0 < m′I ≤ A ≤ mI < MI ≤ B ≤ M ′I, then (2.2) m∇vM m♯vM A♯vB ≤ A∇vB ≤ m′∇vM ′ m′♯vM ′ A♯vB. (ii) If 0 < m′I ≤ B ≤ mI < MI ≤ A ≤ M ′I, then (2.3) M∇vm M♯vm A♯vB ≤ A∇vB ≤ M ′∇vm′ M ′♯vm′ A♯vB. Proof. We use again the function fv(x) = (1−v)+vx m I ≤ A− 1 m and t = M ′ The condition (i) is equivalent to I ≤ M m′ )A♯vB by putting s = M A∇vB ≤ fv( M ′ in this proof. 2 BA− 1 2 ≤ M ′ m′ in (ii) of Theorem 2.1. xv m′ I, so that we get fv( M m )A♯vB ≤ Similarly, the condition (ii) is equivalent to m′ M )A♯vB ≤ A∇vB ≤ fv( m′ M ′ )A♯vB by putting s = m′ 2 BA− 1 M ′ and t = m M ′ I ≤ A− 1 M I ≤ I, so that we get 2 ≤ m M in (i) of Theorem 2.1. fv( m Note that the second inequalities in both (i) and (ii) of Theorem 2.1 and Corollary 2.1 are special cases in Theorem A of our previous paper [3]. (cid:3) 4 S. Furuichi & H.R. Moradi Remark 2.2. It is remarkable that the inequalities fv(t) ≤ fv(x) ≤ fv(s) (0 < s ≤ x ≤ t ≤ 1) given in the proof of Theorem 2.1 are sharp, since the function fv(x) for s ≤ x ≤ t is continuous. So, all result given from Theorem 2.1 are similarly sharp. As a matter of fact, let A = MI and B = mI, then from LHS of (2.3), we infer A∇vB = (M∇vm)I and A♯vB = (M♯vm)I. Consequently, M∇vm M♯vm A♯vB = A∇vB. To see that the constant m∇vM and B = MI, then m♯v M in the LHS of (2.2) can not be improved, we consider A = mI m∇vM m♯vM A♯vB = A∇vB. By replacing A, B by A−1, B−1, respectively, then the refinement and reverse of non- commutative geometric-harmonic mean inequality can be obtained as follows: Corollary 2.2. Let A, B > 0, m, m′, M, M ′ > 0, and v ∈ [0, 1]. (i) If 0 < m′I ≤ A ≤ mI < MI ≤ B ≤ M ′I, then m′!vM ′ m′♯vM ′ A♯vB ≤ A!vB ≤ (ii) If 0 < m′I ≤ B ≤ mI < MI ≤ A ≤ M ′I, then m!vM m♯vM A♯vB. M ′!vm′ M ′♯vm′ A♯vB ≤ A!vB ≤ M!vm M♯vm A♯vB. Now, we give a new sharp reverse inequality for Young's inequality as an additive-type in the following. Theorem 2.2. Let A, B > 0 such that sA ≤ B ≤ tA for some scalars 0 < s ≤ t, and v ∈ [0, 1]. Then (2.4) A∇vB − A♯vB ≤ max {gv (s) , gv (t)} A where gv (x) ≡ (1 − v) + vx − xv for s ≤ x ≤ t. Proof. Straightforward differentiation shows that g′′ tinuous on the interval [s, t], so v (x) = v(1 − v)xv−2 ≥ 0 and gv(x) is con- gv (x) ≤ max {gv (s) , gv (t)} . Therefore, by applying similar arguments as in the proof of Theorem 2.1, we reach the desired inequality (2.4). This completes the proof of theorem. (cid:3) On inequalities of Young type 5 Remark 2.3. We claim that if A, B > 0 such that mI ≤ A, B ≤ MI for some scalars 0 < m < M and v ∈ [0, 1], then A∇vB − A♯vB ≤ max(cid:26)gv (h) , gv(cid:18) 1 h(cid:19)(cid:27) A ≤ L (1, h) log S (h) A holds, where L (x, y) = y−x the Specht's ratio. Indeed, we have the inequalities log y−log x (x 6= y) is the logarithmic mean and the term S (h) refers to (1 − v) + vh − hv ≤ L(1, h) log S(h), (1 − v) + v 1 h − h−v ≤ L(1, h) log S(h), which were originally proved in [8, Lemma 3.2], thanks to S (h) = S(cid:0) 1 Therefore, our result, Theorem 2.2, improves the well-known result by Tominaga [8, Theorem h(cid:1) and L (1, h) = L(cid:0)1, 1 h(cid:1). 3.1], A∇vB − A♯vB ≤ L (1, h) log S (h) A. Corollary 2.3. Let A, B > 0 such that mI ≤ A, B ≤ MI for some scalars 0 < m < M. Then A∇vB − A♯vB ≤ ξA where ξ = max(cid:8) 1 M (M∇vm − M♯vm) , 1 m (m∇vM − m♯vM)(cid:9). Since gv(x) is convex so that we can not obtain a general result on the lower bound for A∇vB − A♯vB. However, if we impose the conditions, we can obtain new sharp inequalities for Young inequalities as an additive-type in the first inequalities both (i) and (ii) in the following proposition. (At the same time, of course, we also obtain the upper bounds straightforwardly.) Proposition 2.1. Let A, B > 0 such that sA ≤ B ≤ tA for some scalars 0 < s ≤ t, v ∈ [0, 1], and gv is defined as in Theorem 2.2. (i) If t ≤ 1, then gv (t) A ≤ A∇vB − A♯vB ≤ gv(s)A. (ii) If s ≥ 1, then gv (s) A ≤ A∇vB − A♯vB ≤ gv(t)A. Proof. It follows from the fact that gv (x) is monotone decreasing for 0 < x ≤ 1 and monotone increasing for x ≥ 1. (cid:3) Corollary 2.4. Let A, B > 0, m, m′, M, M ′ > 0, and v ∈ [0, 1]. (i) If 0 < m′I ≤ A ≤ mI < MI ≤ B ≤ M ′I, then 1 m (m∇vM − m♯vM) A ≤ A∇vB − A♯vB ≤ 1 m′ (m′∇vM ′ − m′♯vM ′) A. (ii) If 0 < m′I ≤ B ≤ mI < MI ≤ A ≤ M ′I, then 1 M (M∇vm − M♯vm) A ≤ A∇vB − A♯vB ≤ 1 M ′ (M ′∇vm′ − M ′♯vm′) A. 6 S. Furuichi & H.R. Moradi Remark 2.4. It is known that for any A, B > 0, A∇vB ≤ A♯vB for v /∈ [0, 1] . Assume gv (x) is defined as in Theorem 2.2. By an elementary computation we have ( g ′ v (x) > 0 g ′ v (x) < 0 for v /∈ [0, 1] and 0 < x ≤ 1 for v /∈ [0, 1] and x > 1 . Now, in the same way as above we have also for any v /∈ [0, 1]: (i) If 0 < m′I ≤ A ≤ mI ≤ MI ≤ B ≤ M ′I, then 1 m′ (m′♯vM ′ − m′∇vM ′) A ≤ A♯vB − A∇vB ≤ 1 m (m♯vM − m∇vM) A. On account of assumptions, we also infer (m′♯vM ′ − m′∇vM ′)I ≤ A♯vB − A∇vB ≤ (m♯vM − m∇vM)I. (ii) If 0 < m′I ≤ B ≤ mI ≤ MI ≤ A ≤ M ′I, then 1 M (M♯vm − M∇vm) A ≤ A♯vB − A∇vB ≤ 1 M ′ (M ′♯vm′ − M ′∇vm′) A. On account of assumptions, we also infer (M♯vm − M∇vm)I ≤ A♯vB − A∇vB ≤ (M ′♯vm′ − M ′∇vm′)I. In addition, with the same assumption to Theorem 2.2 except for v /∈ [0, 1], we have min{gv(s), gv(t)}A ≤ A∇vB − A♯vB, since we have min{gv(s), gv(t)} ≤ gv(x) by g′′ v (x) ≤ 0, for v /∈ [0, 1]. Acknowledgement The author (S.F.) was partially supported by JSPS KAKENHI Grant Number 16K05257. References [1] S.S. Dragomir, On new refinements and reverse of Young's operator inequality, arXiv:1510.01314v1. [2] S. Furuichi, Refined Young inequalities with Specht's ratio, J. Egyptian Math. Soc., 20(1) (2012), 46 -- 49. [3] S. Furuichi, H.R. Moradi and M. Sababheh, New sharp inequalities for operator means, Linear Multilinear Algebra., https://doi.org/10.1080/03081087.2018.1461189 [4] S. Furuichi and H.R. Moradi, On some refinements for mean inequalities, To appear in Rocky Mountain J. Math.https://projecteuclid.org/euclid.rmjm/1524880819 [5] I.H. Gumu¸s, H.R. Moradi and M. Sababheh, More accurate operator means inequalities, J. Math. Anal. Appl., 465(1) (2018), 267 -- 280. [6] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann., 246 (1980), 205 -- 224. [7] W. Liao, J. Wu and J. Zhao, New versions of reverse Young and Heinz mean inequalities with the Kantorovich constant, Taiwanese J. Math., 19(2) (2015), 467 -- 479. On inequalities of Young type 7 [8] M. Tominaga, Specht's ratio in the Young inequality, Sci. Math. Jpn., 5 (2001), 525 -- 530. [9] H. Zuo, G. Shi and M. Fujii, Refined Young inequality with Kantorovich constant, J. Math. Inequal., 5(4) (2011), 551 -- 556. (S. Furuichi) Department of Information Science, College of Humanities and Sciences, Nihon University, 3-25-40, Sakurajyousui, Setagaya-ku, Tokyo, 156-8550, Japan. E-mail address: [email protected] (H.R. Moradi) Young Researchers and Elite Club, Mashhad Branch, Islamic Azad University, Mashhad, Iran E-mail address: [email protected]
1102.5082
1
1102
2011-02-24T20:34:54
On metric characterizations of some classes of Banach spaces
[ "math.FA", "math.MG" ]
The paper contains the following results and observations: (1) There exists a sequence of unweighted graphs $\{G_n\}_n$ with maximum degree 3 such that a Banach space $X$ has no nontrivial cotype iff $\{G_n\}_n$ admit uniformly bilipschitz embeddings into $X$; (2) The same for Banach spaces with no nontrivial type; (3) A sequence $\{G_n\}$ characterizing Banach spaces with no nontrivial cotype in the sense described above can be chosen to be a sequence of bounded degree expanders; (4) The infinite diamond does not admit a bilipschitz embedding into Banach spaces with the Radon-Nikod\'{y}m property; (5) A new proof of the Cheeger-Kleiner result: The Laakso space does not admit a bilipschitz embedding into Banach spaces with the Radon-Nikod\'{y}m property; (6) A new proof of the Johnson-Schechtman result: uniform bilipschitz embeddability of finite diamonds into a Banach space implies its nonsuperreflexivity.
math.FA
math
On metric characterizations of some classes of Banach spaces Mikhail I. Ostrovskii May 29, 2018 Abstract. The paper contains the following results and observations: (1) There exists a sequence of unweighted graphs {Gn}n with maximum degree 3 such that a Banach space X has no nontrivial cotype iff {Gn}n admit uniformly bilipschitz embeddings into X; (2) The same for Banach spaces with no nontrivial type; (3) A sequence {Gn} characterizing Banach spaces with no nontrivial cotype in the sense described above can be chosen to be a sequence of bounded degree expanders; (4) The infinite diamond does not admit a bilipschitz embedding into Banach spaces with the Radon-Nikod´ym property; (5) A new proof of the Cheeger-Kleiner result: The Laakso space does not admit a bilipschitz embedding into Banach spaces with the Radon-Nikod´ym property; (6) A new proof of the Johnson-Schechtman result: uniform bilipschitz embeddability of finite diamonds into a Banach space implies its nonsuperreflexivity. Keywords. Banach space, diamond graphs, expander graphs, Laakso graphs, Lipschitz embedding, Radon-Nikod´ym property 2010 Mathematics Subject Classification: Primary: 46B85; Secondary: 05C12, 46B07, 46B22, 54E35 1 Introduction A mapping F : X → Y between metric spaces X and Y is called a C-bilipschitz embedding if there exists r > 0 such that ∀u, v ∈ X rdX(u, v) ≤ dY (F (u), F (v)) ≤ rCdX(u, v). A sequence {fn} of mappings fn : Un → Zn between metric spaces is a sequence of uniformly bilipschitz embeddings if there is C < ∞ such that all of the embeddings are C-bilipschitz. Many important classes of Banach spaces have been characterized in terms of uniformly bilipschitz embeddings of finite metric spaces. Bourgain [Bou86] proved that a Banach space X is nonsuperreflexive if and only if there exist uniformly bilipschitz embeddings of finite binary trees {Tn}∞ n=1 of all depths into X. Similar characterization of spaces with no type > 1 was obtained by Bourgain, Milman, and Wolfson [BMW86]: a Banach space X has no type > 1 if and only if there exist uniformly bilipschitz embeddings of Hamming cubes {Hn}∞ n=1 into X (see [Pis86] for a simpler proof). Johnson and Schechtman [JS09] found a similar characterization of nonsuperreflexive spaces in terms of diamond graphs [GNRS04] and Laakso graphs [Laa00]. Banach spaces without cotype were character- ized by Mendel and Naor [MN08] in terms of lattice graphs Lm,n whose vertex sets are {0, 1, . . . , m}n, two vertices are joined by an edge if and only if their ℓ∞-distance is equal to 1. Characterizations in terms of bilipschitz embeddability of certain metric spaces are called metric characterizations. Observe that binary trees and Laakso graphs are graphs 1 with uniformly bounded degrees. Degrees of Hamming cubes and the lattice graphs are unbounded. During the seminar "Nonlinear geometry of Banach spaces" (Workshop in Analysis and Probability at Texas A & M University, 2009) Johnson posed the following problem: Find metric characterizations of spaces with no type p > 1 and with no cotype in terms of graphs with uniformly bounded degrees. The first part of this paper is devoted to a solution of this problem. In the second part of the paper we prove results related to another problem posed by Johnson during the mentioned seminar: Find metric characterizations of reflexivity and the Radon-Nikod´ym property (RNP). We prove that Banach spaces containing bilipschitz images of the infinite diamond do not have the RNP, but the converse is not true. We find a new proof of the Cheeger-Kleiner [CK09] result that Banach spaces containing bilipschitz images of the Laakso space do not have the RNP. The author would like to thank Florent Baudier, William B. Johnson, Mikhail M. Po- pov, Beata Randrianantoanina, and Gideon Schechtman for their interest to this work and for useful related discussions. 2 Type and cotype in terms of graphs with uniformly bounded degrees Theorem 2.1. There exist metric characterizations of the classes of spaces with no type > 1 and with no cotype in terms of graphs with maximum degree 3. We start with the characterization of cotype. This case is easier, because the well- known results ([Mat02, Proposition 15.6.1] and [MP76]) imply that any family of finite metric spaces is uniformly bilipschitz embeddable into any Banach space with no cotype. Therefore to prove the cotype part of the theorem we need to show only that the graphs Lm,n are uniformly bilipschitz embeddable into a family of graphs with uniformly bounded degrees. Thus it suffices to prove the following lemma. Lemma 2.2. Let (V (G), dG) be the vertex set of a graph G with its graph distance, and let ε > 0. Then there exist a graph M with maximum degree ≤ 3, ℓ ∈ N, and an embedding F : V (G) → V (M) such that ℓdG(u, v) ≤ dM (F (u), F (v)) ≤ (1 + ε)ℓdG(u, v) ∀u, v ∈ V (G), where dM is the graph distance of M . Proof. Let ∆G be the maximum degree of G and let r ∈ N be such that 3 · 2r−1 ≥ ∆G. Let ℓ ∈ N be such that ℓ+2r ℓ < 1 + ε. We define the graph M in the following way. For each vertex v of G the graph M contains a 3-regular tree of depth r rooted at a vertex which we denote m(v). For each edge uv of G we pick a leaf of the tree rooted at m(v) and a leaf of the tree rooted at m(u) and join them by a path of length ℓ. Leaves picked for different edges are different (this is possible because 3 · 2r−1 ≥ ∆G), and there is no further interaction between the constructed trees and paths. It is easy to see that the maximum degree of M is 3. 2 We map V (G) into V (M) by mapping each v to the corresponding m(v). It remains to show that ℓdG(u, v) ≤ dM (m(u), m(v)) ≤ (ℓ + 2r)dG(u, v) (1) The right-hand side inequality follows from the observation that if u and v are adjacent in G, then dM (m(u), m(v)) ≤ ℓ + 2r, the path of length ℓ + 2r can be constructed as the union of path in M corresponding to uv, and paths from m(u) and m(v) to the corresponding leaves. To prove the left-hand side of (1) we consider a path joining m(u) and m(v). Let m(u1), . . . , m(uk) be the set of roots of those trees which are visited by the path, listed in the order of visits. The description of M implies that u, u1, . . . , uk, v is a uv-walk in G, hence its length is ≥ dG(u, v). In order to get from one tree to another we need to traverse ℓ edges. Hence any path joining m(u) and m(v) has length ≥ ℓdG(u, v). This completes the proof of the lemma and the cotype part of the theorem. Proof of the type part of Theorem 2.1. By [MP76, Theorem 2.3], each space with no type > 1 contains subspaces whose Banach-Mazur distances to ℓd 1 (d ∈ N) are arbitrarily close to 1. Therefore it suffices to check that each of the graphs obtained in a similar way from Hamming cubes admits a uniformly bilipschitz embedding into ℓd 1 for sufficiently large n=1 using the procedure described d. We denote by {Sn}∞ in the proof of Lemma 2.2. We describe an embedding of the vertex set of Sn into ℓk 1. The images of vertices of Sn under this embedding are integer points of ℓk 1, edges of Sn correspond to line segments of length 1 parallel to unit vectors of ℓk 1. Having such a representation of Sn, it remains to show that the identity mappings of vertex sets of Sn endowed with their graph distances and their ℓ1-distances are uniformly bilipschitz. n=1 graphs obtained from {Hn}∞ The graph Hn is n-regular, so we let r ∈ N be such that n ≤ 3 · 2r−1 and consider a rooted 3-regular tree of depth r. This tree can be isometrically embedded into ℓm 1 , where m = 3 + 3 · 2 + · · · + 3 · 2r−1. The embedding is the following: observing that m is the number of edges in the tree, we find a bijection between unit vectors in ℓm 1 and edges of the tree. Now we map the root of the tree to 0 ∈ ℓm 1 ; if v is different from the root, we map v to the sum of unit vectors corresponding to the path from v to the root. We denote by Tr the image of the tree in ℓm 1 . We consider the natural isometric embedding of Hn into ℓn 1 , with images of the vertices being all possible 0, 1-sequences. We pick ℓ in the same way as in Lemma 2.2. We specify the position of the rooted tree corresponding to the vertex v = {θi}n i=1 of Hn in 1 = ℓm ℓk 1 and ℓ · {θi} is a multiple of v considered as a vector in ℓn 1 . 1 ⊕1 ℓn 1 as Tr + ℓ · {θi}, where we mean that Tr ⊂ ℓm We introduce and embed the paths of length ℓ (from the construction of Lemma 2.2) in the following way: Since n is ≤ the number of leaves in Tr, there is a bijection between the unit vectors of ℓn 1 and some subset of leaves of Tr. On the other hand, each edge of Hn is parallel to one of the unit vectors. We add ℓ-paths in the following way. The path corresponding to the edge between v and v + et (et is a unit vector of ℓn 1 ) is the straight 3 line path of length ℓ joining the leaves of Tr + ℓv and Tr + ℓ(v + et); in each of the trees the leaf is chosen in such a way that its ℓm 1 component is the leaf corresponding to et. It is clear that the graph Sn obtained in this way fits the description of M in the proof of Lemma 2.2. Therefore the natural embedding of Hn into Sn is (1 + ε)-bilipschitz if both graphs are endowed with their graph distances. It remains to estimate the bilipschitz constants of natural embeddings of Sn into ℓk 1. Observe that the graph distance between two vertices of Sn cannot be less than the distance between their images in ℓn 1 , because each edge corresponds to a line segment of length 1. It remains to show that the graph distance between two vertices of Sn cannot be much larger than the ℓ1-distance. Let x, y be two vertices of Sn, we need to estimate dSn(x, y) from above in terms of x − y1. 1 ⊕1 ℓm For each set of vertices of the form Tr + ℓv we consider its union with the set of all vertices of ℓ-paths going out of this set. It is easy to see that if both x and y belong to one of such sets, then dSn(x, y) ≤ x − y1. For x ∈ V (Sn) denote the projection of x to ℓn 1 by π(x), and the i-th coordinate of this projection by π(x)i. If the situation described in the previous paragraph does not occur then there exists i ∈ {1, . . . , n} such that π(x)i − π(y)i = ℓ. Let k ≤ n be the number of coordinates for which this equality holds. We have x − y1 ≥ π(x) − π(y)1 ≥ kℓ. To estimate dSn(x, y) from above we construct the following xy-path in Sn. If one of the numbers π(x)i is strictly between 0 and ℓ we start by moving from x in the direction of π(y)i (which in this case should be 0 or ℓ) till we reach a set of the form Tr + ℓwx for some vertex wx of Hn. We do similar thing at the other end of the path (near y): If one of the numbers π(y)i is strictly between 0 and ℓ we end the path by moving from y in the direction of π(x)i (which in this case should be 0 or ℓ) till we reach a set of the form Tr + ℓwy. We find a shortest path between wx and wy in Hn. It is easy to see that it has length k. Now we continue construction of the xy-path in Sn. This path will contain all paths of length ℓ corresponding to the edges of the wxwy-path in Hn. Between these paths we add the pieces of the corresponding trees of the from Tr + ℓu, needed to make a path. As a result we get an xy-path of length < 2ℓ + kℓ + 2(k + 1)r. If 4r ≤ ℓ (we can definitely assume this), we have 2ℓ + kℓ + 2(k + 1)r ≤ 4kℓ. In such a case dSn(x, y) ≤ 4x − y1. This completes the proof of the type part of the theorem. Corollary 2.3. There exists a family {Kn} of constant degree expanders, such that a Banach space X for which there exist uniformly bilipschitz embeddings of {Kn} into X, has no cotype. Proof. Let {Mn} be graphs of maximum degree 3 from the metric characterization of Banach spaces with no cotype. It suffices to show that there exists a family {Kn} of constant degree expanders containing subsets isometric to {Mn}. Consider any family {Gk} of constant d-regular expanders with the growing number of vertices. Let Dn be the diameter of Mn and mn be its number of vertices. It is clear that we may assume 4 without loss of generality that Gn contains a Dn-separated set of cardinality mn. We fix a bijection between this set and V (Mn). We add to Gn edges between vertices corresponding to adjacent vertices of Mn. Since Dn is the diameter of Mn, the obtained graph contains an isometric copy of Mn. The maximum degree of the obtained graph is ≤ d + 3. Its expanding properties are not worse than those of Gn. Adding as many self-loops to it as is needed we get a (d + 3)-regular graph Kn. It is clear that {Kn} is a desired family of (d + 3)-regular expanders. 3 Diamonds and Laakso graphs The diamond graph of level 0 is denoted D0. It has two vertices joined by an edge of length 1. Di is obtained from Di−1 as follows. Given an edge uv ∈ E(Di−1), it is replaced by a quadrilateral u, a, v, b with edge lengths 2−i. We endow Dn with their shortest path metrics. We consider the vertex of Dn as a subset of the vertex set of Dn+1, it is easy to check that this defines an isometric embedding. We introduce Dω as the union of the n=0. For u, v ∈ Dω we introduce dDω(u, v) as dDn(u, v) where n ∈ N vertex sets of {Dn}∞ is any integer for which u, v ∈ V (Dn). Since the natural embeddings Dn → Dn+1 are isometric, it is easy to see that dDn(u, v) does not depend on the choice of n for which u, v ∈ V (Dn). Definition 3.1 ([Jam72] or [Bou83, p. 34]). Let δ > 0. A sequence {xi}∞ δ-tree if xi = 1 2(x2i + x2i+1) and x2i − xi = x2i+1 − xi ≥ δ. i=1 is called a Theorem 3.2. If Dω is bilipschitz embeddable into a Banach space X, then X contains a bounded δ-tree for some δ > 0. It is well-known that Banach spaces with the RNP do not contain bounded δ-trees (see [Bou83, p. 31]). On the other hand there exist Banach spaces without the RNP which do not contain bounded δ-trees, see [BR80, p. 54]. So Theorem 3.2 implies: Corollary 3.3. If Dω is bilipschitz embeddable into a Banach space X, then X does not have the Radon-Nikod´ym property. The converse is not true. Proof of Theorem 3.2. Let f : Dω → X be a bilipschitz embedding. Without loss of generality we assume that δdDω (x, y) ≤ f (x) − f (y) ≤ dDω(x, y) (2) for some δ > 0. Let us show that this implies that the unit ball of X contains a δ-tree. The first element of the tree will be x1 = f (u0) − f (v0), where {u0, v0} = V (D0). Now we consider the quadrilateral u0, a, v0, b. Inequality (2) implies f (a) − f (b) ≥ δ. Consider two pairs of vectors (corresponding to two different paths from u to v in D1): Pair 1: f (v0) − f (a), f (a) − f (u0). Pair 2: f (v0) − f (b), f (b) − f (u0). 5 The inequality f (a) − f (b) ≥ δ implies that at least one of the following is true (f (v0) − f (a)) − (f (a) − f (u0)) ≥ δ or (f (v0) − f (b)) − (f (b) − f (u0)) ≥ δ. Suppose that the first inequality holds. We let x2 = 2(f (v0) − f (a)) and x3 = 2(f (a) − f (u0)). It is clear that both conditions of Definition 3.1 are satisfied. Also, the condition (2) implies that x2, x3 ≤ 1. We continue construction of the δ-tree in the unit ball of X in a similar manner. For example, to construct x4 and x5 we consider the corresponding quadrilateral a, a1, v0, b1 in D2. The inequality f (a1) − f (b1) ≥ δ/2 implies that at least one of the following is true (f (v0) − f (a1)) − (f (a1) − f (a)) ≥ δ/2 or (f (v0) − f (b1)) − (f (b1) − f (a)) ≥ δ/2. Suppose that the second inequality holds. We let x4 = 4(f (v0) − f (b1)) and x5 = 4(f (b1) − f (a)). It is clear that both conditions of Definition 3.1 are satisfied. Also (2) implies that x4, x5 ≤ 1. Proceeding in an obvious way we get a δ-tree in the unit ball of X. 3.1 Finite version and the Johnson-Schechtman characterization of super- reflexivity Definition 3.4 ([Jam72]). A Banach space X has the finite tree property if there exist δ > 0 such that for each k ∈ N the unit ball of X contains a finite sequence {xi : i = 1, . . . , 2k − 1} such that xi = 1 2(x2i + x2i+1) and x2i − xi = x2i+1 − xi ≥ δ for each i = 1, . . . , 2k−1 − 1. It is clear that the proof of Theorem 3.2 implies its finite version: Corollary 3.5. If there exist uniformly bilipschitz embeddings of {Dn}∞ space X, then X has the finite tree property. n=1 into a Banach Combining Corollary 3.5 with the well-known fact (see [Jam72] and [Enf72]) that the finite tree property is equivalent to nonsuperreflexivity, we get the second part of the n=1 into X implies the result in [JS09, p. 181]: uniform bilipschitz embeddability of {Dn}∞ nonsuperreflexivity of X. 3.2 Laakso space Our version of the Laakso space (originally constructed in [Laa00]) is similar to the version from [LP01, p. 290]. However, our version is a countable set (dense in the version of the space from [LP01]). The Laakso graph of level 0 is denoted L0. It consists of two vertices joined by an edge of length 1. The Laakso graph Li is obtained from Li−1 as follows. Each 6 edge uv ∈ E(Li−1) of length 4−i+1 is replaced by a graph with 6 vertices u, t1, t2, o1, o2, v where o1, t1, o2, t2 form a quadrilateral, and there are only two more edges ut1 and vt2, with all edge lengths 4−i. We endow Ln with their shortest path metrics. We consider the vertex of Ln as a subset of the vertex set of Ln+1, it is easy to check that this defines an isometric embedding. We introduce the Laakso space Lω as the union of the vertex sets n=0. For u, v ∈ Lω we introduce dLω(u, v) as dLn(u, v) where n ∈ N is any integer of {Ln}∞ for which u, v ∈ V (Ln). Since the natural embeddings Ln → Ln+1 are isometric, it is easy to see that dLn(u, v) does not depend on the choice of n for which u, v ∈ V (Ln). Our next purpose is to give a new proof of the following result of Cheeger and Kleiner [CK09, Corollary 1.7]: Theorem 3.6. If Lω is bilipschitz embeddable into a Banach space X, then X does not have the Radon-Nikod´ym property. Proof. We do not know whether bilipschitz embeddability of Lω into X implies the ex- istence of a bounded δ-tree in X. To prove Theorem 3.6 we introduce the following definition. Definition 3.7. Let δ > 0. A sequence {xi}∞ x4i−1 + x4i + x4i+1) and (x4i−2 + x4i−1) − (x4i + x4i+1) ≥ δ. i=1 is called a δ-semitree if xi = 1 4 (x4i−2 + Our proof has two steps. First we show that bilipschitz embeddability of Lω into X implies that X contains a bounded δ-semitree. The second step is to show that existence of a bounded δ-semitree in X implies that X does not have the RNP (this is almost standard, based on martingales). Let f : Lω → X be a bilipschitz embedding. Without loss of generality we assume that for some δ > 0. δdLω (x, y) ≤ f (x) − f (y) ≤ dLω(x, y) (3) We need to construct a δ-semitree in the unit ball of X. The first element of the semitree is x1 = f (u0) − f (v0), where {u0, v0} = V (L0). Now we consider the 4-tuple u0, o1, v0, o2. Observe that (3) together with dLω (o1, o2) ≥ 1/2 implies that o1 − o2 ≥ δ/2. Consider two pairs of vectors: Pair 1: f (v0) − f (o1), f (o1) − f (u0). Pair 2: f (v0) − f (o2), f (o2) − f (u0). The inequality f (o1) − f (o2) ≥ δ/2 implies that at least one of the following is true (f (v0) − f (o1)) − (f (o1) − f (u0)) ≥ δ/2 or (f (v0) − f (o2)) − (f (o2) − f (u0)) ≥ δ/2. Suppose that the first inequality holds. We let x2 = 4(f (v0) − f (t2)), x3 = 4(f (t2) − f (o1)), x4 = 4(f (o1) − f (t1)), x5 = 4(f (t1) − f (u0)). It is easy to check that both conditions of Definition 3.7 are satisfied, we even get (x2 + x3) − (x4 + x5) = 4(f (v0) − f (o1)) − (f (o1) − f (u0)) ≥ 2δ. 7 Also, (3) applied to dLω(u0, t1) = dLω(t1, o1) = dLω (o1, t2) = dLω(t2, v0) = 1/4 implies that x2, x3, x4, x5 ≤ 1. We continue our construction of the δ-semitree in the unit ball of X in a similar manner. For example, to construct x6, x7, x8, and x9, we consider the 6-tuple corresponding to the edge t2v0 of L1 and repeat the same procedure as above for u0v0. Proceeding in an obvious way we get a δ-semitree in the unit ball of X. To show that presence of a bounded δ-semitree implies absence of the RNP we use the same argument as for ε-bushes in [BL00, p. 111]. We construct an X-valued martingale {fn}∞ n=0 on [0, 1]. We let f0 = x1. The function f2 is defined on four quarters of [0, 1] by x2, x3, x4, x5, respectively. To define the function f3 we divide [0, 1] into 16 equal subintervals, and define f3 as x6, . . . , x21, on the respective subintervals, etc. It is clear that we get a sequence of uniformly bounded functions. The first condition in the definition of a δ-semitree implies that this sequence is a martingale. The second condition implies that it is not convergent almost everywhere because it shows that on each interval of the form (cid:2) k 4n (cid:3) the average value of fn − fn+1 over the first half of the interval is ≥ δ/4, this implies that fn(t) − fn+1(t) ≥ δ/4 on a subset in [0, 1] of measure ≥ 1 2 . It remains to apply [BL00, Theorem 5.8]. 4n , k+1 References [BL00] [Bou86] Y. Benyamini, J. Lindenstrauss, Geometric Nonlinear Functional Analysis, volume 1, American Math- ematical Society, Providence, R.I., 2000. J. Bourgain, The metrical interpretation of superreflexivity in Banach spaces, Israel J. Math., 56 (1986), no. 2, 222–230. [BMW86] J. Bourgain, V. Milman, H. Wolfson, On type of metric spaces, Trans. Amer. Math. Soc., 294 (1986), [BR80] [Bou83] [CK09] [Enf72] no. 1, 295–317. J. Bourgain and H. P. Rosenthal, Martingales valued in certain subspaces of L1, Israel J. Math., 37 (1980), no. 1-2, 54–75. R. D. Bourgin, Geometric aspects of convex sets with the Radon-Nikod´ym property, Lecture Notes in Mathematics, 993, Springer-Verlag, Berlin, 1983. J. Cheeger, B. Kleiner, Differentiability of Lipschitz maps from metric measure spaces to Banach spaces with the Radon-Nikod´ym property, Geom. Funct. Anal., 19 (2009), no. 4, 1017–1028. P. Enflo, Banach spaces which can be given an equivalent uniformly convex norm, in: Proceedings of the International Symposium on Partial Differential Equations and the Geometry of Normed Linear Spaces (Jerusalem, 1972), Israel J. Math., 13 (1972), 281–288. [GNRS04] A. Gupta, I. Newman, Y. Rabinovich, A. Sinclair, Cuts, trees and ℓ1-embeddings of graphs, Combi- natorica, 24 (2004) 233–269; Conference version in: 40th Annual IEEE Symposium on Foundations of Computer Science, 1999, pp. 399–408. [Jam72] [JS09] [Laa00] R. C. James, Some self-dual properties of normed linear spaces, in: Symposium on Infinite-Dimensional Topology (Louisiana State Univ., Baton Rouge, La., 1967), pp. 159–175. Ann. of Math. Studies, No. 69, Princeton Univ. Press, Princeton, N. J., 1972. W. B. Johnson, G. Schechtman, Diamond graphs and super-reflexivity, Journal of Topology and Anal- ysis, 1 (2009), 177–189. T. J. Laakso, Ahlfors Q-regular spaces with arbitrary Q > 1 admitting weak Poincare inequality, Geom. Funct. Anal., 10 (2000), no. 1, 111–123. 8 [LP01] U. Lang, C. Plaut, Bilipschitz embeddings of metric spaces into space forms, Geom. Dedicata, 87 (2001), 285–307. [Mat02] J. Matousek, Lectures on Discrete Geometry, Springer-Verlag, New York, 2002. [MP76] B. Maurey, G. Pisier, S´eries de variables al´eatoires vectorielles ind´ependantes et propri´et´es g´eom´e- triques des espaces de Banach, Studia Math., 58 (1976), no. 1, 45–90. [MN08] M. Mendel, A. Naor, Metric cotype, Ann. Math., 168 (2008), 247–298. [Pis86] G. Pisier, Probabilistic methods in the geometry of Banach spaces, (Varenna, 1985), 167–241, Lecture Notes in Math., 1206, Springer, Berlin, 1986. in: Probability and analysis Department of Mathematics and Computer Science St. Johns University 8000 Utopia Parkway, Queens, NY 11439, USA e-mail: [email protected] 9
1302.7040
2
1302
2013-04-04T01:20:44
On matrix inequalities between the power means: counterexamples
[ "math.FA" ]
We prove that the known sufficient conditions on the real parameters $(p,q)$ for which the matrix power mean inequality $((A^p+B^p)/2)^{1/p}\le((A^q+B^q)/2)^{1/q}$ holds for every pair of matrices $A,B>0$ are indeed best possible. The proof proceeds by constructing $2\times2$ counterexamples. The best possible conditions on $(p,q)$ for which $\Phi(A^p)^{1/p}\le\Phi(A^q)^{1/q}$ holds for every unital positive linear map $\Phi$ and $A>0$ are also clarified.
math.FA
math
On matrix inequalities between the power means: counterexamples Koenraad M.R. Audenaert1,∗ and Fumio Hiai2,† 1 Department of Mathematics, Royal Holloway, University of London, Egham TW20 0EX, United Kingdom 2 Tohoku University (Emeritus), Hakusan 3-8-16-303, Abiko 270-1154, Japan Abstract We prove that the known sufficient conditions on the real parameters (p, q) for which the matrix power mean inequality ((Ap + Bp)/2)1/p ≤ ((Aq + Bq)/2)1/q holds for every pair of matrices A, B > 0 are indeed best possible. The proof proceeds by constructing 2× 2 counterexamples. The best possible conditions on (p, q) for which Φ(Ap)1/p ≤ Φ(Aq)1/q holds for every unital positive linear map Φ and A > 0 are also clarified. 2010 Mathematics Subject Classification: Primary 15A45, 47A64 Key Words and Phrases: Matrix, operator, power mean, Jensen inequality, sym- metric norm, joint convexity 1 Introduction For each n ∈ N we write Mn for the n × n complex matrix algebra and Pn for the set of positive definite matrices in Mn. For each non-zero real parameter p and for every A, B ∈ Pn, the p-power mean of A, B is (cid:18)Ap + Bp 2 (cid:19)1/p , (1.1) ∗E-mail: [email protected] †E-mail: [email protected] 1 which is also defined for positive invertible operators on an arbitrary Hilbert space. In particular, it is the arithmetic mean when p = 1, and it is the harmonic mean when p = −1. Moreover, when p = 0, it is defined by continuity as = exp(cid:18) log A + log B p→0(cid:18)Ap + Bp (cid:19)1/p lim 2 2 (cid:19), (1.2) which is the so-called Log-Euclidean mean, a kind of geometric mean but different from that in the sense of operator means [11]. In fact, (1.1) is not an operator mean except when p = ±1. In this paper we are concerned with conditions on p and q for the validity of the matrix inequality between the power means (cid:18)Ap + Bp 2 (cid:19)1/p ≤(cid:18)Aq + Bq 2 (cid:19)1/q . (1.3) A more general result involving positive linear maps is known under suitable assump- tions on p, q in [6, 12, 13, 14] (see Theorems 2.1 and 2.2 below). Our interest here is showing that these sufficient conditions of p, q are best possible for (1.3) to hold. Although the result is naturally expected, no rigorous proof is known to the best of our knowledge. This question for the best possible conditions of p, q showed up in some concavity/convexity problem of a certain matrix function in [10]. It is useful to write power means in terms of a positive linear map of block-diagonal matrices. Defining a positive linear map Φ : M2n → Mn by Φ(cid:18)(cid:20)A X Y B(cid:21)(cid:19) := A + B 2 for matrices in M2n partitioned in blocks in Mn, one can write for A, B ∈ Pn Φ(cid:18)(cid:20)A 0 0 B(cid:21)p(cid:19)1/p =(cid:18)Ap + Bp 2 (cid:19)1/p . Therefore, it is also interesting to determine p, q for which the inequality Φ(Ap)1/p ≤ Φ(Aq)1/q (1.4) (1.5) holds for every unital positive linear map Φ. Most fundamental in such matrix/operator inequalities are Choi's inequality [4] (extending Davis [5]) and Hansen and Pedersen's Jensen inequality [7]. The paper is organized as follows. In Section 2 we state in more precise terms our problem on the best possible p, q for matrix inequalities (1.3) and (1.5) together with the known affirmative results. A motivation coming from [10] is also explained. Section 3 is the body of the proof of our main result by constructing counterexamples to (1.3), all of which are given by 2 × 2 matrices. Those are further reformulated to give counterexamples to (1.5) for Φ : M3 → M2. 2 2 Result and motivation The main aim of this paper is to determine the range of real parameters p, q for which the matrix inequality between the power means in (1.3) holds. Before stating the main result we first recall the affirmative result, which is known to hold in a more general setting of (1.5). Let H and K be general Hilbert spaces. Let B(H) be the algebra of all bounded linear operators on H and B(H)++ the set of all positive invertible operators on H. Let Φ : B(H) → B(K) be a positive linear map that is unital, i.e., Φ(IH) = IK, where IH denotes the identity operator on H. Then the map A ∈ B(H)++ 7−→ Φ(Ap)1/p ∈ B(K)++ (2.1) can be defined for every p ∈ R with p 6= 0. Indeed, for every A∈ B(H)++ and for every p 6= 0, since Ap ≥ δIH for some δ > 0, Φ(Ap) ≥ δIK so that Φ(Ap) ∈ B(K)++. Moreover, the following convergence in the operator norm is straightforward: lim p→0 for every A ∈ B(H)++. Indeed, Φ(Ap)1/p = exp Φ(log A) (2.2) 1 p log Φ(Ap) = 1 p log Φ(IH + p log A + o(p)) = 1 p log(IK + pΦ(log A) + o(p)) = Φ(log A) + o(p), where o(p) means that o(p)/p → 0 in the operator norm as p → 0. So we shall write Φ(Ap)1/p when p = 0 to mean exp Φ(log A). Under the above assumption, we state the following result which can be considered folklore. Theorem 2.1. Let p, q ∈ R. The operator inequality Φ(Ap)1/p ≤ Φ(Aq)1/q holds for every A ∈ B(H)++ if (p, q) satisfies one of the following conditions: p = q, 1 ≤ p < q, p < q ≤ −1, p ≤ −1, q ≥ 1, 1/2 ≤ p < 1 ≤ q, p ≤ −1 < q ≤ −1/2. (2.3)   Proof. For the convenience of the reader we give a concise proof using Choi's inequality [4, Theorem 2.1]. When 1 ≤ p < q, we have Φ(Ap) ≤ Φ(Aq)p/q so that Φ(Ap)1/p ≤ Ψ(Aq)1/q. When p ≤ −1 and q ≥ 1, or when 1/2 ≤ p < 1 ≤ q, we have Φ(Ap)1/p ≤ Φ(A) ≤ Φ(Aq)1/q. The proof is similar for the remaining cases. 3 Next, let H1 and H2 be Hilbert spaces and Φi : B(Hi) → B(K) be positive linear maps, i = 1, 2, such that Φ1(IH1) + Φ2(IH2) = IK. Define a unital positive linear map Φ : B(H1 ⊕ H2) → B(K) by Φ(cid:18)(cid:20)A X Y B(cid:21)(cid:19) := Φ1(A) + Φ2(B) for A ∈ B(H1) and B ∈ B(H2). For this Φ, restricting map (2.1) to A ⊕ B defines (A, B) ∈ B(H1)++ × B(H2)++ 7−→ (Φ1(Ap) + Φ2(Bp))1/p ∈ B(K)++. When p = 0, this means exp(Φ1(log A)+Φ2(log B)) by (2.2). Therefore, the next result is a special case of Theorem 2.1, which was shown in [12, 13, 14] (see also [6, Chapter 4]). In fact, results in more general forms were given there. Theorem 2.2. Let Φi, i = 1, 2, be as above. Then the operator inequality (Φ1(Ap) + Φ2(Bp))1/p ≤ (Φ1(Aq) + Φ2(Bq))1/q holds for every A ∈ B(H1)++ and B ∈ B(H2)++ if (p, q) satisfies one of the conditions in (2.3). Obviously, when Φ1(X) = Φ2(X) = (1/2)X for X ∈ B(H), the expressions in (2.1) and (2.2) reduce to the power mean in (1.1) and the Log-Euclidean mean in (1.2), respectively. Hence, the above theorem says that, in particular, the matrix inequality between the power means in (1.3) holds if (p, q) satisfies one of (2.3). It is natural to expect that the converse is also true, that is, (2.3) is the optimal range of (p, q) for which (1.3) holds true. For this converse direction, it seems that no rigorous proof is known so far. Now, the following is our main result, which completely settles the converse direction. Theorem 2.3. Let p, q ∈ R, and assume that matrix inequality (1.3) holds for every A, B ∈ P2. Then (p, q) satisfies one of the conditions in (2.3). To prove the theorem, we need to provide counterexamples to (1.3) for any (p, q) outside the range given in (2.3), which will be done in the next section. It turns out that all counterexamples are 2 × 2 matrices. Restricted to the case q = 1, the theorem says the well-known fact [8, Proposition 3.1] that the function tp on (0,∞) is 2-convex if and only if either 1 ≤ p ≤ 2 or −1 ≤ p ≤ 0, so 2-convexity implies operator convexity in this case. Theorem 2.3, with Theorem 2.1, shows that when Φ : M4 → M2 is (1.4) for n = 2, matrix inequality (1.5) holds for every A ∈ P4 if and only if (p, q) satisfies one of (2.3). However, we can reformulate counterexamples in Theorem 2.3 to obtain the following better result. The proof will be given in the last of the next section. 4 Theorem 2.4. Let p, q ∈ R, and assume that matrix inequality (1.5) holds for every unital completely positive linear map Φ : M3 → M2 and every A, B ∈ P3. Then (p, q) satisfies one of the conditions in (2.3). Related to the above theorem, the following remarks are worth mentioning: (1) In particular, when q = 1, the above theorem says that the Jensen inequality Φ(A)p ≤ Φ(Ap) holds for every unital (completely) positive linear map Φ : M3 → M2 and every A ∈ P3 if and only if either 1 ≤ p ≤ 2 or −1 ≤ p ≤ 0. (2) Choi [4] gave a convenient counterexample when Φ : M3 → M2 is the compres- sion map taking A ∈ M3 to the 2 × 2 top left corner of A. Choi's example is 1 0 1 0 0 1 1 1 1 A :=    , for which Φ(A)4 6≤ Φ(A4). Since this A is not positive definite, we take Then a numerical computation shows that the signs of the eigenvalues of Φ(Bp)−Φ(B)p are 2 0 1 0 1 1 1 1 2 B := A + I3 =   > 0. −, + if p < −1, +, + if −1 < p < 0, −,− if 0 < p < 1, +, + if 1 < p < 2, −, + if p > 2.   Thus, Φ(B)p ≤ Φ(Bp) holds only if either 1 ≤ p ≤ 2 or −1 ≤ p ≤ 0, and it holds reversed only if 0 ≤ p ≤ 1. (3) The matrix sizes 3 and 2 in Φ : M3 → M2 of Theorem 2.4 are minimal. Indeed, it is well-known that when ϕ is a positive linear functional on Mn, we have f (ϕ(A)) ≤ ϕ(f (A)) for every Hermitian A ∈ Mn and every convex function f defined on an interval containing the eigenvalues of A. Also, it is known [2, Theorem 2.2] that when Φ : M2 → Mn is a unital positive linear map, the inequality f (Φ(A)) ≤ Φ(f (A)) holds true for every Hermitian A ∈ M2 and every convex function f as above. Furthermore, we have the next result showing that the situation is also similar for inequality (1.5). Theorem 2.5. Let Φ : M2 → Mn be a unital positive linear map. Then (1.5) holds true for every A, B ∈ P2 and every p, q ∈ R with p ≤ q. Proof. The proof is similar to that of [2, Theorem 2.2]. Let p < q be arbitrary and let A ∈ P2. We may assume by continuity that A has eigenvalues λ1 > λ2 such that 5 λ1λp 2 6= λ2λp 1 and λ1λq 2 6= λ2λq 1. Then the computation in [2] gives Φ(Ap) = Φ(A) − 2 1 − λp λp λ1 − λ2 λq 1 − λq λ1 − λ2 2 λ2λp λ2λq 2 1 − λ1λp λ1 − λ2 1 − λ1λq λ1 − λ2 2 , . Φ(Aq) = Φ(A) − Since λ2In ≤ Φ(A) ≤ λ1In, the result follows since ≤(cid:18)λq λ1 − λ2 (cid:19)1/p 1 − λ1λp x − 2 1 − λp (cid:18)λp λ1 − λ2 λ2λp 2 2 1 − λq λ1 − λ2 that is, x − λ2λq λ1 − λ2 (cid:19)1/q 1 − λ1λq 2 , (cid:18) x − λ2 λ1 − λ2 for any x ∈ [λ2, λ1]. λp 1 + λ1 − x λ1 − λ2 λp 2(cid:19)1/p ≤(cid:18) x − λ2 λ1 − λ2 λq 1 + λ1 − x λ1 − λ2 λq 2(cid:19)1/q In the rest of the section we explain what motivated us to prove the optimality of conditions (2.3) for the validity of (1.3). In [10] we discussed joint concavity/convexity of the trace function (A, B) ∈ Pn × Pm 7−→ Tr(cid:8)Φ(Ap)1/2Ψ(Bq)Φ(Ap)1/2(cid:9)s, where p, q, s are real parameters, n, m, l ∈ N, and Φ : Mn → Ml and Ψ : Mm → Ml are (strictly) positive linear maps. We are interested in extending concavity/convexity results under trace to those under symmetric (anti-) norms. (The notion of symmetric anti-norms was introduced in [3].) For instance, we are interested in joint convexity of the norm function (A, B) ∈ Pn × Pm 7−→(cid:13)(cid:13)(cid:8)Φ(Ap)1/2Ψ(Bq)Φ(Ap)1/2(cid:9)s(cid:13)(cid:13), where k · k is a symmetric norm on Ml. This joint convexity for any symmetric norm can be reduced to that for the Ky Fan k-norms for k = 1, . . . , l. Although the problem for all Ky Fan norms seems difficult, we could settle in [10] the special case where k = 1, i.e., k · k is the operator norm k · k∞ (another special case where k = l is the original situation under trace). In [10] we proved Theorem 2.6. Under the above assumption, the function is jointly convex if one of the following six conditions is satisfied: (A, B) ∈ Pn × Pm 7−→(cid:13)(cid:13)(cid:8)Φ(Ap)1/2Ψ(Bq)Φ(Ap)1/2(cid:9)s(cid:13)(cid:13)∞  −1 ≤ p, q ≤ 0 and s > 0, −1 ≤ p ≤ 0, 1 ≤ q ≤ 2, p + q > 0 and s ≥ 1/(p + q),  1 ≤ p ≤ 2, −1 ≤ q ≤ 0, p + q > 0 and s ≥ 1/(p + q), and their counterparts where (p, q, s) is replaced with (−p,−q,−s). 6 (2.4) Moreover, for the optimality of the above conditions in (2.4) for (p, q, s) we proved Theorem 2.7. The function (A, B) ∈ Pn × Pn 7−→ k(Ap/2BqAp/2)sk∞ (2.5) is jointly convex for every n ∈ N (or equivalently, for fixed n = 2) if and only if (p, q, s) satisfies one of the conditions in (2.4) and their counterparts of (−p,−q,−s) in place of (p, q, s). The "if " part of Theorem 2.7 is an obvious special case of Theorem 2.6. To prove the "only if " part, we observed that, for each n ∈ N, p, q 6= 0 and s > 0, if (2.5) is jointly convex then (cid:18)A1/q + B1/q 2 (cid:19)q ≤(cid:18) A−1/p + B−1/p 2 (cid:19)−p holds for every A, B ∈ Pn. In this way, the matrix inequality between the power means shows up, and the restriction on (p, q) obtained in Theorem 2.3 is crucial to prove Theorem 2.7. So we need to prove Theorem 2.3 to complete the proof of Theorem 2.7 in [10], which is our main motivation here, though Theorem 2.3 is certainly of independent interest. 3 Counterexamples This section is mostly devoted to the proof of Theorem 2.3 by constructing counterex- amples. It is obvious that the condition p ≤ q is necessary for (1.3) to hold for the numerical function (i.e., for A = aI and B = bI with a, b ∈ (0,∞)). From the obvious identities 2 (cid:19)1/p =(cid:26)(cid:18) (A−1)−p + (B−1)−p (cid:18)Ap + Bp (cid:19) =(cid:26)exp(cid:18)log A−1 + log B−1 exp(cid:18) log A + log B (cid:19)−1/p(cid:27)−1 (cid:19)(cid:27)−1 2 2 2 it is also obvious that, for each n ∈ N, (1.3) holds for every A, B ∈ Pn if and only if (1.3) with (−q,−p) in place of (p, q) holds for every A, B ∈ Pn. Therefore, it suffices to provide counterexamples for any (p, q) such that either −1 < p < 1/2 and q > max{0, p}, or 1/2 ≤ p < q < 1. Below we divide our job into three cases which cover all of such (p, q). , p 6= 0, , 3.1 Case −1 < p < 1/2, p 6= 0 and q > max{0, p} For each x, y > 0 and θ ∈ R define A, Bθ ∈ P2 by Bθ :=(cid:20)cos θ − sin θ 0 x(cid:21) , A :=(cid:20)1 0 cos θ (cid:21)(cid:20)1 0 sin θ 0 y(cid:21)(cid:20) cos θ − sin θ cos θ(cid:21) . sin θ 7 Lemma 3.1. Let p, q ∈ R \ {0} and x, y > 0 be such that xp + yp 6= 2, xq + yq 6= 2 and ((xp + yp)/2)1/p 6= ((xq + yq)/2)1/q. Then we have det(cid:26)(cid:18)Aq + Bq = θ2"1 −(cid:18) Ap + Bp (cid:19)1/q 2 θ θ (cid:19)1/p(cid:27) 2 q(2 − xq − yq) (cid:27)(cid:26)(cid:18)xq + yq (1 − xq)(1 − yq) 2 − xq − yq(cid:27)2(cid:26)1 −(cid:18)xp + yp 1 − yq 2 (cid:19)1/q −(cid:18) xp + yp 2 (cid:19)1/p(cid:27)(cid:26)1 −(cid:18) xq + yq 2 (cid:19)1/p(cid:27) 2 (cid:19)1/q(cid:27)# 2(cid:26)(1 − xp)(1 − yp) p(2 − xp − yp) − −(cid:26) 1 − yp 2 − xp − yp − + o(θ2) as θ → 0. Proof. We have Ap + Bp θ =(cid:20)2 − (1 − yp) sin2 θ (1 − yp) sin 2θ 2 (1 − yp) sin 2θ xp + yp + (1 − yp) sin2 θ(cid:21) 2 = G + θH + θ2K + o(θ2), where G :=(cid:20)2 0 xp + yp(cid:21) , H :=(cid:20) 0 0 1 − yp 1 − yp 0 (cid:21) , K :=(cid:20)−(1 − yp) 0 0 1 − yp(cid:21) . We apply the Taylor formula with Fr´echet derivatives (see e.g., [9, Theorem 2.3.1]) to obtain (Ap + Bp θ )1/p = G1/p + D(x1/p)(G)(θH + θ2K) + 1 2 D2(x1/p)(G)(θH, θH) + o(θ2), where the second and the third terms in the right-hand side are the first and the second Fr´echet derivatives of X ∈ P2 7→ X 1/p ∈ P2 at G, respectively. By Daleckii and Krein's derivative formula (see [1, Theorem V.3.3], [9, Theorem 2.3.1]) we have D(x1/p)(G)(θH + θ2K) (x1/p)[1](2, 2) (x1/p)[1](2, xp + yp) (x1/p)[1](2, xp + yp) (x1/p)[1](xp + yp, xp + yp)(cid:21) ◦ (θH + θ2K) =(cid:20) =" = θ" 1 p−1 1 p 2 21/p−(xp+yp)1/p 2−xp−yp 0 21/p−(xp+yp)1/p 2−xp−yp 1 p (xp + yp) 1 21/p−(xp+yp)1/p 2−xp−yp p2 + θ2"− 1 1 (1 − yp) p−1(1 − yp) 0 2−xp−yp 21/p−(xp+yp)1/p p−1# ◦ (θH + θ2K) # (1 − yp) p−1(1 − yp)# , 0 p(xp + yp) 0 1 1 8 where (x1/p)[1] denotes the first divided difference of x1/p and ◦ means the Schur (or Hadamard) product. For the second divided difference of x1/p we compute (x1/p)[2](2, 2, xp + yp) = (cid:0) 1 1 p−1(xp + yp) + (xp + yp)1/p , p − 1(cid:1)21/p − 1 21/p − 2 p (xp + yp) p2 (2 − xp − yp)2 p−1 +(cid:0) 1 (2 − xp − yp)2 1 p − 1(cid:1)(xp + yp)1/p , (x1/p)[2](2, xp + yp, xp + yp) = and hence we have 1 2 D2(x1/2)(G)(θH, θH) = θ2  (xp+yp)+(xp+yp)1/p 1 p −1 p−1)21/p− 1 ( 1 p 2 (2−xp−yp)2 0 (1 − yp)2 (1 − yp)2 (In the above computation we have used the assumption that xp + yp 6= 2.) Therefore, it follows that 21/p− 2 p (xp+yp) 1 p −1 +( 1 (2−xp−yp)2 0 p−1)(xp+yp)1/p .   where α(1,1) p := − 1 2p (1 − yp) + (cid:18) Ap + Bp 2 θ (cid:19)1/p ="1 + α(1,1) p α(1,2) p θ θ2 α(1,2) p θ + α(2,2) p θ2# + o(θ2), (3.1) 2 (cid:0) xp+yp (cid:1)1/p = − 2p(2 − xp − yp)2 (2 − 2p) − (xp + yp) + 2p2−1/p(xp + yp)1/p (1 − yp)2 2−1/p(1 − yp)2(cid:8)21/p − (xp + yp)1/p(cid:9) 2p(2 − xp − yp) − 1 (1 − yp) + 2p (1 − xp)(1 − yp) = − 2p(2 − xp − yp) − (1 − yp)(cid:8)1 − ( xp+yp (2 − xp − yp) is not written down here since it is unnecessary in the computation (1 − yp)2 (1 − yp)2(cid:8)1 − ( xp+yp (2 − xp − yp)2 )1/p(cid:9) (2 − xp − yp)2 )1/p(cid:9) . , 2 2 α(1,2) p := p (The form of α(2,2) below.) By assumption ((xp + yp)/2)1/p 6= ((xq + yq)/2)1/q, we arrive at det(cid:26)(cid:18)Aq + Bq = θ2"(cid:8)α(1,1) (cid:19)1/q −(cid:18)Ap + Bp p (cid:9)(cid:26)(cid:18) xq + yq q − α(1,1) 2 (cid:19)1/p(cid:27) −(cid:8)α(1,2) (cid:19)1/p(cid:27) 2 (cid:19)1/q −(cid:18)xp + yp q − α(1,2) 2 2 θ θ p (cid:9)2# + o(θ2). The above formula inside the big bracket is equal to the sum of the following ∆1 and ∆2: ∆1 := 1 2(cid:26) (1 − xp)(1 − yp) p(2 − xp − yp) − q(2 − xq − yq) (cid:27)(cid:26)(cid:18) xq + yq (1 − xq)(1 − yq) 2 (cid:19)1/q −(cid:18)xp + yp 2 (cid:19)1/p(cid:27), 9 2 2 + ∆2 :=(− (1 − yq)2(cid:0)1 − ( xq+yq (2 − xq − yq)2 2 (cid:19)1/q ×(cid:26)(cid:18)xq + yq −((1 − yq)(cid:0)1 − ( xq+yq (2 − xq − yq) (1 − yp)2(cid:0)1 − ( xp+yp )1/q(cid:1) (2 − xp − yp)2 2 (cid:19)1/p(cid:27) −(cid:18)xp + yp (1 − yp)(cid:0)1 − ( xp+yp )1/q(cid:1) − (2 − xp − yp) Letting wp := 1 − ((xp + yp)/2)1/p we furthermore compute 2 2 )1/p(cid:1) ) )1/p(cid:1) )2 . ∆2 =(− (1 − yq)2wq (2 − xq − yq)2 + −( (1 − yq)wq (2 − xq − yq) − 2 − xp − yp − (2 − xp − yp)2)(−wq + wp) (1 − yp)2wp (2 − xp − yp))2 (1 − yp)wp 2 − xq − yq(cid:27)2 1 − yq wpwq, =(cid:26) 1 − yp and the lemma follows from the above expressions of ∆1 and ∆2. Now, let −1 < p < 1/2, p 6= 0 and q > max{0, p}. We prove that (cid:18)Ap + Bp 2 θ (cid:19)1/p 6≤(cid:18)Aq + Bq 2 θ (cid:19)1/q for some x, y > 0 and some θ > 0. Suppose on the contrary that (cid:18)Ap + Bp 2 θ (cid:19)1/p ≤(cid:18)Aq + Bq 2 θ (cid:19)1/q for all x, y > 0 and all θ > 0. Let 0 < x < 1 and y = x2. Then it is clear that (cid:18)xp + x2p 2 (cid:19)1/p xp + x2p 6= 2, 2 (cid:19)1/p = x(cid:18) 1 + xp xq + x2q < 2, < x(cid:18) 1 + xq 2 (cid:19)1/q =(cid:18)xq + x2q 2 (cid:19)1/q . Hence, by Lemma 3.1 we must have 1 2(cid:26) (1 − xp)(1 − x2p) p(2 − xp − x2p) − −(cid:26) 1 − x2p 2 − xp − x2p − ≥ 0. q(2 − xq − x2q) (cid:27)(cid:26)(cid:18) xq + x2q (1 − xq)(1 − x2q) 2 − xq − x2q(cid:27)2(cid:26)1 −(cid:18)xp + x2p 1 − x2q 2 2 2 (cid:19)1/q (cid:19)1/p(cid:27) −(cid:18)xp + x2p (cid:19)1/q(cid:27) (cid:19)1/p(cid:27)(cid:26)1 −(cid:18) xq + x2q 2 (3.2) 10 x(1 + xp)1/p 21/p − ≈ 21/p − 21/q 1 p + 1 q 2 x (3.3) 2 When 0 < p < 1/2 and q > p, we have as x ց 0 x(1 + xq)1/q (cid:18)xq + x2q (cid:19)1/q 1 − x2p 2 − xp − x2p − (cid:19)1/p −(cid:18)xp + x2p 1 − x2q 2 − xq − x2q = and 21/q = 2 xp − xq − x2p + x2q − xp+2q + x2p+q (2 − xp − x2p)(2 − xq − x2q) xp 4 . ≈ Therefore, the dominant term of the left-hand side of (3.2) is 21/p − 21/q p + 1 21+ 1 q (cid:18) 1 2p − 1 2q(cid:19)x − x2p 16 < 0 thanks to 2p < 1 when x > 0 is sufficiently small. This contradicts (3.2). When −1 < p < 0 and q > 0, we have the same estimation (3.3), and moreover and (1 − xq)(1 − x2q) q(2 − xq − x2q) ≈ − xp p (1 − xp)(1 − x2p) p(2 − xp − x2p) − 1 − x2q 1 − x2p 2 − xq − x2q ≈ 1 − 2 − xp − x2p − p (cid:19) − 21/p − 21/q p + 1 21+ 1 q (cid:18)− xp+1 1 2 = 1 2 as x ց 0. 1 4 < 0 Therefore, the left-hand side of (3.2) is dominantly thanks to p + 1 > 0 for x > 0 sufficiently small, and we have a contradiction again. 3.2 Case p = 0 < q For x, y > 0 let A, Bθ ∈ P2 be the same as in Section 3.1. The following is the counterpart of Lemma 3.1 in the case p = 0. The expression here can easily be obtained by taking the limit of that in Lemma 3.1 as p → 0. However, deriving the expression in this way is not a rigorous proof, so we sketch an independent proof. Lemma 3.2. Let q ∈ R \ {0} and x, y > 0 be such that xy 6= 1, xq + yq 6= 2 and x 6= y (hence √xy 6= ((xq + yq)/2)1/q). Then we have − exp(cid:18)log A + log Bθ 2 θ 2 det(cid:26)(cid:18)Aq + Bq = θ2"− 1 (cid:19)1/q 2(cid:26) log x · log y −(cid:26) log y log xy − + o(θ2) as θ → 0. log xy + (cid:19)(cid:27) q(2 − xq − yq) (cid:27)(cid:26)(cid:18) xq + yq (1 − xq)(1 − yq) 2 − xq − yq(cid:27)2 1 − yq 2 (cid:19)1/q (cid:0)1 − √xy(cid:1)(cid:26)1 −(cid:18)xq + yq − √xy(cid:27) 2 (cid:19)1/q(cid:27)# 11 Proof. We have where log A + log Bθ =(cid:20) log y · sin2 θ − log y · sin 2θ 2 − log y · sin 2θ log xy − log y · sin2 θ(cid:21) 2 = G + θH + θ2K + o(θ2), 0 G :=(cid:20)0 0 log xy(cid:21) , H :=(cid:20) 0 − log y − log y 0 (cid:21) , K :=(cid:20)log y 0 0 − log y(cid:21) . As in the proof of Lemma 3.1, 2 H 2 (cid:19) + θ2 K exp(cid:18)log A + log Bθ = eG/2 + D(ex)(G/2)(cid:18)θ 2 (cid:19) = θ" 2 (cid:19) = θ2"− log2 y D(ex)(G/2)(cid:18)θ D2(ex)(G/2)(cid:18)θ + θ2 K H 2 H 2 , θ H 0 1 2 H 1 2 D2(ex)(G/2)(cid:18)θ (1−√xy) log y 2 (cid:19) + (1−√xy) log y 2 log xy − (1−√xy) log2 y H 2 , θ 2 (cid:19) + o(θ2), (cid:21) , # , # + θ2(cid:20) log y √xy log2 y 2 log xy + (1−√xy) log2 y 0 − √xy log y log2 xy log xy log xy log2 xy 0 0 0 0 2 2 where we have used assumption xy 6= 1. Therefore, we write α(1,2) θ 0 √xy + α(2,2) exp(cid:18) log A + log Bθ θ2 2 0 θ2# + o(θ2), (3.4) θ 0 α(1,2) 0 (cid:19) ="1 + α(1,1) (1 − √xy) log2 y log2 xy where α(1,1) 0 := α(1,2) 0 := log2 y log y 2 − 2 log xy − (1 − √xy) log y . log xy = log x · log y 2 log xy − (1 − √xy) log2 y log2 xy , θ Since √xy 6= ((xq + yq)/2)1/q by assumption, we obtain, by (3.4) and (3.1) with q, (cid:19)1/q det(cid:26)(cid:18)Aq + Bq = θ2"(cid:8)α(1,1) q − α(1,1) 0 (cid:9)2(cid:21) + o(θ2). Letting w0 := 1− √xy as well as wq := 1− ((xq + yq)/2)1/q we compute the expression − exp(cid:18)log A + log Bθ 2 (cid:19)1/q 0 (cid:9)(cid:26)(cid:18) xq + yq (cid:19)(cid:27) − √xy(cid:27) −(cid:8)α(1,2) q − α(1,2) 2 2 in the above big bracket as (cid:26)− (1 − xq)(1 − yq) 2q(2 − xq − yq) − 2 log xy (cid:27)(cid:26)(cid:18)xq + yq log x · log y 2 (cid:19)1/q − √xy(cid:27) 12 1 w0 log2 y (1 − yq)wq +(cid:26)− 2 − xq − yq + 2(cid:26)(1 − xq)(1 − yq) = − 2q(2 − xq − yq) 2 − xq − yq(cid:27)2 1 − yq −(cid:26) log y log xy − w0wq, log2 xy (cid:27)(−wq + w0) −(cid:26) (1 − yq)2wq (2 − xq − yq)2 − 2 (cid:19)1/q 2 log xy (cid:27)(cid:26)(cid:18)xq + yq − √xy(cid:27) log x · log y + w0 log y log xy (cid:27)2 and the assertion follows. Now, let q > 0. We suppose that exp(cid:18)log A + log Bθ 2 (cid:19) ≤(cid:18)Aq + Bq 2 θ (cid:19)1/q for all x, y > 0 and all θ > 0. Let 0 < x < 1 and y = x2, so xq + yq 6= 2 and x 6= y. Hence, by Lemma 3.2 we must have log x + 1 3 − 2(cid:26)2 −(cid:26) 2 3 − As x ց 0 we have 2 (cid:19)1/q q(2 − xq − x2q) (cid:27)(cid:26)(cid:18) xq + x2q (1 − xq)(1 − x2q) 2 − xq − x2q(cid:27)2 1 − x2q (1 − x3/2)(cid:26)1 −(cid:18)xq + x2q (cid:19)1/q (cid:18)xq + x2q − x3/2 ≈ 1 21/q x 2 2 − x3/2(cid:27) (cid:19)1/q(cid:27) ≥ 0. so that the left-hand side of (3.5) is dominantly (3.5) − 1 3 · 21/q x log x −(cid:18)2 3 − 1 2(cid:19)2 = − 1 3 · 21/q x log x − 1 36 < 0, a contradiction. Hence it has been shown that, for every q > 0, exp(cid:18)log A + log Bθ 2 for some x, y > 0 and some θ > 0. 3.3 Case 0 < p < q < 1 (cid:19) 6≤(cid:18)Aq + Bq 2 θ (cid:19)1/q For θ ∈ R define 2 × 2 positive semidefinite matrices Bθ :=(cid:20) cos2 θ 0 0(cid:21) , A :=(cid:20)2 0 cos θ sin θ cos θ sin θ sin2 θ (cid:21) . Indeed, the latter is Bθ in Section 3.1 with y = 0 while the former is slightly different from A in Section 3.1 with x = 0. 13 Lemma 3.3. For every p, q ∈ (0, 1), (cid:19)1/q −(cid:18)Ap + Bp det(cid:26)(cid:18)Aq + Bq 2 (cid:19)1/p(cid:18)2q + 1 = −θ2(cid:18)2p + 1 (cid:19)1/p(cid:27) 2 (cid:19)1/q(cid:18) 1 2p + 1 − 2 2 θ θ 1 2q + 1(cid:19)2 + o(θ2) as θ → 0. Proof. Since (Ap + Bp θ )/2 is singular at θ = 0 and x1/p is singular at x = 0, the Taylor formula applied in Sections 3.1 and 3.2 cannot be used. However, a direct approximate computation is not difficult as below. Since B is a rank one projection, we write Ap + Bp θ 2 where =(cid:20) 2p+1−sin2 θ sin 2θ 2 4 sin 2θ sin2 θ 4 2 (cid:21) = 2p + 1 4 (cid:20)1 + a b b 1 − a(cid:21) , a := 1 − 2 sin2 θ 2p + 1 , b := sin 2θ 2p + 1 . b 1 − a(cid:21) has the eigenvalues 1 + c and 1 − c with c := √a2 + b2 −b (cid:21), respectively, from which one can b (cid:21) and (cid:20)c − a θ b compute Observe that (cid:20)1 + a (< 1) and the eigenvectors are (cid:20)c + a (cid:19)1/p (cid:18)Ap + Bp 4 (cid:19)1/p(cid:20)c + a c − a =(cid:18)2p + 1 4 (cid:19)1/p" (1+c)1/p+(1−c)1/p =(cid:18)2p + 1 As θ ց 0 we compute 2 2c b 2 −b (cid:21)(cid:20)(1 + c)1/p 0 0 −b (cid:21)−1 (1 − c)1/p(cid:21)(cid:20)c + a c − a b + (1+c)1/p−(1−c)1/p a (1+c)1/p−(1−c)1/p 2c b (1+c)1/p−(1−c)1/p b a# . (1+c)1/p+(1−c)1/p 2 2c − (1+c)1/p−(1−c)1/p 2c a = 1 − 2θ2 2p + 1 + o(θ2), b = 2θ 2p + 1 + o(θ), c2 = a2 + b2 = 1 − 2p+2θ2 (2p + 1)2 + o(θ2) so that and c = 1 − 2p+1θ2 (2p + 1)2 + o(θ2), 1 c = 1 + 2p+1θ2 (2p + 1)2 + o(θ2) (1 + c)1/p = 21/p(cid:18)1 − 2pθ2 p(2p + 1)2(cid:19) + o(θ2), (1 − c)1/p = o(θ2) 14 thanks to p ∈ (0, 1). Therefore, the (1, 1) entry of ((Ap + Bp θ )/2)1/p is α(1,1) p =(cid:18)2p + 1 1 1 + 2 2pθ2 2pθ2 p−1(cid:18)1 − 4 (cid:19)1/p(cid:26)2 p−1(cid:18)1 − p +1 (cid:26)2 − (cid:18)1 − p(2p + 1)2(cid:19) p(2p + 1)2(cid:19)(cid:18)1 + 2p+1θ2 2p+1θ2 p(2p + 1)2 + (2p + 1)2 − 2p + p p(2p + 1)2 θ2(cid:19) + o(θ2). 1 21/p 2 (2p + 1)1/p (2p + 1)1/p = = 2p+1θ2 (2p + 1)2(cid:19)(cid:18)1 − 2θ2 2p + 1(cid:19)(cid:27) + o(θ2) 2θ2 2p + 1(cid:27) + o(θ2) The (2, 2)-entry of ((Ap + Bp θ )/2)1/p is α(2,2) 4 (cid:19)1/p(cid:26)2 p =(cid:18)2p + 1 p−1(cid:18)1 + − 2 (2p + 1)1/p p +1 (cid:26)− = 2 1 1 1 2pθ2 p−1(cid:18)1 − p(2p + 1)2(cid:19) (2p + 1)2(cid:19)(cid:18)1 − 2p+1θ2 2p+1θ2 (2p + 1)2 + 2θ2 2p + 1(cid:27) + o(θ2) 1 p−2 = (2p + 1) 21/p θ2 + o(θ2). 2pθ2 p(2p + 1)2(cid:19)(cid:18)1 − 2θ2 2p + 1(cid:19)(cid:27) + o(θ2) The (1, 2)-entry of ((Ap + Bp θ )/2)1/p is α(1,2) p =(cid:18)2p + 1 4 (cid:19)1/p 2 1 p−1(cid:18)1 + 2p+1θ2 (2p + 1)2(cid:19)(cid:18)1 − 2pθ2 p(2p + 1)2(cid:19) 2θ 2p + 1 + o(θ2) 1 p−1 = (2p + 1) 21/p θ + o(θ2). By the above estimate for ((Ap + Bp θ )/2)1/p and the same for ((Aq + Bq θ )/2)1/q we obtain θ θ 2 (cid:19)1/q det(cid:26)(cid:18) Aq + Bq −(cid:18)Ap + Bp 2 =(cid:8)α(1,1) q − α(1,1) p (cid:9)(cid:8)α(2,2) =(cid:26) (2q + 1)1/q − q −1 −(cid:26) (2q + 1) (cid:19)1/p(cid:27) q − α(2,2) p (cid:9) −(cid:8)α(1,2) (cid:27)(cid:26)(2q + 1) (cid:27)2 (2p + 1) 21/p (2p + 1)1/p 21/p 21/q 21/q 21/q p−1 − 1 1 1 p (cid:9)2 q − α(1,2) q −2 1 (2p + 1) 21/p − 1 p−2 (cid:27)θ2 θ2 + o(θ2) (2p + 1) p−2(2q + 1)1/q (2p + 1)1/p(2q + 1) 1 q −1 =(cid:26)− 1 p + 1 q 2 − 1 p + 1 q 2 15 + 2(2p + 1) 1 p−1(2q + 1) 1 q −1 1 p + 1 q 2 = −(cid:18) 2p + 1 2 (cid:19)1/p(cid:18)2q + 1 2 (cid:19)1/q(cid:18) 1 2p + 1 − (cid:27)θ2 + o(θ2) 2q + 1(cid:19)2 1 θ2 + o(θ2) as θ → 0. Now, let 0 < p < q < 1. Suppose that ((X p + Y p)/2)1/p ≤ ((X q + Y q)/2)1/q for all X, Y ∈ P2. By continuity this holds for all 2 × 2 positive semidefinite X, Y too so that (cid:18)Ap + Bp 2 θ (cid:19)1/p ≤(cid:18)Aq + Bq 2 θ (cid:19)1/q holds for any θ > 0. Then, by Lemma 3.3 we must have −(cid:18) 1 2p + 1 − 1 2q + 1(cid:19)2 ≥ 0, which implies that p = q, a contradiction. 3.4 Proof of Theorem 2.4 To prove Theorem 2.4, we may, in the same way as above for Theorem 2.3, provide counterexamples for the three cases of Sections 3.1 -- 3.3. This can easily be done by using the same examples as above. Case 3.1. Define a unital CP map (i.e., completely positive linear map) Φ : M3 → M2 by Φ(Z) := Z[1, 2] + UθZ[1, 3]U∗θ 2 for Z ∈ M3, where Z[i, j] denotes the principal submatrix of Z on rows and columns i and j, and For a diagonal matrix Z := diag(1, x, y), since cos θ (cid:21) . Uθ :=(cid:20)cos θ − sin θ sin θ Φ(Z p)1/p =(cid:18)Ap + Bp 2 θ (cid:19)1/p with A and Bθ in Section 3.1, we have a counterexample for this case in the same way as in Section 3.1. Case 3.2. By the same Φ and Z as in Case 3.1 we have a counterexample as in Section 3.2 since Φ(Z p)1/p for p = 0 is exp((log A + log Bθ)/2). 16 Case 3.3. Define a unital CP map Φ : M3 → M2 by Φ(Z) := Z[1, 3] + UθZ[2, 3]U∗θ 2 , where Uθ is as in Case 3.1. For a diagonal matrix Z := diag(2, 1, 0), since Φ(Z p)1/p = ((Ap + Bp θ )/2)1/p with A and Bθ in Section 3.3, we have a counterexample for this case. Acknowledgments The authors would like to thank the anonymous referee whose suggestions are quite helpful to improve the paper. In particular, Theorem 2.4 is based on the referee's suggestion. References [1] R. Bhatia, Matrix Analysis, Springer, New York, 1996. [2] R. Bhatia and R. Sharma, Some inequalities for positive linear maps, Linear Al- gebra Appl. 436 (2012), 1562 -- 1571. [3] J.-C. Bourin and F. Hiai, Norm and anti-norm inequalities for positive semi- definite matrices, Internat. J. Math. 22 (2011), 1121 -- 1138. [4] M.-D. Choi, A Schwarz inequality for positive linear maps on C∗-algebras, Illinois J. Math. 18 (1974), 565 -- 574. [5] C. Davis, A Schwarz inequality for convex operator functions, Proc. Amer. Math. Soc. 8 (1957), 42 -- 44. [6] T. Furuta, J. Mi´ci´c Hot, J. Pecari´c and Y. Seo, Mond-Pecari´c Method in Operator Inequalities, Element, Zagreb, 2005. [7] F. Hansen and G. K. Pedersen, Jensen's inequality for operators and Lowner's theorem, Math. Ann. 258 (1982), 229 -- 241. [8] F. Hansen and J. Tomiyama, Differential analysis of matrix convex functions II, J. Inequal. Pure Appl. Math. 10 (2009), Article 32, 5 pp. [9] F. Hiai, Matrix Analysis: Matrix Monotone Functions, Matrix Means, and Ma- jorization (GSIS selected lectures), Interdisciplinary Information Sciences 16 (2010), 139 -- 248. [10] F. Hiai, Concavity of certain matrix trace and norm functions, preprint (arXiv:1210.7524). 17 [11] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann. 246 (1980), 205 -- 224. [12] J. Mi´ci´c, Z. Pavi´c and J. Pecari´c, Jensen's inequality for operators without oper- ator convexity, Linear Algebra Appl. 434 (2011) 1228 -- 1237. [13] J. Mi´ci´c and J. Pecari´c, Order among quasi-arithmetic means of positive operators, II, Sci. Math. Jpn. 71 (2010), 93 -- 109. [14] J. Mi´ci´c, J. Pecari´c and Y. Seo, Order among quasi-arithmetic means of positive operators, Math. Rep. (Bucur.) 14(64) (2012), 71 -- 86. 18
1801.02824
1
1801
2018-01-09T07:20:25
A density result for homogeneous Sobolev spaces on planar domains
[ "math.FA", "math.CA" ]
We show that in a bounded simply connected planar domain $\Omega$ the smooth Sobolev functions $W^{k,\infty}(\Omega)\cap C^\infty(\Omega)$ are dense in the homogeneous Sobolev spaces $L^{k,p}(\Omega)$.
math.FA
math
A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ Abstract. We show that in a bounded simply connected planar domain Ω the smooth Sobolev functions W k,∞(Ω)∩C∞(Ω) are dense in the homogeneous Sobolev spaces Lk,p(Ω). 1. Introduction By the result of Meyers-Serrin [16] it is known that C ∞(Ω) is dense in W k,p(Ω) for every open set Ω in Rd. The space C ∞(Rd) is not always dense in W k,p(Ω), for example when Ω is a slit disk. However, a slit disk is not a very appealing example as it is not the interior of its closure. Counterexamples for the density satisfying Ω = int(Ω) were given by Amick [1] and Kolsrud [9]. In fact, in these examples even C(Ω) is not dense in W k,p(Ω). Going further in counterexamples, O'Farrell [18] constructed a domain satisfying Ω = int(Ω) where W k,∞(Ω) is not dense in W k,p(Ω) for any k and p. The domain constructed by O'Farrell was infinitely connected. From the recent results of Koskela-Zhang [14] and Koskela-Rajala-Zhang [13] we can conclude that this is necessary for such constructions in the plane, since W 1,∞(Ω) is dense in W 1,p(Ω) for all finitely connected bounded planar domains (see also the earlier work by Giacomini-Trebeschi [4]). Further examples of domains where W 1,p(Ω) is not dense in W 1,q(Ω) were constructed by Koskela [11] and Koskela-Rajala-Zhang [13]. In this note we continue the study of density of W k,∞(Ω) in W k,p(Ω). Let us remark that such density clearly holds in the case where the Sobolev functions in W k,p(Ω) can be extended to Sobolev functions defined on the whole R2. By work of Jones [8], this is true when ∂Ω is a quasi-circle. (See also the works [6, 5, 7].) Geometric characterizations of Sobolev extension domains are known, especially in the planar simply connected domains when k = 1, see [3, 10, 19, 12]. Being an extension domain is only a sufficient condition for the density. For example, there are Jordan domains Ω and functions f ∈ W 1,p(Ω) that cannot be extended to a function in W 1,p(R2). However, global smooth functions are dense in W 1,p(Ω) for any Jordan domain and any p ∈ [1,∞], see Lewis [15] and Koskela-Zhang [14]. For W k,p(Ω) with k ≥ 2 this is still unknown. In [20] Smith-Stanoyevitch-Stegenga studied the density of C ∞(R2) as well as the density of functions in C ∞(Ω) with bounded derivatives, in W k,p(Ω). For the latter class they Date: October 1, 2018. 2000 Mathematics Subject Classification. Primary 46E35. Key words and phrases. Sobolev space, homogeneous Sobolev space, density. All authors partially supported by the Academy of Finland. 1 2 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ obtained a density result assuming Ω to be starshaped or to satisfy an interior segment condition. For the smaller class of functions C ∞(R2) they also required an extra assumption on the boundary points to be m2-limit points. (See also Bishop [2] for a counterexample on a related question.) The result of Koskela-Zhang [14] showing that W 1,∞(Ω) is dense in W 1,p(Ω) for every bounded simply connected planar domain was generalized to higher dimensions by Koskela- Rajala-Zhang [13]. They showed that simply connectedness is not sufficient to give such a density result, but Gromov hyperbolicity in the hyperbolic distance is. In this paper we provide another generalization to the Koskela-Zhang result by going to higher order Sobolev spaces. We show that if we restrict attention to the homogenous norm, then being simply connected is sufficient for domains in the plane. For a domain Ω ⊂ R2 and p ∈ [1,∞), by homogenous Sobolev space Lk,p(Ω) we mean functions with p-integrable distributional derivatives of order k; Lk,p(Ω) = {u ∈ L1 loc(Ω) : ∇αu ∈ Lp(Ω), if α = k}, with semi-norm Pα=k k∇αukLp(Ω), where α is any 2-vector of non-negative integers and α is its ℓ1-norm. The (non-homogenous) Sobolev space W k,p(Ω) is defined as with norm Pα≤k k∇αukLp(Ω). Theorem 1.1. Let k ∈ N, p ∈ [1,∞) and Ω ⊂ R2 be a bounded simply connected domain. Then the subspace W k,∞(Ω) ∩ C ∞(Ω) is dense in the space Lk,p(Ω). loc(Ω) : ∇αu ∈ Lp(Ω), if α ≤ k}, W k,p(Ω) = {u ∈ L1 The approach in [13] differs from ours in that there the approximating functions are de- fined via shifting matters to the disk via the Riemann mapping. Instead, we directly make a Whitney decomposition of the domain and a rough reflection to define our approximating sequence. We achieve this via an elementary use of simply connectedness in the plane. In both of these approaches the values of the function in a suitable compact set are used to define a smooth function in the entire domain which approximates the original function in Sobolev norm. For this we employ similar tools as used by Jones in [8]. In p-Poincar´e domains, that is domains Ω where a p-Poincar´e inequality ZΩ u − uDp dx ≤ CZΩ ∇up dx holds, we can bound the integrals of the lower order derivatives by the integrals of the higher order ones and thus we obtain the following corollary to our Theorem 1.1. Corollary 1.2. Let k ∈ N, p ∈ [1,∞) and Ω ⊂ R2 be a bounded simply connected p- Poincar´e domain. Then W k,∞(Ω) ∩ C ∞(Ω) is dense in the space W k,p(Ω). For instance Holder-domains are p-Poincar´e domains for p ≥ 2, see Smith-Stegenga [21]. It still remains an open question whether Corollary 1.2 holds if one drops the assumption of being a p-Poincar´e domain. Next we come to the question of density of C ∞(R2) functions in Lk,p(Ω) in our setting of bounded simply connected domains. We have the following corollary which is analogous to A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 3 [14, Corollary 1.2], where it is shown that Ω being Jordan is sufficient. A small modification of the argument there applies to our situation as well. See the end of Section 4 for the proof. Corollary 1.3. Let k ∈ N, p ∈ [1,∞) and Ω ⊂ R2 be a Jordan domain. Then C ∞(R2) is dense in the space Lk,p(Ω). In Section 2, we collect the necessary ingredients which will be used for defining the approximating sequence; these include a suitable Whitney-type decomposition of a simply connected domain and a local polynomial approximation of Sobolev functions. In Section 3 we describe a partition of the domain using the Whitney-type decomposition of Section 2, which is needed for obtaining a suitable partition of unity. Then in Section 4, we define the approximating sequence and present the necessary estimates for proving Theorem 1.1 and Corollary 1.3. 2. Preliminaries For sets A, B ⊂ R2 we denote the diameter of A by diam(A) and the distance between A and B by dist(A, B). We denote by B(x, r) the open ball with center x ∈ R2 and radius r > 0 and more generally, by B(A, r) the open r-neighbourhood of a set A ⊂ R2. Given a connected set E ⊂ R2 and points x, y ∈ E, we define the inner distance dE(x, y) between x and y in E to be the infimum of lengths of curves in E joining x to y. (Notice that in general the infimum might have value ∞.) We write the inner distance in E between sets A, B ⊂ E as distE(A, B). With a slight abuse of notation, by a curve γ we refer to both, a continuous mapping γ : [0, 1] → R2 and its image γ([0, 1]). Given two curves γ1, γ2 : [0, 1] → R2 such that γ1(1) = γ2(0), we denote by γ1 ∗ γ2 : [0, 1] → R2 the concatenated curve γ1 ∗ γ2(t) = γ1(2t) for t ≤ 1/2 and γ1 ∗ γ2(t) = 2t − 1 for t ≥ 1/2. We denote the length of a curve γ by L(γ). We will use the following facts in plane topology whose proofs can be found in the book of Newman [17, Chapter VI, Theorem 5.1 and Chapter V, Theorem 11.8]. Lemma 2.1. Let Ω be a simply connected domain in R2 and γ : [0, 1] → R2 a continuous curve that is injective on (0, 1), whose endpoints γ(0) and γ(1) are in ∂Ω and interior γ((0, 1)) in Ω. Then Ω \ γ has two connected components, both of which are simply con- nected. In the case where Ω is Jordan and γ is homeomorphic to a closed interval, the two connected components of Ω \ γ have boundaries γ ∪ J1 and γ ∪ J2, where J1 and J2 are the two connected components of ∂Ω \ γ. 2.1. A dyadic decomposition. Although it is standard to consider a Whitney decom- position of a domain in Rd (see for instance Whitney [23] or the book of Stein [22, Chapter VI]), we will use a precise construction of such a decomposition. We present this construc- tion below. Here and later on we denote the sidelength of a square Q by l(Q). For notational convenience we start the Whitney decomposition below from squares with sidelength 2−1. Formally, by rescaling, we may consider all bounded domains Ω ⊂ R2 to 4 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ have diam(Ω) ≤ 1 in which case no Whitney decomposition would have squares larger than the ones used below regardless of the starting scale. Definition 2.2 (Whitney decomposition). Let Ω ⊂ R2 be a bounded (simply connected) open set. Let Qn be the collection of all closed dyadic squares of sidelength 2−n. Define a Whitney decomposition as F := Sn∈N Fn where the sets Fn are defined recursively as follows. Define  Q ∈ Q1 : [Q′∈Q1  F1 := Q′∩Q6=∅ and  Q′ ⊂ Ω   Q ∈ Qn+1 : Q 6⊂ Fn and [Q′∈Qn+1  Q′∩Q6=∅  Q′ ⊂ Ω  , Q. Lemma 2.3. A Whitney decomposition given by Definition 2.2 has the following properties. Fn+1 := where Fn =Sj≤nSQ∈ Fj (W1) Ω =SQ∈ F Q (W2) l(Q) < dist(Q, Ωc) ≤ 3√2l(Q) = 3diam(Q) for all Q ∈ F (W3) int Q1 ∩ int Q2 = ∅ for all Q1, Q2 ∈ F, Q1 6= Q2 (W4) If Q1, Q2 ∈ F and Q1 ∩ Q2 6= ∅, then l(Q1) l(Q2) ≤ 2. Proof. Although the proof is very elementary, we give it here for completeness. For (W1), take any x ∈ Ω and n ∈ N such that x ∈ Q ∈ Qn, where 2−n+2√2 < dist(x, Ωc) ≤ 2−n+3√2. Then for any Q′ ∈ Qn with Q′ ∩ Q 6= ∅ we have Q′ ⊂ Ω. Hence by definition either Q ∈ Fn or x ∈ Q ⊂ Q′′ ∈ Fi for some i < n. In order to see (W2), let Q ∈ Fn. Then all Q′ ⊂ Ω for all Q′ ∈ Qn with Q′ ∩ Q 6= ∅. Consequently, dist(Q, Ωc) > 2−n = l(Q). For the upper bound, suppose distQ, Ωc > 3√22−n. Let Q2 ∈ Qn−1 be such that Q ⊂ Q2. Then dist(Q2, Ωc) > √22−n+1 and so Q3 ⊂ Ω for all Q3 ∈ Qn−1 for which Q2 ∩ Q3 6= ∅. Thus Q2 ∈ Fn−1 or Q2 ⊂ Q4 ∈ Fi for some i < n − 1. In either case, Q /∈ Fn giving a contradiction. Property (W3) holds by the recursion in the definition and the fact that the dyadic squares are nested. Suppose (W4) is not true. Then there exist Q1 ∈ Fn and Q2 ∈ Fm with n < m − 1 and Q1 ∩ Q2 6= ∅. Let Q3 ∈ Fn+1 be such that Q2 ⊂ Q3. Then Q′ ⊂ Ω Q′ ⊂ [Q′∈Qn [Q′∈Qn+1 Q′∩Q36=∅ Q′∩Q16=∅ and so either Q3 ∈ Fn+1 or Q3 ⊂ Fn. In both cases Q2 ⊂ Fn+1 and so Q2 /∈ Fm. (cid:3) A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 5 PSfrag replacements Q0 Ω Γ′ eA′ eϕ−1(eA0) eϕ−1(eA) eϕ−1(T (eA0)) Figure 1. A core part Dn is selected from the Whitney decomposition of Ω by taking the connected component containing Q0 of the interior of the union of Whitney squares with sidelength at least 2−n. By a chain of dyadic squares {Qi}m i=1 we mean a collection of sets Qi ∈ F such that Qi ∩ Qi+1 is a non-degenerate line segment for all i ∈ {1, . . . , m − 1}. We say that the chain connects Q1 and Qm. 2.2. Approximating polynomials. We record here the following two Lemmas from [8] which will be used when estimating the approximation in Section 4. Lemma 2.4 (Lemma 2.1, [8]). Let Q be any square in R2 and P be a polynomial of degree k defined in R2. Let E, F ⊂ Q be such that E,F > ηQ where η > 0. Then kPkLp(E) ≤ C(η, k)kPkLp(F ). Given a function u ∈ C ∞(Ω) and a bounded set E ⊂ Ω, we define (see [8]) the polynomial approximation of u in E , Pk(u, E) to be the polynomial of order k − 1 which satisfies ZE ∇α(u − Pk(u, E)) = 0 for each α = (α1, α2) such that α = α1 + α2 ≤ k − 1. Once k is fixed, we denote the polynomial approximation of u in a dyadic square Q as PQ The next lemma is a consequence of Poincar´e inequality for Lipschitz domains. Lemma 2.5 (Lemma 3.1, [8]). Let Ω ⊂ R2 be a bounded simply connected domain and F i=1 in F be a chain of a Whitney decomposition of Ω. Fix α such that α ≤ k. Let {Qi}m dyadic squares in F .Then we have k∇α(PQ1 − PQm)kLp(Q1) ≤ Cl(Q1)k−αk∇kukLp(Sm i=1 Qi), where ∇ku is the vector (∇αu)α=k normed by the ℓ2-norm and C = C(m). In what follows, given β = (β1, β2) and α = (α1, α2), we write β ≤ α if the inequality holds coordinate-wise. 6 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ 3. Decomposition of the domain From now on we fix a bounded simply connected domain Ω ⊂ R2 and a Whitney decom- position F of Ω given by Definition 2.2. For our purposes we need to choose at each level a nice enough subcollection of Fn, namely we take connected components of the Whitney decomposition (see Figure 1). More precisely we fix Q0 ∈ F1 and for each n ∈ N let Cn be the connected component of the interior of Fn that has int Q0 as a subset. We define and using this the families of squares Fn,j := {Q ∈ Fj : int Q ⊂ Cn} Fn := Fn,n, Dn := [j≤n Fn,j and the corresponding sets for two of the above collections by Fn := [Q∈Fn Q and Dn := [Q∈Dn Q = C n. The collection of boundary layer squares in Dn is denoted by ∂Dn :=nQ ∈ Dn : Q ∩ (Ω \ Dn) 6= ∅o . With this notation we have the following lemma. Lemma 3.1. The above collections have the properties: (i) Dn ⊂ Dn+1 for all k ∈ N. (ii) Ω =Sn∈N Dn. (iii) If Q1, Q2 ∈ Fn and Q1 ∩ Q2 is a singleton, then there exists Q3 ∈ Dn for which (iv) If Q ∈ ∂Dn, then Q ∈ Fn. (v) If Q ∈ ∂Dn, then Q ∩ (Ω \ Fn) 6= ∅. (vi) The set Cn is simply connected. Q1 ∩ Q2 ∩ Q3 6= ∅. Proof. The property (i) is obvious by the definitions of Fn,j and Dn since Cn ⊂ Cn+1. For (ii) it suffices to prove that for every Q ∈ F there exists n ∈ N so that Q ∈ Dn. Let Q ∈ Fn. Since Ω is connected and open, there exists a path γ in Ω joining Q to Q0. By the fact that Fj ⊂ int Fj+1 and the property (W1) of the decomposition F we have that Ω = ∪j∈Nint Fj. Then by the compactness of γ there exists m ≥ n so that γ ⊂ int Fm. Hence Q ∈ Dm. For (iii) let Q1, Q2 ∈ Fn be so that Q1 ∩ Q2 is a singleton {q}. Assume that the claim is false. Then for the two squares Q ∈ Qn that intersect both Q1 and Q2 it is true that Q 6∈ Fn and Q 6⊂ Q′ for all Q′ ∈ Fn−1. Let q1 and q2 be the centres of the squares Q1 and Q2 respectively. Consider a curve γ′ : [0, 1] → Ω for which γ′ 1 = q1 and γ′ ⊂ Cn. Such a curve exists by the definition of Cn. We may also assume that γ′ is an injective 0 = q1, γ′ A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 7 PSfrag replacements Q1 q1 Q q Q2 γ q2 eϕ−1(eA0) eϕ−1(eA) eϕ−1(T (eA0)) Figure 2. The constructed Jordan curve γ in the proof of Lemma 3.1 ((iii)) has in its interior domain a dyadic square Q that also has to be an element of Dn. curve. Let t0 := sup{t : γ′(t) ∈ Q1} and t1 := inf{t ≥ t0 : γ′(t) ∈ Q2}. Define a Jordan curve γ := γ1 ∗ γ2 ∗ γ′[t0,t1]∗γ3 ∗ γ4, where γ1, γ2, γ3 and γ4 correspond to the line segments [q, q1], [q1, γ′ t1, q2] and [q2, q] respectively. By Jordan curve theorem γ divides R2 into two components, one of which is precompact (see Figure 2). Denote the precompact component by A. t0], [γ′ For small enough ball B around q we have by the definition of γ that B \ γ has exactly two components. Since γ is a Jordan curve one of those components has to contain an interior point of A and thus the whole component lies inside A. On the other hand that component has to intersect with one of the dyadic squares in Qn touching both Q1 and Q2 (but being different from Q1 and Q2). Let Q ∈ Qn be that square. Now for all the neighbouring squares Q ∈ Qn (except the opposite one) of Q either Q ∩ γ([0, 1]) 6= ∅ implying that Q ∈ Dn or Q is in the precompact component of R2 \ γ([0, 1]) and thus by simply connectedness Q ⊂ Ω. Since Q1 ∈ Fn, also the opposite square of Q is a subset of Ω. Hence Q ∈ Fn or Q ⊂ Q′ ∈ Fn−1 which is a contradiction. Thus we have proven (iii). In order to see (iv), suppose that there exists Q ∈ ∂Dn such that Q /∈ Fn. Then Q ∈ Fn,i ⊂ Fi for some i < n. By Property (W4), for all the Q′ ∈ F with Q′ ∩ Q 6= ∅ we have Q′ ∈ Fj for j ≤ i + 1 ≤ n. Thus, Q′ ⊂ Dn and Q /∈ ∂Dn giving a contradiction. If property (v) fails for some Q ∈ ∂Dn, then for every Q′ ∈ F with Q′ ∩ Q 6= ∅ we have Q′ ∈ Fi for some i ≤ n. Thus, again Q′ ⊂ Dn and Q /∈ ∂Dn giving a contradiction. Finally, we prove property (vi). Since Cn is open it suffices to prove that every Jordan curve is loop homotopic to a constant loop. Suppose this is not the case. Then there exists a Jordan curve γ that is not homotopic to a constant loop, and a point x ∈ Ω \ Cn that lies inside γ. In particular there exists Q ∈ Qn such that Q 6⊂ Dn which lies inside γ and for which Q ∩ Dn is an edge of a square. Now by similar argument as in (iii) we conclude that Q ∈ Dn, which is a contradiction. (cid:3) 8 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ The next lemma shows that we can connect the boundary of Dn to the boundary of Ω with a short curve in the complement of Dn. Lemma 3.2. For each point x ∈ ∂Dn, there exists an injective curve γ : [0, 1] → R2 so that γ(0) = x, γ(1) ∈ ∂Ω, γ(0, 1) ⊂ Ω \ int Dn and L(γ) ≤ 2√2l(Q). Proof. Let Q ∈ ∂Dn be such that x ∈ Q ∩ ∂Dn. By Lemma 3.1 (v) we have that there exists a square Q′ ∈ Qn touching Q at x so that Q′ /∈ Fn and Q′ 6⊂ Q for every Q ∈ Fj, j < n. Thus, there exists a neighbouring square Q′′ ∈ Qn of Q′ and a point y ∈ ∂Ω ∩ Q′′. Let γ1 be a curve corresponding to a line segment connecting x to a point z ∈ Q′ ∩ Q′′ and let γ2 be a curve corresponding to a line segment connecting z to y. Moreover, let t0 := inf{t : γ2 t0 ∈ ∂Ω. Define a curve γ := γ1∗ γ2[0,t0]. For γ we have that γ(0, 1) ⊂ (Ω\ int Dn)∩ (Q′∪ Q′′), γ(0) = x, γ(1) ∈ ∂Ω and L(γ) ≤ d(Q′) + d(Q′′) = 2√2l(Q). (cid:3) Observe that by Lemma 3.1 (iv) we have ∂Dn =SQ∈∂Dn(Q∩∂Dn). Thus, by Lemma 3.1 (iii) we have that ∂Dn is locally homeomorphic to the real line. Since by Lemma 3.1 (vi) Cn is simply connected, we have that ∂Dn = ∂Cn is connected. Hence, ∂Dn is a Jordan curve. Thus, we may write t ∈ ∂Ω}. Since ∂Ω is closed, we have that γ2 ∂Dn = Ln[i=1 Ii, (3.1) where Ii = [yi, yi+1] is an edge of a square in Fn with vertices yi and yi+1, and y1 = yLn+1. For the rest of the paper we fix a constant M > (4√2 + 2). However, the following lemma is true for any M > 0 and with C depending on M. Lemma 3.3. There exists C ∈ N so that for any n ∈ N and x, y ∈ ∂Dn with d∂Dn(x, y) ≥ 2−nC, and for any γ in Ω\int Dn connecting x to y we have that γ∩(Ω \ B(x, M2−n)) 6= ∅. In particular, L(γ) ≥ M2−n. Proof. By taking a slightly larger C, namely C + 2, we may assume that x = yi and y = yj for some i and j, where yi, yj are two endpoints of intervals from the collection {Ii} forming the boundary as noted above. Moreover, by symmetry we may assume that i < j and j − i ≤ n + 1 − j. Since each Ii is a side for two squares in Qn, by taking C large enough, we obtain H2(B(x, 2(M + 1)2−n)) = π(2(M + 1)2−n)2 < 1 2 C(2−n)2 ≤ H2([ Q), where the union is taken over all Q ∈ Qn having Im as one of it sides for some i < m ≤ j−1. Therefore, one of the intervals Im1, for i < m1 ≤ j−1, has to intersect with the complement of the ball B(x, 2M2−n). Let Q′ 1 ∈ ∂Dn be the boundary square corresponding to that interval and let q1 ∈ Im1 \ B(x, 2M2−n). By symmetry, there also exists Q′ 2 ∈ ∂Dn whose side is some Im2 with m2 /∈ {i+1, i+1, . . . , j−1} such that there is q2 ∈ Im2 \B(x, 2M2−n). Suppose now that there exists a curve γ in Ω\int Dn joining x to y with γ ⊂ B(x, M2−n). We may assume that γ is injective, and by compactness that γ(t) ∈ Ω \ Dn for every t ∈ (0, 1). Then, for i = 1, 2 we have that B(Q′ i)) ⊂ B(q, M2−n) and hence i, 2√2l(Q′ A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 9 Q′′ 2 Q′ 2 Q′ 1 Q′′ 1 PSfrag replacements x y γ D3 i there is a neighbouring square Q′′ 1 or Q′′ 1) will be a subset of Dn. Figure 3. In the proof of Lemma 3.3 we assume towards a contradiction that x and y can be connected by a short curve γ in Ω\ Dn. This will imply that one more square in Qn (here Q′′ i, 2√2l(Q′)) ∩ γ = ∅. Now by definition of Q′ i which is not a subset of Dn, see Figure 3. We claim that either Q′′ B(Q′ i ∈ Qn of Q′ 2 lies inside the Jordan curve γ′ obtained by concatenating the curve γ and the part of the boundary, denoted by γ′′, obtained from the intervals {Ih}j−1 h=i, or by concatenating γ and ∂Dn \ γ′′. This can be seen in the following way. Consider Ω h−→ R2 ֒→ S2, where h is a homeomor- phism and the inclusion R2 ֒→ S2 is the inverse of the stereographic projection. Under this composite map S2\Dn is a simply connected domain. Hence, by Lemma 2.1 (S2\Dn)\γ has exactly two components whose boundaries are the two connected components of ∂Dn \ γ together with γ. Thus, (Ω \ Dn) \ γ = (S2 \ Dn) \ γ has exactly two components. Since ∂Q′′ 1 ∩ ∂Dn and ∂Q′′ 2 ∩ ∂Dn are in two different connected components of ∂Dn \ γ, we conclude that Q′′ 2 are in different components of (Ω \ Dn) \ γ. We denote the Q′′ that lies inside the Jordan curve by Q′′. Since Q′′ ⊂ B(Q′,√2l(Q′)), we have that every neighbouring square of Q′′ either lies inside γ′ or is an element of ∂Dn. In particular, by the simply connectedness of Ω they all are subsets of Ω. Hence, Q′′ ⊂ Dn which is a contradiction. Thus, we have proven that γ ∩ (Ω \ B(x, M2−n)) 6= ∅. (cid:3) Let us now partition Ω \ Dn in the following way. Recall (3.1). Notice that for large enough n we have that Ln ≥ 2C. Define x1 := y1 and then xm := y(m−1)C until Ln+1−(m− 1 and Q′′ i 10 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ H7 x8 γ8 H6 γ7 x7 γ6 x6 H5 γ5 x5 H4 x4 γ4 H3 x3 γ3 D3 x1 γ1 H8 H1 x2 γ2 H2 Figure 4. Here the domain Ω is decomposed into the core part D3 and eight boundary parts Hi. A neighbourhood H8 of H8 is also illustrated. 1)C < 2C. Notice that for every i 6= j we have d∂Dn(xi, xj) ≥ 2−nC. We now partition the set Ω \ Dn up to Lebesgue measure zero into connected sets { Hj}m j=1 where Hj is the open set bounded by γj, γj+1 given by Lemma 3.2 for points xj and xj+1, and Jj := SC(j+1) Ii (with interior in Ω \ Dn). This partition is well defined by Lemma 2.1. Notice that since L(γi) ≤ M for all i, we have that γi∩γj = ∅ for all i 6= j. Let us define Hj as the connected component containing Hj of the set Ω ∩(cid:16) Hj ∪ BR2(γj ∪ γj+1 ∪ Jj, δ)(cid:17), where δ = 2−n−3. See Figure 4 for an illustration of the decomposition. Although the decomposition depends on n, for simplicity we do not display the dependence in the notation. A crucial property of our decomposition is the following lemma. Lemma 3.4. We have Hj ∩ Hi 6= ∅, if and only if i − j ≤ 1 in a cyclic manner. Proof. Trivially γi+1 ∈ Hi ∩ Hi+1. Thus, we only need to show that Hj ∩ Hi 6= ∅ implies i − j ≤ 1. We may assume that i 6= j. Let x ∈ Hi ∩ Hj. Suppose first that x ∈ Hi. Then, by (path) connectedness of Hj there exists a path γ in Hj from x to Hj. Let i=Cj t0 := inf{t ∈ [0, 1] : γ(t) /∈ Hi}. Suppose now that x ∈ Dn. Since δ < 2−n Then, γ(t0) /∈ Hi but γ(t) ∈ Hi∩Hj. Thus it suffices to consider the case when x /∈ Hi∪ Hj. 2 , we have that x ∈ Q for some Q ∈ ∂Dn. Then, there are neighbouring squares Qi, Qj ∈ Qn of Q for which Qi ∩ Hi 6= ∅ and Qj ∩ Hj 6= ∅. Since δ is small, we may choose the Qi, Qj so that Qi ∩ Qj 6= ∅. If Qi = Qj or if Qi and Qj have a common edge, then there is a curve γ′ in Qi ∪ Qj from Hi to Hj with L(γ′) < 2δ. If Qi ∩ Qj is a singleton, then by Lemma 3.1 (iii) the neighbouring square Q′ 6= Q of both A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 11 Qi and Qj lies in Ω \ int Dn. Indeed, if this were not the case, then Q′, Q ∈ Fn and Q′ ∩ Q is a singleton, implying that Qi ∈ Dn or Qj ∈ Dn. Thus, there exists a curve γ′ in Ω \ Dn joining Qi and Qj with L(γ′) < 4δ. Now, we have Qi ∩ Ji 6= ∅ or Qi ∩ (γi ∪ γi+1) 6= ∅. Notice that γi ∩ Ji 6= ∅ 6= γi+1 ∩ Ji. By Lemma 3.2 we have max(l(γi), l(γi+1)) ≤ 2√2 · 2−n. Combining these observations with the analogous ones for Qj, we have that Ji and Jj can be connected by a curve in Ω \ Dk with length less than 4δ + 4√2 · 2−n < 2−nM. Hence, we have by Lemma 3.3 that dist∂Dn(Ji, Jj) ≤ C. Thus, i − j ≤ 1 in cyclical manner. We are left with the case where x ∈ Ω \ (Dn ∪ Hi ∪ Hj). By definition we have that B(Dn, 2δ) ⊂ Ω. Thus, if dist(x, Ji) < δ, we may join x to Ji by a curve in Ω \ int Dn with length less than δ. If dist(x, Ji) ≥ δ, then x ∈ B(γm, δ), where m ∈ {i, i + 1}. By path connectedness of Hi there is a curve γ in Hi connecting x to γi ∪ γi+1 ∪ Ji. We want to prove that x can be joined to γm in δ-neighbourhood of γm. If (a subcurve of) γ is not such a curve, then we may define t0 := inf{t ∈ [0, 1] : γ(t) ∈ B(Dn, δ)}. Then, γ[0,t0]⊂ B(γm, δ). Therefore, there exists a point y ∈ γm with d(γ(t0), y) < δ. In particular, the line segment [γ(t0), y] lies in (Ω \ Dn) ∩ B(γm, δ) and thus we have proven that there exists a curve γ′ in (Ω\ Dn)∩ B(γm, δ) connecting x to γm. By the definition of γm we have that γ′ ⊂ B(γm(0), 2√2·2−n +δ). By the same argument for j we conclude that Ji and Jj can actually be connected by a curve γ in (Ω \ int Dn)∩ B(γ(0), 4√2 · 2−n + 2δ). Hence, by Lemma 3.3 dist∂Dn(Ji, Jj) < C, and thus i − j ≤ 1 in cyclical manner. (cid:3) 4. Approximation In this section we finish the proof of Theorem 1.1 by making a partition of unity using the decomposition of Ω constructed in Section 3 and by approximating a given function by polynomials in this decomposition. Recall that our aim is to show that for any u ∈ Lk,p(Ω) and ǫ > 0 there exists a function uǫ ∈ W k,∞(Ω) ∩ C ∞(Ω) with k∇ku − ∇kuǫkLp(Ω) . ǫ. By noting that Lk,p(Ω) ∩ C ∞(Ω) is dense in Lk,p(Ω) we may assume that function u ∈ Lk,p(Ω) ∩ C ∞(Ω). From now on, let u and ǫ > 0 be fixed. Using the notation from Section 3, we write the domain Ω as the union of the core part i=1. For each Hi we let Ii be the collection of squares Q Dn and the boundary regions {Hi}l in ∂Dn such that Q ∩ Hi 6= ∅, which are bounded in number independently of n. We need to decide what polynomial to attach to each set Hi. For this purpose, for each 1 ≤ i ≤ l we assign a square Qi ∈ Ii. We call Qi the associated square of Hi. Given Q ∈ Ii we set PQ := Si+1 j=i−1{Q′ ∈ Ij}, which is a collection of squares from a suitable neighbourhood of Q. Recall the approximating polynomials PQ introduced in Section 2.2. We abbreviate Pi = PQi for the associated squares Qi. We make a smooth partition of unity by using a Euclidean mollification. (Compare to [14] where the inner distance in Ω was used for the mollification.) Let ρr denote a standard Euclidean mollifier supported in B(0, r). We start with a collection of functions { ψi}l i=0, 12 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ i=0 by setting ψi = ψi/Pl where ψ0 = χDn ∗ ρ2−n−5 and ψi = (cid:0)χ Hi ∗ ρ2−n−5(cid:1) Hi for i ≥ 1. Using this we obtain a partition of unity {ψi}l Now the partition of unity {ψi}l (1) The function ψ0 is supported in B(Dn, 2−n 10 ). (2) For i ≥ 1 the function ψi is supported in Hi. (3) For all i, 0 ≤ ψi ≤ 1. (4) P ψi ≡ 1 on Ω. (5) For all i, ∇αψi ≤ Cα2−nα for all multi-indeces α. We will fix n later such that for the function uǫ defined as i=0 satisfies the following. ψj. j=0 uǫ(x) := u(x)ψ0(x) + lXi=1 ψi(x)Pi(x) for x ∈ Ω, we have Note that uǫ = u on Dn−1; indeed Dn−1 ∩ ψi = ∅ for i ≥ 1, see Lemma 3.1 (iv). First of all, we consider only n large enough so that k∇ku − ∇kuǫkLp(Ω) < Cǫ. Now, we need to show that n can actually be chosen large enough so that also k∇kukLp(Ω\Dn−1) ≤ ǫ. So, we compute for Q ∈ Ii and α = k k∇kuǫkLp(Ω\Dn−1) ≤ Cǫ. k∇αuǫkLp(Q) ≤Xβ≤α +Xβ≤αXj (cid:18)ZQ ∇βu − ∇βPi(x)p∇α−βψ0(x)p dx(cid:19)1/p (cid:18)ZQ ∇βPj(x) − ∇βPi(x)p∇α−βψj(x)p dx(cid:19)1/p =: A1 + A2, (4.1) (4.2) (4.3) where A1 and A2 are the first and second terms on the right hand side of the inequality and we used that for β < α, Pj ∇α−βψj = 0 and order of Pi is at most k − 1. We first estimate A1 as 2n(α−β)k∇βu − ∇βPikLp(Q) 2n(α−β)(k∇βPi − ∇βPQkLp(Q) + k∇βu − ∇βPQkLp(Q)) 2n(α−β)2n(β−k)k∇kukLp(∪ Q) A1 .Xβ≤α .Xβ≤α .Xβ≤α . k∇kukLp(∪ Q), A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 13 where in the third inequality we used that Qi (associated square of Hi) and Q may be joined by a chain of bounded number of squares from Ii by our construction, and therefore we may apply Lemma 2.5. Similarly we estimate A2 as (cid:18)ZQ ∇βPj(x) − ∇βPi(x)p∇α−βψj(x)p dx(cid:19)1/p i+1Xj=i−1 (k∇βPj − ∇βPQkLp(Q) + k∇βPi − ∇βPQkLp(Q)) (4.4) 2n(α−β) i+1Xj=i−1 A2 .Xβ≤α .Xβ≤α .Xβ≤α . k∇kukLp(∪ Q), 2n(α−β)2n(β−k)k∇kukLp(∪ Q) where again in the second inequality we used that if ψj(x) 6= 0 for x ∈ Q ∈ Ii then by our construction Qj and Q can be joined by a chain of bounded number of squares as j is either i − 1, i or i + 1 (cyclically); and therefore we can apply Lemma 2.5. For Q ∈ F \ Dn such that Q ∩ spt(ψ0) 6= ∅, we assign to Q a square Q′ ∈ Ii, such that 6= ∅. Note that such a square Q′ exists by our construction. Then Q and Q′ can Q ∩ Q′ be joined by a chain of bounded (by an absolute constant) number of squares from Dn+1. We choose such a chain for Q and denote it by BQ. We also set Jn := {Q ∈ F \ Dn : Q ∩ spt(ψ0) 6= ∅}. We estimate using Lemma 2.5 exactly as above (see (4.2)) to obtain for α = k k∇αuǫkLp(Q) ≤Xβ≤α +Xβ≤αXj (cid:18)ZQ ∇βu − ∇βPQ(x)p∇α−βψ0(x)p dx(cid:19)1/p (cid:18)ZQ ∇βPj(x) − ∇βPQ(x)p∇α−βψj(x)p dx(cid:19)1/p (4.5) =: B1 + B2. Again, we estimate separately, 2n(α−β)k∇βu − ∇βPQkLp(Q) B1 .Xβ≤α . k∇kukLp(Q) 14 and DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ (cid:18)ZQ ∇βPj(x) − ∇βPQ(x)p∇α−βψj(x)p dx(cid:19)1/p 2n(α−β) i+1Xj=i−1 (k∇βPj − ∇βPQ′kLp(Q) + k∇βPQ′ − ∇βPQkLp(Q)) 2n(α−β)2n(β−k)(k∇kukLp( S Q′′ ∈P Q′′) + k∇kukLp( S Q′′ ∈B Q′′) Q′ Q′ i+1Xj=i−1 B2 .Xβ≤α .Xβ≤α .Xβ≤α . k∇kukLp( Q′′). S Q′ S B Q′ Q′′ ∈P Next we note that ∇kuǫ ≡ 0 in Hi\ ∪j6=i spt(ψj) and we compute for α = k k∇αuǫkLp(Q) k∇αuǫkLp(Q) + XQ∈Jn,Q∩Hi6=∅ k∇αuǫkLp((spt(ψj )∩spt(ψi))\ (4.6) S Q′′) Q′′ ∈∂Dn T Jn The terms in the first and second summands have been estimated earlier. Denoting H ′ (∪i+1 j=i−1spt(ψj) ∩ spt(ψi)) \ Q′′, we estimate now the third one; i := where we used the facts that for β < α, ∇α−βPj(ψj) = 0 and ψ0 ≡ 0 in H ′ inequality, Lemma 2.4 in the second inequality since H ′ C coming from Lemma 3.2 and in the third inequality we used Lemma 2.5. Remark 4.1. Note that for each Q ∈ Ii we have PQ = PQi where Qi is the associated square of Hi. We note that any Q′ ∈ ∂Dn occurs in at most three distinct collections PQi. Moreover any Q ∈ Dn+1 appears in only a bounded number of the collections BQ′′, where Q′′ ∈ Jn. In particular, any Q′ ∈ ∂Dn appears in only a bounded number of the collections BQ′′, where Q′′ ∈ Jn. The bounds are provided by absolute constants coming from volume comparison. i in the first i ⊂ CQi for some absolute constant + k∇αuǫkLp(Hi) ≤ XQ∈ Qi i+1Xj=i−1 SQ′′∈∂Dn T Jn i) .Xβ≤α .Xβ≤α .Xβ≤α . k∇kukLp( S k∇αuǫkLp(H ′ 2n(α−β) Q′), Q′ ∈PQi i+1Xj=i−1 k∇βPj − ∇βPikLp(H ′ i) 2n(α−β)(k∇βPj − ∇βPikLp(Qi) 2n(α−β)2n(β−k)k∇kukLp( S Q′ ∈PQi (4.7) Q′) A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 15 Now it follows from equations (4.3), (4.4), (4.5), (4.6) and (4.7) that k∇αuǫkLp(Ω\Cn) .Xi k∇kukLp(Hi) + k∇kukLp( S Q) + k∇kukLp( S Q∈∂Dn . k∇kukLp( S Q) Q∈∂Dn Q) + k∇kukLp( S Q∈Jn Q′∈BQ Q′) (4.8) Q∈Jn when α = k. By Remark 4.1 we may choose n such that k∇kukLp( S Q∈∂Dn Q) + k∇kukLp( S Q∈Jn Q) + k∇kukLp( S Q∈Jn Q′∈BQ Q′) < ǫ. Then, the claim follows from (4.1) and (4.8). Remark 4.2. We note that when k = 1 we may take the function to be smooth as well as bounded for showing the density of W 1,∞(Ω) in W 1,p(Ω). This is because truncations approximate the functions in W 1,p(Ω). This allows us to boundealso approximate the Lp norm of u. Indeed let u ∈ W 1,p(Ω) ∩ C ∞(Ω) ∩ L∞(Ω) such that kukL∞ ≤ M. De- compose the domain as in the above construction; then choose n large enough such that kukW 1,p(Ω\Dn−1) ≤ ǫ and MΩ\ Dn−1 < ǫ. Then it follows from estimates in the proof that the function uǫ defined as above approximates u in W 1,p(Ω) with error given by ǫ. This conclusion is the content of [14]. Finally, let us show how the smooth approximation in Jordan domains is done. Proof of Corollary 1.3. The argument we need follows the one used to prove [14, Corollary 1.2]. As in [14], given a bounded Jordan domain we approximate it from outside by a nested sequence of Lipschitz and simply connected domains Gs which are obtained for example by taking the complement of the unbounded connected component of the union Whitney squares larger than 2−s from the Whitney decomposition of the complementary Jordan domain of Ω. Then, we note that for given n, taking sn large enough, we have that the squares in ∂Dn are Whitney type sets in Gsn, meaning they have diameters comparable to the distance from the boundary of Gsn. Note that Gsn ⊂ B(Ω, 2−sn+5) are simply connected. Now the set Gsn \ ¯Cn (recall that Cn is a suitable connected component of the interior of the union of the Whitney squares of scale less than 2−n) can be decomposed in the same way as Ω\ ¯Cn was decomposed into the sets Hi in Section 3. We may then follow the argument used in the proof of Theorem 1.1 to obtain an approx- imating sequence of functions un in Gsn which are in the space W k,∞(Gsn) ∩ Lk,p(Gsn) ∩ C ∞(Gsn). By multiplying with a smooth cut-off function that is 1 on Ω and compactly supported in Gsn, we obtain a sequence of global smooth functions having the desired properties. (cid:3) 16 DEBANJAN NANDI, TAPIO RAJALA, AND TIMO SCHULTZ References 1. Charles J. Amick, Approximation by smooth functions in Sobolev spaces, Bull. London Math. Soc. 11 (1979), no. 1, 37 -- 40. MR 535794 2. Christopher J. Bishop, A counterexample concerning smooth approximation, Proc. Amer. Math. Soc. 124 (1996), no. 10, 3131 -- 3134. MR 1328340 3. Stephen M. Buckley and Pekka Koskela, Criteria for imbeddings of Sobolev-Poincar´e type, Internat. Math. Res. Notices (1996), no. 18, 881 -- 901. MR 1420554 4. Alessandro Giacomini and Paola Trebeschi, A density result for Sobolev spaces in dimension two, and applications to stability of nonlinear Neumann problems, J. Differential Equations 237 (2007), no. 1, 27 -- 60. MR 2327726 5. V. M. Gol'dshteın and Yu. G. Reshetnyak, Quasiconformal mappings and Sobolev spaces, Mathematics and its Applications (Soviet Series), vol. 54, Kluwer Academic Publishers Group, Dordrecht, 1990, Translated and revised from the 1983 Russian original, Translated by O. Korneeva. MR 1136035 6. V. M. Gol'dsteın, T. G. Latfullin, and S. K. Vodop'janov, A criterion for the extension of functions 2 from unbounded plane domains, Sibirsk. Mat. Zh. 20 (1979), no. 2, 416 -- 419, 464. of the class L1 MR 530508 7. Vladimir Gol'dsteın and Serge Vodop'anov, Prolongement de fonctions diff´erentiables hors de domaines plans, C. R. Acad. Sci. Paris S´er. I Math. 293 (1981), no. 12, 581 -- 584. MR 647686 8. Peter W. Jones, Quasiconformal mappings and extendability of functions in Sobolev spaces, Acta Math. 147 (1981), no. 1-2, 71 -- 88. MR 631089 9. Torbjorn Kolsrud, Approximation by smooth functions in Sobolev spaces, a counterexample, Bull. London Math. Soc. 13 (1981), no. 2, 167 -- 169. MR 608104 10. Pekka Koskela, Extensions and imbeddings, J. Funct. Anal. 159 (1998), no. 2, 369 -- 383. MR 1658090 11. 12. Pekka Koskela, Tapio Rajala, and Yi Ru-Ya Zhang, A geometric characterization of planar sobolev , Removable sets for Sobolev spaces, Ark. Mat. 37 (1999), no. 2, 291 -- 304. MR 1714767 extension domains, Preprint. 13. , A density problem for Sobolev spaces on Gromov hyperbolic domains, Nonlinear Anal. 154 (2017), 189 -- 209. MR 3614650 14. Pekka Koskela and Yi Ru-Ya Zhang, A density problem for Sobolev spaces on planar domains, Arch. Ration. Mech. Anal. 222 (2016), no. 1, 1 -- 14. MR 3519964 15. John L. Lewis, Approximation of Sobolev functions in Jordan domains, Ark. Mat. 25 (1987), no. 2, 255 -- 264. MR 923410 16. Norman G. Meyers and James Serrin, H = W , Proc. Nat. Acad. Sci. U.S.A. 51 (1964), 1055 -- 1056. MR 0164252 17. M. H. A. Newman, Elements of the topology of plane sets of points, Cambridge, At the University Press, 1951, 2nd ed. MR 0044820 18. Anthony G. O'Farrell, An example on Sobolev space approximation, Bull. London Math. Soc. 29 (1997), no. 4, 470 -- 474. MR 1446566 19. Pavel Shvartsman, On Sobolev extension domains in Rn, J. Funct. Anal. 258 (2010), no. 7, 2205 -- 2245. MR 2584745 20. Wayne Smith, Alexander Stanoyevitch, and David A. Stegenga, Smooth approximation of Sobolev functions on planar domains, J. London Math. Soc. (2) 49 (1994), no. 2, 309 -- 330. MR 1260115 21. Wayne Smith and David A. Stegenga, Holder domains and Poincar´e domains, Trans. Amer. Math. Soc. 319 (1990), no. 1, 67 -- 100. MR 978378 22. Elias M. Stein, Singular integrals and differentiability properties of functions, Princeton Mathematical Series, No. 30, Princeton University Press, Princeton, N.J., 1970. MR 0290095 23. Hassler Whitney, Analytic extensions of differentiable functions defined in closed sets, Trans. Amer. Math. Soc. 36 (1934), no. 1, 63 -- 89. MR 1501735 A DENSITY RESULT FOR HOMOGENEOUS SOBOLEV SPACES ON PLANAR DOMAINS 17 Department of Mathematics and Statistics, P.O. Box 35 (MaD), FI-40014 University of Jyvaskyla, Finland E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected]
1603.01427
3
1603
2016-12-08T10:56:34
Splines are Universal Solutions of Linear Inverse Problems with Generalized-TV regularization
[ "math.FA" ]
Splines come in a variety of flavors that can be characterized in terms of some differential operator L. The simplest piecewise-constant model corresponds to the derivative operator. Likewise, one can extend the traditional notion of total variation by considering more general operators than the derivative. This leads us to the definition of the generalized Beppo-Levi space M, which is further identified as the direct sum of two Banach spaces. We then prove that the minimization of the generalized total variation (gTV) over M, subject to some arbitrary (convex) consistency constraints on the linear measurements of the signal, admits nonuniform L-spline solutions with fewer knots than the number of measurements. This shows that non-uniform splines are universal solutions of continuous-domain linear inverse problems with LASSO, L1, or TV-like regularization constraints. Remarkably, the spline-type is fully determined by the choice of L and does not depend on the actual nature of the measurements.
math.FA
math
Splines are universal solutions of linear inverse problems with generalized-TV regularization Michael Unser∗ Julien Fageot∗ December 9, 2016 John Paul Ward∗† Abstract Splines come in a variety of flavors that can be characterized in terms of some differential operator L. The simplest piecewise-constant model corresponds to the derivative operator. Likewise, one can extend the traditional notion of total variation by considering more general operators than the derivative. This results in the definition of a generalized total variation semi-norm and of its corresponding native space, which is further identified as the direct sum of two Banach spaces. We then prove that the minimization of the generalized total variation (gTV), subject to some arbitrary (convex) consistency constraints on the linear measurements of the signal, admits nonuniform L-spline solutions with fewer knots than the number of measurements. This shows that nonuniform splines are universal solutions of continuous-domain linear inverse problems with LASSO, L1, or total-variation-like regular- ization constraints. Remarkably, the type of spline is fully determined by the choice of L and does not depend on the actual nature of the measurements. Keywords. Sparsity, total variation, splines, inverse problems, compressed sensing AMS subject classifications. 41A15, 47A52, 94A20, 46E27, 46N20, 47F05, 34A08, 26A33 1 Introduction Imposing sparsity constraints is a powerful paradigm for solving ill-posed inverse problems and/or for reconstructing signals at super-resolution [6]. This is usually achieved by formulating the task as an optimization problem that includes some form of (cid:96)1 regularization [49]. The concept is central to the theory of compressed sensing (CS) [9, 19] and is currently driving the development of a new generation of algorithms for the reconstruction of biomedical images [36]. The primary factors that have contributed to making sparsity a remarkably popular research topic during the past decade are as follows: • the possibility of recovering the signal from few measurements (CS) with a theoretical guarantee of perfect recovery under strict conditions [7, 10, 19]; • the availability of fast iterative solvers for this class of problems [4, 14, 26, 39]; • the increasing evidence of the superiority of the sparsity-promoting schemes over the classi- cal linear reconstruction (including the Tikhonov (cid:96)2 regularization) in a variety of imaging modalities [36]. ∗Biomedical Imaging Group, École polytechnique fédérale de Lausanne (EPFL), CH-1015 Lausanne, Switzer- land †Department of Mathematics, University of Central Florida, Orlando, FL, USA. Present address: Department of Mathematics, North Carolina A&T State University, Greensboro, NC, USA 1 The approach developed in this paper is also driven by the idea of sparsity. However, it deviates from the standard paradigm because the recovery problem is formulated in the con- tinuous domain under the practical constraint of a finite number of linear measurements. The ill-posedness of the problem is then dealt with by searching for a solution that is consistent with the measurements and that minimizes a generalized version of the total-variation (TV) semi- norm-the continuous-domain counterpart of (cid:96)1 regularization. Our major finding (Theorem 1) is that the extremal points of this kind of recovery problem are nonuniform splines whose type is matched to the regularization operator L. The powerful aspect is that the result holds in full generality, as long as the problem remains convex. The only constraint is that the linear inverse problem should be well-posed over the (very small) null space of the regularization operator, which is the minimal requirement for any valid regularization scheme. In particular, Theo- rem 1 gives a theoretical explanation of the well-documented observation that total variation regularization-the simplest case of the present theory with L = D (derivative operator)-tends to produce piecewise-constant solutions [11, 41]. Recognizing the intimate connection between linear inverse problems and splines is also helpful for discretization purposes because it provides us with a parametric representation of the solution that is controlled by the regularization opera- tor L. In that respect, our representer theorems extend some older results on spline interpolation with minimum L1-norms, including the adaptive regression splines of Mammen and van de Geer [38] and the functional analytic characterization of Fisher and Jerome [27]. There is a connec- tion as well with the work of Steidl et al. on splines and higher-order TV [48], although their formulation is strictly discrete and restricted to the denoising problem. 2 Linear Inverse Problems: Current Status and Motivation Our notational convention is to use bold letters to denote ordinary vectors and matrices to distinguish them from their infinite-dimensional counterparts; that is, functions (such as s) and linear operators (such as L). Simply stated, the inverse problem is to recover a signal s from a finite set of linear measurements y = y0(s) + n ∈ RM where n is a disturbance term that is usually assumed to be small and independent of s. In most real-world problem the unknown signal lives in the continuum so that it is appropriate to view it as an element of some Banach space B. Then, by the assumption of linearity, there exists a set of functionals νm ∈ B(cid:48) (the continuous dual of B) with m = 1, . . . , M such that the noise-free measurements are given by y0 = ν(s) = ((cid:104)ν1, s(cid:105), . . . ,(cid:104)νM , s(cid:105)). The measurement functionals νm are governed by the underlying physics (forward model) and assumed to be known. Since the signal s ∈ B is an infinite-dimensional entity and the number of measurements is finite, the inverse problem is obviously ill-posed, not to mention the fact that the true measurements y are typically only approximate versions of y0 since they are corrupted by noise. 2.1 Finite-Dimensional Formulation The standard approach for the resolution of such inverse problems is to select some finite- dimensional reconstruction space V = span{ϕn}N n=1 ⊂ B. Based on the (simplifying) assumption that s ∈ V, one then converts the original noise-free forward model into the discretized version y0 = Ax, where x ∈ RN represents the expansion coefficients of s in the basis {ϕn}N n=1. Here, A is the so-called sensing matrix of size (M × N ) whose entries are given by [A]m,n = (cid:104)νm, ϕn(cid:105). The basic assumption made by the theory of compressed sensing is that there exists a finite- dimensional basis (or dictionary) {ϕn}N n=1 that "sparsifies" the class of desired signals with the property that (cid:107)x(cid:107)0 ≤ K0 for some fixed K0 which is (much) smaller than N; in other words, it 2 should be possible to synthesize the signal exactly by restricting the expansion to no more than K0 atoms in the basis {ϕn}N n=1 [20, 23, 40]. The signal recovery is then recast as the constrained optimization problem arg min x∈RN (cid:107)x(cid:107)1 s.t. (cid:107)y − Ax(cid:107)2 2 ≤ 2, (1) Instead of basing the recovery on the synthesis formula s =(cid:80) where the minimization of the (cid:96)1 norm promotes sparse solutions [49]. The role of the right- hand-side inequality is to encourage consistency between the noisy measurements y and their noise-free restitution y0 = Ax. The popularity of (1) stems from the fact that the theory of CS guarantees a faithful signal recovery from M > 2K0 measurements under strict conditions on A (i.e., restricted isometry) [7, 10, 19]. n xnϕn ∈ V, one can adopt an alternative analysis or regularization point of view. To that end, one typically assumes that s is discretized in some implicit "pixel" basis with expansion coefficients s = (s1, . . . , sN ) ∈ RN, where the sn are the samples of the underlying signal. The corresponding system matrix (forward model) is denoted by H : RN → RM. Given some appropriate regularization operator L : RN → RN(cid:48), the idea then is to exploit the property that the transformed version of the signal, Ls, is sparse. This translates into the optimization problem (cid:107)Ls(cid:107)1 s.t. (cid:107)y − Hs(cid:107)2 2 ≤ 2, arg min s∈RN (2) which is slightly more involved than (1). The two forms are equivalent only when N(cid:48) = N and L is invertible, the connection being A = HL−1. For computational purposes, (2) is often converted into the equivalent unconstrained version of the problem (cid:0)(cid:107)y − Hs(cid:107)2 2 + λ(cid:107)Ls(cid:107)1 (cid:1), arg min s∈RN (3) where λ ∈ R+ is an adjustable regularization parameter that needs to tuned such that (cid:107)y − Hs(cid:107)2 2 = 2. One of the preferred choices for L is the finite-difference operator-or the discrete version of the gradient in dimensions higher than one. This corresponds to the "total-variation" reconstruction method, which is widely used in applications [3, 11, 30, 41]. The sparsity-promoting effect of these discrete formulations and the conditions under which the expansion coefficients of the signal can be recovered are fairly well understood [28, 54]. What is less satisfactory is the intrinsic interdependence between the sparsity constraints and the choice of the appropriate reconstruction space, which makes it difficult to deduce rates of convergence and error estimates relating to the underlying continuous-domain recovery problem. Infinite-Dimensional Formulation the property that such a function admits the (unique) expansion s =(cid:80) 2.2 Recently, Adcock and Hansen have addressed the above limitation by formulating an infinite- dimensional theory of CS [1]. The measurements are the same as before, but the unknown signal is now a function s : Rd → R. For the purpose of illustration, we take d = 1 and s ∈ BV(R) with n wnψn in the (properly normalized) Haar wavelet basis {ψn}. It is known that the condition s ∈ BV(R) implies the inclusion of w = (wn) in weak-(cid:96)1(Z)-a space that is slightly larger than (cid:96)1(Z) [12]. Conversely, one can force the inclusion in BV(R) by imposing a bound on the (cid:96)1 norm of these coefficients. If one further assumes that the signal is sparse in the Haar basis, one can recast the reconstruction 3 problem as (cid:107)w(cid:107)1 s.t. min w∈(cid:96)1(Z) M(cid:88) m=1 (cid:12)(cid:12)ym − (cid:104)νm, (cid:88) n wnψn(cid:105)(cid:12)(cid:12)2 ≤ 2, (4) which is the infinite-dimensional counterpart of the synthesis formulation (1). The key question is to derive conditions on how to choose the νm to guarantee recovery of wavelet coefficients up to a certain scale. This has been done in [1, 2], which means that the issue of convergence is now reasonably well understood for the synthesis form of the recovery problem. In our framework, MD(R) is the space of functions on R of bounded (total) variation, which is slightly larger than BV(R) because it also includes constant signals. This allows us to close the circle by enforcing a regularization on the "true" total variation of the solution, which is associated with the derivative operator D = d dx. This results in the functional optimization problem s = arg min f∈MD(R) TV(f ) = (cid:107)Df(cid:107)M s.t. (cid:107)y − ν(f )(cid:107)2 2 = (cid:12)(cid:12)ym − (cid:104)νm, f(cid:105)(cid:12)(cid:12)2 ≤ 2, (5) M(cid:88) m=1 K(cid:88) which is the continuous-domain counterpart of (2). Now, the motivation for our present theory is that (5) corresponds to a special case of Theorem 1 with L = D and the closed compact convex set C ⊂ RM being specified by the inequality on the right-hand-side of (5); that is, C(y) = {z ∈ RM : (cid:107)y − z(cid:107)2 2 ≤ 2}. The key is that the differentiation operator D is spline- admissible in the sense of Definition 1: Its causal Green's function is the Heaviside (or unit-step) function ρD(x) = 1+(x) whose rate of growth is n0 = 0, while its null space ND = span{p1} with p1(x) = 1 is composed of all constant-valued signals. This implies that the extreme points of (5) necessarily take the form s(x) = b1 + ak1+(x − xk) (6) k=1 knots at the xk with the property that D{s} = (cid:80)K with K ≤ M. This corresponds to a piecewise-constant signal with jumps of size ak at the xk, as illustrated in Figure 1. The solution also happens to be a polynomial spline of degree 0 with k=1 akδ(· − xk) = wδ, which is a weighted sum of shifted Dirac impulses (the innovation of the spline), as shown on the bottom of Figure 1. In view of Definition 2, the solution (6) can also be described as a nonuniform L-spline with L = D. The remarkable aspect of this result is that the parametric form (6) is universal, in the sense that it does not dependent on the measurement functionals νm. To the best of our knowledge, this is the first mathematical explanation of the well-known observation that TV regularization tends to enforce piecewise-constant solutions. The other interesting point is that one can interpret the solution as the best K-term representation of the signal within an infinite- dimensional dictionary that consists of a constant signal p1 plus a continuum of shifted Green's functions (i.e., {1+(· − τ )}τ∈R), making the connection with the synthesis views (1) and (4) of the problem. Also, note that the described sparsifying effect is much more dramatic than that of the finite-dimensional setting since one is collapsing a continuum (integral representation) into a discrete and finite sum. We shall now show that the mechanism at play is very general and transposable to a much broader class of regularization operators L and data-fidelity terms, as well as for the multidi- mensional setting. 4 Figure 1: Prototypical solution of a linear inverse problem with total-variation regularization. The signal is piecewise-constant; in other words, it is a nonuniform L-spline with L = D (deriva- tive operator). The application of D uncovers the innovation wδ: The Dirac impulses are located at the points of discontinuity (knots), while their height (weight) encodes the magnitude of the corresponding jump. 2.3 Road Map of the Paper The remainder of this paper is organized as follows: After setting the notation, we present and discuss of our main representer theorem (Theorem 1) in Section 3. We also provide a refined version for the simpler interpolation scenario (Theorem 2). We then proceed with the review of primary applications in Section 4. The mathematical tools for proving our results are developed in the second half of the paper. The first enabling component is the tight connection between splines and operators, which is the topic of Section 5. In particular, we present an operator-based method to synthesize a spline from its innovation, which requires the construction of an appropriate right-inverse operator (Theorem 4). The existence of such inverse operators is fundamental to the characterization of the native spaces associated with our generalized total-variation criterion (gTV) (Theorem 5), as we show in Section 6. The actual proof of Theorems 1 and 2 is given in Section 7. It relies on a preparatory result (generalized Fisher-Jerome theorem) that establishes the impulsive form of the solutions of some abstract minimization problem over the space M(Rd) of bounded Borel measures. We conclude the paper in Section 8 with a brief discussion of open issues. 3 Representer Theorems for Generalized Total Variation Although we are considering a finite number of measurements, we are formulating the reconstruc- tion problem in the continuous domain. This calls for a precise specification of the underlying functional setting. 5 a1x1s(x)w(x)=ddxs(x)xx 3.1 Notation The space of tempered distribution is denoted by S(cid:48)(Rd) where d gives the number of dimensions. This space is made of continuous linear functionals µ : ϕ (cid:55)→ (cid:104)µ, ϕ(cid:105) acting on the Schwartz' space S(Rd) of smooth and rapidly decaying test functions on Rd [29, 33]. We shall primarily work with the space M(Rd) of regular, real-valued, countably additive Borel measures on Rd, which is also known (by the Riesz-Markov theorem) to be the continuous dual of C0(Rd): the Banach space of continuous functions on Rd that vanish at infinity equipped with the supremum norm (cid:107) · (cid:107)∞ [42, Chap. 6]. Since S(Rd) is dense in C0(Rd), this allows us to define M(Rd) as M(Rd) = {w ∈ S(cid:48) (Rd) : (cid:107)w(cid:107)M = sup ϕ∈S(Rd):(cid:107)ϕ(cid:107)∞=1 (cid:104)w, ϕ(cid:105) < ∞}, (7) by ϕ (cid:55)→ (cid:104)w, ϕ(cid:105) =(cid:82) and also to extend the space of test functions to ϕ ∈ C0(Rd). The action of w will be denoted Rd ϕ(x)w(x)dx where the right-hand side stands for the Lebesgue integral of ϕ with respect to the underlying measure1. The bottom line is that M(Rd) is the Banach space associated with the norm (cid:107)·(cid:107)M which returns the "total variation" of the measure that specifies w. Two key observations in relation to our goal are: 1. the compatibility of the L1 and total-variation norms with the former being stronger than the latter. Indeed, (cid:107)f(cid:107)L1(Rd) = (cid:107)f(cid:107)M for all f ∈ L1(Rd); 2. the inclusion of Dirac impulses in M(Rd), but not in L1(Rd). Specifically, δ(· − x0) ∈ M(Rd) for any fixed offset x0 ∈ Rd with (cid:104)δ(· − x0), ϕ(cid:105) = ϕ(x0) for all ϕ ∈ C0(Rd). We shall monitor the algebraic rate of growth/decay of (ordinary) functions of the variable x ∈ Rd by verifying their inclusion in the Banach space L∞,α0(Rd) = {f : Rd → R s.t. (cid:107)f(cid:107)∞,α0 < +∞}, (8) −α0(cid:1) (cid:0)f (x)(1 + (cid:107)x(cid:107)) where (cid:107)f(cid:107)∞,α0 = ess sup x∈Rd ··· xmd d ∈ L∞,α0(Rd) for α0 ≥ m = m1 + ··· + md. with α0 ∈ R. For instance, xm = xm1 A linear operator whose output is a function is represented with a roman capital letter (e.g., L). The action of L on the signal s is denoted by s (cid:55)→ L{s}, or Ls for short. Such an operator is said to be shift-invariant if it commutes with the shift operator s (cid:55)→ s(· − x0); that is, if L{s(· − x0)} = L{s}(· − x0) for any admissible signal s and x0 ∈ Rd. 1 3.2 Main Result on the Optimality of Splines Since the solution is regularized, the constrained minimization is performed over some native space ML(Rd) that is tied to some admissible differential operator L, such as D, D2 (second derivative), or ∆ (Laplacian) for d > 1. Definition 1 (Spline-admissible operator). A linear operator L : ML(Rd) → M(Rd), where ML(Rd) ⊃ S(Rd) is an appropriate subspace of S(cid:48)(Rd), is called spline-admissible if 1The use of w(x)dx in the integral is a slight abuse of notation when the measure is not absolutely continuous with respect to the Lebesgue measure. 6 1. it is shift-invariant; 2. there exists a function ρL : Rd → R of slow growth (the Green's function of L) such that L{ρL} = δ, where δ is the Dirac impulse. The rate of polynomial growth of ρL is n0 = inf{n ∈ N : ρL ∈ L∞,n(Rd)}. 3. the (growth-restricted) null space of L, NL = {q ∈ L∞,n0(Rd) : L{q} = 0}, has the finite dimension N0 ≥ 0. The native space of L, ML(Rd), is then specified as ML(Rd) = {f ∈ L∞,n0(Rd) : (cid:107)Lf(cid:107)M < ∞}. (9) It is largest function space for which the generalized total variation gTV(f ) = (cid:107)Lf(cid:107)M is well-defined under the finite-dimensional null-space constraint (cid:107)Lf(cid:107)M = 0 ⇔ f ∈ NL, for any f ∈ ML(Rd). This also means that gTV is only a semi-norm on ML(Rd). However, it can be turned into a proper norm by factoring out the null space of L. We rely on this property and the finite dimensionality of NL to prove that ML(Rd) is a bona fide Banach space (see Theorem 5). Having set the functional context, we now state our primary representer theorem. Theorem 1 (gTV optimality of splines for linear inverse problems). Let us assume that the following conditions are met: inition 1. 1. The regularization operator L : ML(Rd) → M(Rd) is spline-admissible in the sense of Def- 2. The linear measurement operator ν : f (cid:55)→ ν(f ) =(cid:0)(cid:104)ν1, f(cid:105), . . . ,(cid:104)νM , f(cid:105)(cid:1) maps ML(Rd) → RM and is weak*-continuous on ML(Rd) =(cid:0)CL(Rd)(cid:1)(cid:48). 3. The recovery problem is well-posed over the null space of L: ν(q1) = ν(q2) ⇔ q1 = q2, for any q1, q2 ∈ NL. Then, the extremal points of the general constrained minimization problem β = min f∈ML(Rd) (cid:107)Lf(cid:107)M s.t. ν(f ) ∈ C, (10) where C is any (feasible) convex compact subset of RM, are necessarily nonuniform L-splines of the form s(x) = akρL(x − xk) + bnpn(x) (11) with parameters K ≤ M, {xk}K RN0. Here, {pn}N0 The full solution set of (10) is the convex hull of those extremal points. n=1 is a basis of NL and L{ρL} = δ so that β = (cid:107)Ls(cid:107)M =(cid:80)K k=1 with xk ∈ Rd, a = (a1, . . . , aK) ∈ RK, and b = (b1, . . . , bN0) ∈ k=1 ak = (cid:107)a(cid:107)1. n=1 K(cid:88) k=1 N0(cid:88) 7 The key property here is L{s} = (cid:80)K Theorem 1 is a powerful existence result that points towards the universality of nonuniform k=1 akδ(· − xk), which follows from L-spline solutions. Conditions 1-3 in Definition 1 and is consistent with the more detailed characterization of splines presented in Section 5. For the time being, it suffices to remark that these splines are smooth (i.e., infinitely differentiable) everywhere, except at their knot locations {xk}. Although the extremal problem is defined over a continuum, the remarkable outcome is that the problem admits solutions that are intrinsically sparse, with the level of sparsity being measured by the minimum number K of required spline knots. In particular, this explains why the solution of a problem with a TV/L1-type constraint on the derivative (resp., the second derivative) is piecewise-constant (resp., piecewise linear when L = D2) with breakpoints at xk. The other pleasing aspect is the direct connection between the functional concept of generalized TV and the (cid:96)1-norm of the expansion coefficients a. We observe that the solution is made up of two components: an adaptive one that is specified by {xk} and a, and a linear regression term (with expansion coefficients b) that describes the component in the null space of the operator. Since b does not contribute to (cid:107)Ls(cid:107)M, the opti- mization tends to maximize the contribution of the null-space component. The main difficulty in finding the optimal solution is that K and (xk) are problem-dependent and unknown a priori. We have mentioned in Section 2.2 that the semi-norm (cid:107)Df(cid:107)M yields the classical total variation of a function in 1D. Unfortunately, there is no such direct connection for d > 1, the reason being that the multidimensional gradient ∇ is not spline-admissible because it is a vector operator. Instead, as a proxy for the popular total variation of Rudin and Osher [41], we suggest using the (fractional) Laplacian semi-norm (cid:107)(−∆)γ/2f(cid:107)M with γ ≥ d, which is endowed with the same invariance and null-space properties. According to Theorem 1, such a γth-order regularization results in extremal points that are nonuniform polyharmonic splines [21, 37]. 3.3 Connection with Unconstrained Problem The statement in Theorem 1 is remarkably general. In particular, it covers the generic regularized least-squares problem (cid:32) M(cid:88) m=1 fλ = arg min f∈ML(Rd) (cid:33) ym − (cid:104)νm, f(cid:105)2 + λ(cid:107)Lf(cid:107)M , (12) which is commonly used to formulate linear inverse/compressed-sensing problems [6, 9, 19, 23, 2 ≤ 2} = B(y; ), which is a 26]. The connection is obtained by taking C = {z ∈ RM : (cid:107)y − z(cid:107)2 ball of diameter  centered on the measurement vector y = (y1, . . . , yM ). Indeed, since the data- fidelity term is (strictly) convex, the extreme points s of (10) saturate the inequality such that (cid:107)y − ν(s)(cid:107)2 2 = 2 and gTV is minimized with α = α() = (cid:107)Ls(cid:107)M. In the unconstrained form (12), the selection of a fixed λ ∈ R+ results in a particular value of the data error (cid:107)y− ν(fλ)(cid:107)2 2 = (cid:48)(λ) with the optimal solution fλ = s(cid:48) having the same total variation as if we were looking at the primary problem (10) with C = B(y; (cid:48)). To get further insights on the optimization problem (12), we can look at two limit cases. When λ → ∞, the solution must take the form f∞ = p ∈ NL so that (cid:107)Lf∞(cid:107)M = 0. It then 2 ≤ (cid:107)y(cid:107)2 < ∞. On the contrary, when λ → 0, the minimization will follows that (cid:107)y − ν(f∞)(cid:107)2 force the data term (cid:107)y−ν(f0)(cid:107)2 2 to vanish. Theorem 1 then ensures the existence of a nonuniform "interpolating" L-spline f0(x) with ν(f0) = y and minimum gTV semi-norm. 8 3.4 Generalized Interpolation In the latter interpolation scenario, the convex set C reduces to the single point C = {y ∈ RM}. This configuration is of special theoretical relevance because it enables us to refine our upper bound on the number K of spline knots. Theorem 2 (Generalized spline interpolant). Under Assumptions 1-3 of Theorem 1, the extremal points of the (feasible) generalized interpolation problem (cid:107)Lf(cid:107)M s.t. arg min ν(f ) = y (13) f∈ML(Rd) are nonuniform L-splines of the same form (11) as in Theorem 1, but with K ≤ (M − N0). 4 Application Areas We first briefly comment on the admissibility conditions in Theorem 1 and indicate that the restrictions are minimal. To the best of our knowledge, the continuity of the measurement operator ν is a necessary requirement for the mathematical analysis of any inverse problem. The difficulty here is that our native space ML(Rd) = (CL(Rd))(cid:48) is non-reflexive, which forces us to rely on the weak*-topology. The continuity requirement in Theorem 1 is therefore equivalent to νm ∈ CL(Rd) for m = 1, . . . , M where the Banach structure of the predual space CL(Rd) is laid out in Theorem 6. In particular, we refer to the norm inequality (25), which suggests that Condition 2 in Theorem 1 is met by picking νm ∈ L1,−n0(Rd) where the latter is the Banach space associated with the weighted L1-norm (cid:90) Rd (cid:107)f(cid:107)L1,−n0 = f (x)(1 + (cid:107)x(cid:107))n0dx. (14) ML(Rd) = (cid:0)CL(Rd)(cid:1)(cid:48) ⊂ (cid:0)L1,−n0(Rd)(cid:1)(cid:48) In fact, L1,−n0(Rd) is the predual of the space L∞,n0(Rd) defined by (8), which implies that = L∞,n0(Rd). The condition νm ∈ L1,−n0(Rd) is a mild algebraic decay requirement that turns out to be satisfied by the impulse response of most physical devices. As for the requirement that the inverse problem is well defined over the null space of L (Condition 3), it a prerequisite to the success of any regularization scheme. Otherwise, there is simply no hope of turning an ill-posed problem into a well-posed one. For instance, in the introductory example with classical total-variation regularization, the constraint is that ν should have at least one component νm such that (cid:104)νm, 1(cid:105) (cid:54)= 0 which, again, is very mild requirement. Next, we discuss examples of signal recovery that are covered by Theorems 1 and 2. The standard setting is that one is given a set of noisy measurements y = ν(s) + "noise" of an unknown signal s and that one is trying to recover s from y based on the solution of (12), or some variant of the problem involving some other (convex) data term-the most favorable choice being the log likelihood of the measurement noise. We shall then close the discussion section by briefly making the connection with a class of inverse problems in measure space; that is, the case L = Identity. Interpolation 4.1 The task here is to reconstruct a continuous-domain signal from its (possibly noisy) nonuniform samples {s(xm)}M m=1, which is achieved by searching for the function s : Rd → R that fits the samples while minimizing (cid:107)Ls(cid:107)M. This corresponds to the problem setting in Theorem 1 with 9 νm = δ(· − xm) and C = B(y; ), where y denotes the measurement vector. Hence, the admis- −1 sibility condition νm ∈ CL(Rd) is equivalent to (L φ )∗{δ(· − xm)} = gφ(xm,·) ∈ C0(Rd), where the boundedness is ensured by the stability condition in Theorem 4. The more technical conti- nuity requirement is achieved when ρL is continuous (Hölder exponent r0 > 0). This happens when the order of the differential operator is greater than one, which seems to exclude2 simple operators such as D (piecewise-constant approximation). This limitation notwithstanding, our theoretical results are directly applicable to the problems of adaptive regression splines [38] with L = DN, the construction of shape-preserving splines [35], as well as a whole range of variations including TV denoising. 4.2 Generalized Sampling The setting is analogous to the previous one, except that the samples are now observed through a sampling aperture φ ∈ L1,−n0(Rd) so that νm = φ(· − xm) [24, 50]. The function φ may, for example, correspond to the point-spread function of a microscope. Then, the recovery problem is equivalent to a deconvolution [18]. Since the measurements are obtained by integration of s against an ordinary function νm ∈ L1,−n0(Rd), there is no requirement for the continuity of ρL because of the implicit smoothing effect of φ. This means that essentially no restrictions apply. 4.3 Compressed Sensing The result of Theorem 1 is highly relevant to compressed sensing, especially since the underlying L1/TV signal-recovery problem is formulated in the continuous domain. We like to view (11) as the prototypical form of a piecewise-smooth signal that is intrinsically sparse with sparsity K = (cid:107)a(cid:107)0. The model also conforms with the notion of a finite rate of innovation [56]. If we know that the unknown signal s has such a form, then Theorem 1 suggests that we can attempt to recover it from an M-dimensional linear measurement y = ν(s) by solving the optimization problem (10) with C = B(y; ), which is in agreement with the predominant paradigm in the field. While the theorem states that M ≥ K, common sense dictates that we should take M > Nfreedom, where Nfreedom = 2K + N0 is the number of degrees of freedom of the underlying model. The difficulty, of course, is that a subset of those parameters (the spline knots xk) induce a model dependency that is highly nonlinear. Inverse Problems in the Space of Measures 4.4 Some of the theoretical results of this paper are also of direct relevance for inverse problems that are formulated in the space M(Rd) of measures [5]. The prototypical example is the recovery of the location (with super-resolution precision) of a series of Dirac impulses from noisy measurements, which may be achieved through the continuous-domain minimization of the total variation of the underlying measure [8, 16, 22, 25]. The Fisher-Jerome theorem [27, Theorem 1] as well as our extension for the unbounded domain Rd and arbitrary convex sets (Theorem 7) support this kind of algorithm, as they guarantee the existence of sparse solutions-understood as a sum of Dirac spikes-for this family of problems. 2We can bypass this somewhat artificial limitation by replacing the ideal sampler by a quasi-ideal sampling device that involves a mollified version of a Dirac impulse. 10 5 Splines and Operators We now switch to the explanatory part of the presentation. The first important concept that is implicit in the statement of Theorems 1 and 2 is the powerful association between splines and operators, the idea being that the selection of an admissible operator L specifies a corresponding type of splines [46, 47][55, Chapter 6]. Sorted by increasing complexity, the three types of operators that are of relevance to us are: (i) ordinary differential operators, which are polynomials of the derivative operator D = d dx [13, 46, 52]; (ii) partial differential operators such as the Laplacian ∆ (or some polynomial 2 with γ ∈ R+ whose Fourier thereof); and (iii) fractional derivatives such as Dγ or (−∆) symbols are (jω)γ and (cid:107)ω(cid:107)γ, respectively [21, 51, 53]. It can be shown that all linear-shift- invariant operators of Type (i) and all elliptic operators of Type (ii) are spline-admissible in the sense of Definition 1. This is also known to be true for fractional derivatives and fractional Laplacians with γ ≥ d [21, 51]. Let us mention that the issue of making sure that the null space of the operator L is finite- dimensional is often nontrivial for d > 1. It is a fundamental aspect that is addressed in the L2 theory of radial basis functions and polyharmonic splines with the definition of the appropriate native spaces [58, Chapter 10]. Here, we have chosen to bypass some of these technicalities by including a growth restriction(cid:0)i.e., the condition that q ∈ L∞,n0(Rd)(cid:1) in the definition of NL. A to a zero of multiplicity at least m + 1 of the frequency response (cid:98)L(ω) at ω = ω0 (see [55, fundamental property in that respect is that the finite-dimensional null space of a LSI operator can only include exponential polynomial components of the form xmej(cid:104)ω0,x(cid:105), which correspond γ Proposition 6.1 p. 118] and [32, Section 6]). Once it is established that L is spline-admissible, one can rely on the following unifying distributional definition of a spline. Definition 2 (Nonuniform L-spline). A function s : Rd → R of slow growth (i.e., s ∈ L∞,n0(Rd) with n0 ≥ 0) is said to be a nonuniform L-spline if (cid:88) L{s} = akδ(· − xk) = wδ, (15) k where (ak) is a sequence of weights and the Dirac impulses are located at the spline knots {xk}. The generalized function L{s} = wδ is called the innovation of the spline because it contains the crucial information for its description: the positions {xk} of the knots and the amplitudes (ak) of the corresponding discontinuities. The one-dimensional brands of greatest practical interest are the polynomial splines with L = Dm [15, 45] and the exponential splines [13, 46, 52] with L = cmDm +··· + c1D + c0I, where I = D0 denotes the identity operator. Their multidimensional counterparts are the polyharmonic splines with L = (−∆)γ/2 [21, 37] and the Sobolev splines with L = (I − ∆)γ/2 for γ ≥ d [57]. The connection with the theory of Sobolev spaces is that the Green's functions of (−∆)γ/2 (resp., (I − ∆)γ/2) are the kernels of the Riesz (resp., Bessel) potentials [31]. For a constructive use of Definition 2, we also need to be able to re-synthesize the spline s from its innovation. In the case of our introductory example with L = D (see Figure 1), one simply integrates wδ, which yields (6) (up to the integration constant b1) owing to the property that D−1{δ(· − xk)}(x) =(cid:82) x−∞ δ(τ − xk)dτ = 1+(x − xk). In principle, the same inversion procedure is applicable for the generic operator L and amounts to substituting the δ distribution in (15) by the Green's function ρL. The only delicate part is the proper handling of the "integration 11 constants" (the part of the solution that lies in the null space of the operator), which is achieved through the specification of N0 linear boundary conditions of the form (cid:104)φn, s(cid:105) = 0. We now show that the underlying functionals φ = (φ1, . . . , φN0) can be incorporated in the −1 φ . Our construction requires that φ first specification of an appropriate right-inverse operator L be matched to a basis of NL such as to form a biorthogonal system. We note that this is always feasible as long as the φn are linearly independent with respect to NL. (An explicit construction is given in the proof of Theorem 2.) Definition 3. The pair (φ, p) = (φn, pn)N0 if {pn}N0 φn ∈ N (cid:48) canonical basis. n=1 is called a biorthogonal system for NL ⊂ ML(Rd) n=1 is a basis of NL and the vector of "boundary" functionals φ = (φ1, . . . , φN0) with L satisfy the biorthogonality condition φ(pn) = en where en is the nth element of the operator L stability by relying on Theorem 3 whose proof is given in Appendix A. (cid:80)N0 The interest of such a system is that any q ∈ NL has a unique representation as q = n=1(cid:104)φn, q(cid:105)pn with associated norm (cid:107)φ(q)(cid:107)2. The fundamental requirement for our formulation is the stability/continuity of the inverse −1 φ : M(Rd) → ML(Rd). Since ML(Rd) ⊂ L∞,n0(Rd) by construction, we can control Theorem 3. The generic linear operator G : w (cid:55)→ f = (cid:82) (cid:0)g(x, y) (1 + (cid:107)x(cid:107)) M(Rd) → L∞,α0(Rd) with α0 ∈ R if and only if its kernel g is measurable and Rd g(·, y)w(y)dy continuously maps −α0(cid:1) < ∞. (16) ess sup x,y∈Rd This allows us to characterize the desired operator in term of its Schwartz' kernel (or gener- n=1 be a biorthogonal system for NL ⊂ Rd gφ(·, y)ϕ(y)dy −1 −1 φ : ϕ (cid:55)→ L φ ϕ =(cid:82) alized impulse response) gφ(x, y) = L −1 φ {δ(· − y)}(x). Theorem 4 (Stable right-inverse of L). Let (φn, pn)N0 ML(Rd) ⊂ L∞,n0(Rd). Then, there exists a unique operator L such that LL −1 φ ϕ = ϕ −1 φ ϕ) = 0 φ(L (right-inverse property) (boundary condidions) for all ϕ ∈ S(Rd). The kernel of this operator is gφ(x, y) = ρL(x − y) − N0(cid:88) pn(x)qn(y), n=1 (17) (18) (19) with ρL such that LρL = δ and qn(y) = (cid:104)φn, ρL(· − y)(cid:105). Moreover, if gφ satisfies the stability −1 φ admits a continuous extension M(Rd) → L∞,n0(Rd) with condition (16) with α0 = n0, then L (17) and (18) remaining valid for all ϕ ∈ M(Rd). The proof of Theorem 4 is given in Appendix B. Since the choice of the N0 linear boundary functionals φn is essentially arbitrary, there is flexibility in defining admissible inverse operators. The important ingredient for our formulation is the existence of such inverses with the unconditional guarantee of their stability (see Theorem 5 below). To put this result into context, we now provide some illustrative examples. For L = DN0, (n−1)! for n = 1, . . . , N0, where we have that n0 = (N0 − 1), ρDN0 (x) = n0! , and pn(x) = xn−1 xn0 + 12 the polynomial basis is biorthogononal to φ with φn(x) = (−1)(n−1)δ(n−1)(x). This (canonical) −1 choice of boundary functionals then translates into the construction of an inverse operator L φ that imposes the vanishing of the function and its derivatives at the origin. By applying (19) and recognizing the binomial expansion of (x − y)n0, we simplify the expression of the kernel of this operator as (x − y)n0 − n0(cid:88) (cid:40) (x−y)n0 n0! + n0! − (x−y)n0 n0! + n=0 xn n! (−y)n0−n (n0 − n)! 1(0,x](y), x ≥ 0 1(x,0](y), x < 0. gφ(x, y) = = The crucial observation here is that the function gφ(x,·) with x ∈ R fixed is compactly supported and bounded. Moreover, (cid:107)gφ(x,·)(cid:107)∞ = gφ(x, 0) = xn0 n0! so that gφ obviously satisfies the stability bound (16) with α0 = n0. By contrast, the condition fails for the conventional shift-invariant inverse ϕ (cid:55)→ ρDN0 ∗ ϕ (n0-fold integrator), which stresses out the non-trivial stabilizing effect of the second correction term in (19). The other important consequence of the correction is the vanishing of gφ(x, y) as y → ±∞ for any fixed x ∈ Rd, contrary to its leading term (x− y)n0 + /n0! which does not decay (and even grows) as y → −∞. The primary usage of the inverse operators of Theorem 4 is the resolution of differential equations of the form for some w ∈ M(Rd). (φ, p), we readily show that (20) admits a unique solution in ML(Rd), which is given by Indeed, by invoking the properties of L φ(s) = (b1, . . . , bN0) −1 φ and the biorthogonality of Ls = w s.t. (20) s = L −1 φ w + bnpn. N0(cid:88) n=1 For the particular case of the spline innovation wδ in Definition 2, we find that (cid:88) N0(cid:88) s = −1 φ {δ(· − xk)} + akL bnpn which, upon substitution of the kernel given by (19), results in a form that is the same as (11) in Theorem 1 modulo some adjustment of the constants bn. k n=1 6 Native or Generalized Beppo-Levi Spaces The search for the solution of our optimization problem is performed over the native space ML(Rd) defined by (9), which is the largest space over which our gTG regularization functional is well defined. The delicate aspect is that ML(Rd) is specified in terms of a semi-norm, in analogy with the definition of the classical Beppo-Levi spaces of order n ∈ N and exponent p ≥ 1, written as Bp,n(Rd) = {f ∈ S(cid:48)(Rd) : (cid:107)∂mf(cid:107)Lp < ∞ for all multi-indices m = n} [17, 34]. Hence, in 1D, the proposed definition of MDn(R) is a slight extension of B1,n(R). In higher dimensions, it can be shown3 that Bp,2n(Rd) = {f ∈ L∞,2n−1(Rd) : (cid:107)(−∆)nf(cid:107)Lp < ∞}, so that there also exists a close connection between B1,2n(Rd) and M(−∆)n(Rd). 3The argument is that the only functions that are harmonic and of slow growth are polynomials. 13 The crucial point for our formulation is that ML(Rd) also happens to be a complete normed (or Banach) space when equipped with the proper direct-sum topology. We shall now make this structure explicit with the help of the inverse operators defined in Theorem 4. Since the principle is similar to the characterization of the Beppo-Levi spaces, we shall also refer to ML(Rd) as a generalized Beppo-Levi space. Theorem 5 (Banach-space structure of native space). Let L be a spline-admissible operator, ML(Rd) its native space defined by (9), and (φ, p) some biorthogonal system for its null space NL. Then, the following equivalent conditions hold: 1. The right-inverse operator L −1 φ specified by Theorem 4 isometrically maps M(Rd) → ML(Rd) ⊂ L∞,n0(Rd), while its kernel necessarily fullfills the stability condition (cid:0)gφ(x, y) (1 + (cid:107)x(cid:107)) −n0(cid:1) < ∞. Cφ = sup x,y∈Rd (21) (22) 2. Every f ∈ ML(Rd) admits a unique representation as where w = L{f} ∈ M(Rd) and p =(cid:80)N0 −1 φ w + p, f = L n=1(cid:104)f, φn(cid:105)pn ∈ NL. 3. ML(Rd) is a Banach space equipped with the norm (cid:107)f(cid:107)L,φ = (cid:107)Lf(cid:107)M + (cid:107)φ(f )(cid:107)2. Proof. As preparation, we define a subset of ML(Rd) as ML,φ(Rd) = {f ∈ ML(Rd) : φ(f ) = 0}. (23) Since the boundary conditions φ(f ) = 0 are linear, ML,φ(Rd) is clearly a vector space. We now show that it is a Banach space when equipped with the norm (cid:107) · (cid:107)L = (cid:107)L{·}(cid:107)M. By definition, (cid:107)·(cid:107)L is a semi-norm on ML(Rd), meaning that it fulfills the properties of a norm, except for the unicity condition. To establish the latter on ML,φ(Rd), we consider f ∈ ML,φ(Rd) such that (cid:107)f(cid:107)L = 0, which is equivalent to f ∈ NL. Since f ∈ ML,φ(Rd) (by hypothesis), we have that n=1(cid:104)φn, f(cid:105)pn = 0, as expected. This proves that ML,φ(Rd) is isometrically isomorphic to M(Rd) and, hence, a Banach space. Alternatively, one can also view ML,φ(Rd) as a concrete transcription (or representative within the equivalence class) of the abstract quotient space ML(Rd)/NL. φ(f ) = 0, from which we deduce that f =(cid:80)N0 1. Existence and Stability of Inverse Operators. We have just revealed that L is a bijective, norm-preserving mapping ML,φ(Rd) → M(Rd). This allows us to invoke the bounded-inverse theorem, which ensures the existence and boundedness (here, an isometry) of the inverse operator −1 L−1 : M(Rd) → ML,φ(Rd). The relevant L−1 is precisely the unique operator L φ identified in −1 φ w) = 0 for all w ∈ M(Rd). Finally, we Theorem 4, as it imposes the boundary condition φ(L use the same technique as in the proof of Theorem 3 to establish the necessity of the stability condition (16) with α0 = n0. 2. Direct Sum Decomposition. Since the system (φ, p) is biorthogonal, the operator ProjNL : n=1(cid:104)φn, f(cid:105)pn is a continuous projection operator ML(Rd) → NL(Rd). It follows that any f (cid:55)→(cid:80)N0 14 element f ∈ ML(Rd) has a unique decomposition as f = f1 + q, where q = ProjNL {f} ∈ NL and f1 = (f − q) with φ(f1) = 0. This last condition implies that f1 ∈ ML,φ(Rd) so that f1 has −1 φ w, where w = Lf1 = Lf ∈ M(Rd). Since ML,φ ∩NL = {0}, a unique representation as f1 = L this expresses the structural property that ML(Rd) = ML,φ(Rd) ⊕ NL. 3. Identification of the Underlying Norm. Any element p ∈ NL is uniquely characterized by {f} ∈ NL with its expansion coefficients φ(p) in the basis p. The same holds true for q = ProjNL φ(q) = φ(f ) for any f ∈ ML(Rd). Since ML,φ(Rd) and NL are both Banach spaces, we can equip their direct sum ML(Rd) with the composite norm (cid:107)f(cid:107)L,φ = (cid:107)w(cid:107)M + (cid:107)φ(f )(cid:107)2, with the guarantee that the Banach-space property is preserved. For the converse implication, we simply identify ML,φ(Rd) as the closed subspace of ML(Rd) with the property that (cid:107)f(cid:107)L,φ = (cid:107)Lf(cid:107)M. The connection with the L-spline s of Definition 2 is that s ∈ ML(Rd) if and only if the (cid:80)K (cid:96)1-norm of its spline weights a = (a1, . . . , aK) is finite. Indeed, we have that (cid:107)Ls(cid:107)M = (cid:107)wδ(cid:107)M = k=1 ak = (cid:107)a(cid:107)(cid:96)1, owing to the property that (cid:107)δ(· − xk)(cid:107)M = 1. We note that the choice of gTV is essential here since the simpler (and a priori only slightly more restrictive) L1-norm regularization (cid:107)Ls(cid:107)L1 would exclude the spline solutions that are of interest to us because δ /∈ L1(Rd). Our final ingredient is the identification of the predual space of ML(Rd), which is denoted and CL,p(Rd) be the image of C0(Rd) by L∗ : C0(Rd) → CL,p(Rd). Then, ML =(cid:0)CL(Rd)(cid:1)(cid:48) where by CL(Rd). Theorem 6 (Predual of native space). Let (φ, p) be a biorthogonal system of NL ⊂ L∞,n0(Rd) CL(Rd) = CL,p(Rd) ⊕ N (cid:48) norm n=1. CL(Rd) is a Banach space equipped with the L = span{φn}N0 L with N (cid:48) φ =(cid:0)L −1∗ −1 φ where L (cid:107)f(cid:107)(cid:48) (cid:1)∗ is the adjoint of L −1∗ φ f(cid:107)∞ + (cid:107)p(f )(cid:107)2. L,φ = (cid:107)L −1 φ . Moreover, there exists a constant C > 0 such that L,φ ≤ C(cid:107)f(cid:107)L1,−n0 (25) (cid:107)f(cid:107)(cid:48) (24) for any f ∈ L1,−n0(Rd). : f (cid:55)→ q = (cid:80)N0 The proof is given in Appendix C. The direct-sum decomposition in Theorem 6 is achieved n=1(cid:104)pn, f(cid:105)φn with (cid:107)q(cid:107) = (cid:107)p(f )(cid:107)2 = (cid:107)p(q)(cid:107)2, by means of the operator ProjN (cid:48) which relies on the biorthogonality of (φ, p) to project CL(Rd) onto N (cid:48) L(Rd). This also means that CL,p(Rd) can be defined as CL,p(Rd) = {f ∈ CL(Rd) : p(f ) = 0}, in direct analogy with the definition of ML,φ(Rd) in (23). L 7 Proof of Theorems 1 and 2 Our technique of proof will be to first establish the optimality of innovation-type solutions of the form that appear in Definition 2 for general linear inverse problems defined on M(Rd) (Theorem 7) and to then transfer the result to ML(Rd) with the help of the stable inverse operators specified in Theorem 4. The first step is achieved by generalizing an earlier result by Fisher and Jerome [27]. with some norm (cid:107) · (cid:107)N . The generic element of H is f = (w, p) with (cid:107)f(cid:107)H = (cid:107)w(cid:107)M + (cid:107)p(cid:107)N . Let H be the direct sum of M(Rd) =(cid:0)C0(Rd)(cid:1)(cid:48) and a finite-dimensional space N equipped 15 Theorem 7 (Generalized Fisher-Jerome theorem). Let F : H → RM with M ≥ N0 = dim(N ) be a weak*-continuous linear map such that (26) for some constant B > 0 and every p ∈ N . Let C be a convex compact subset of RM such that U = F −1(C) = {(w, p) ∈ H : F (w, p) ∈ C} is nonempty (feasibility hypothesis). Then, B(cid:107)p(cid:107)N ≤(cid:107)F (0, p)(cid:107)2 V = arg min (w,p)∈U (cid:107)w(cid:107)M is a nonempty, convex, weak*-compact subset of H with extremal points (wδ, p) of the form K(cid:88) wδ = akδ(· − xk) with K ≤ M and xk ∈ Rd for k = 1, . . . , K, and min(w,p)∈U (cid:107)w(cid:107)M =(cid:80)K k=1 k=1 ak. (27) Theorem 7 is the most technical component of our formulation as it involves the weak* topology. The details of the proof are laid out in Appendix D together with a precise definition of the underlying concepts. The essence of Theorem 7 is very similar to Fisher-Jerome's original result [27, Theorem 1], except for two crucial points: (i) the fact that they are only considering measures defined over a bounded domain Ω ⊂ Rd (or, by extension, on a compact metric space), and (ii) the nature of the constraints which, in their case, is limited to coordinatewise inequalities of the form z1,m ≤ [F (w, p)]m ≤ z2,m. These differences are substantial enough to justify a new, self-contained proof. In particular, we believe that our extension for functions defined on Rd (beyond the compact Hausdorff framework of [27]) is essential for covering nonlocal operators such as fractional derivatives, and for deploying Fourier-domain/signal-processing techniques. Our primary constraint for the validity of Theorem 7 is the existence of the lower bound (26). We now show that this property is implicit in the statement of the hypotheses of Theorem 1. (cid:80)N0 n=1 be a biorthogonal system of NL ⊂ ML(Rd) such that q = n=1(cid:104)φn, q(cid:105)pn for all q ∈ NL. Then, Condition 3 in Theorem 1 is equivalent to the existence of Proposition 1. Let (φn, pn)N0 a constant 0 < B such that with (cid:107)q(cid:107)2NL = (cid:107)φ(q)(cid:107)2 n=1 (cid:104)φn, q(cid:105)2. 2 =(cid:80)N0 B(cid:107)q(cid:107)NL ≤ (cid:107)ν(q)(cid:107)2, ∀q ∈ NL (28) While there are softer ways of establishing this equivalence, we have chosen an explicit ap- proach that also serves as background for the proof of Theorem 2. Proof. Any q ∈ NL has a unique expansion q =(cid:80)N0 n=1 cnpn with c = φ(q) and (cid:107)q(cid:107)NL = (cid:107)c(cid:107)2. The property that q is uniquely determined by its measurements b = ν(q) is therefore equivalent to c also being the solution of the overdetermined system Pc = b with P = [ν(p1) ··· ν(pN0)]. (29) 16 It is well known that such a system is solvable if and only if (PT P) is invertible and that its (least-squares) solution is given by This characterization then yields the norm estimate c = (PT P) −1PT b. (cid:107)q(cid:107)NL = (cid:107)c(cid:107)2 ≤ σmax(P) σ2 min(P) (cid:107)ν(q)(cid:107)2, min(P)/σmax(P). where σmin(P) = σmin(PT ) and σmax(P) are the minimum and maximum singular values of P, respectively. Finally, the invertibility of (PT P) is equivalent to σ2 min(P) = λmin(PT P) > 0, while the continuity assumption on ν ensures that σmax(P) < ∞. The constant is then given by B = σ2 Proof of Theorem 1. Let (φ, p) be a biorthogonal system for NL. Then, by Theorem 5, any −1 function f ∈ ML(Rd) has a unique decomposition as f = L φ w + p with w = Lf ∈ M(Rd) and p ∈ NL. This allows us to interpret the measurement process f (cid:55)→ ν(f ) = (cid:104)ν, f(cid:105) as the linear map F : H → RM such that −1 φ w(cid:105) + (cid:104)ν, p(cid:105) (cid:104)ν, f(cid:105) = (cid:104)ν, L −1∗ φ ν, w(cid:105) + (cid:104)ν, p(cid:105) = F (w, p). = (cid:104)L −1∗ is an isometry CL(Rd) → C0(Rd). Hence, the weak*- We also know from Theorem 6 that L φ continuity of ν : ML(Rd) → RM is equivalent to the weak*-continuity of F : H → RM. The complementary lower bound is given by Proposition 1 as B(cid:107)p(cid:107)ML,φ ≤(cid:107)ν(p)(cid:107)2 = (cid:107)F (0, p)(cid:107)2. are of the form (p, wδ) with wδ =(cid:80)K With this new representation, the constrained minimization problem is equivalent to the one considered in Theorem 7 with N = NL, which ensures that all extreme points of the solution set k=1 akδ(· − xk), K ≤ M, and xk ∈ Rd. Upon application of −1 φ wδ +p, where p is a suitable component the (stable) right-inverse operator, this maps into s = L that is in the null space of the operator. Finally, we use the explicit kernel formula (19) and the procedure outlined at the end of Section 5 to convert this representation into (11), which removes the artificial dependence upon φ. Proof of Theorem 2. From the proof of Proposition 1, we know that the minimal singular value of the cross-product matrix P = [p1 ··· pM ]T with pm = ν(pm) ∈ RN0 is non-vanishing. The geometric implication is that span{p}M m=1 = RN0. Since the corresponding system is redundant, we can always identify a subset of these row vectors that forms a basis of RN0. Without loss of generality, we now assume that this subset is {pm}N0 m=1 and that the corresponding subma- trix P0 = [p1 ··· pN0]T is therefore invertible. In other words, we have identified a reduced vector of measurement functionals ν0 = (ν1, . . . , νN0) that is linearly independent with respect to NL. This, in turn, allows us to construct the boundary functional φ0 = P−1 0 ν0 that meets the biorthogonal requirement φ0(pn) = P−1 In effect, this yields a biorthogonal system with the property that N (cid:48) Coming back to our interpolation problem, we define y = (y0, y1) with y0 = (y1, . . . , yN0) ∈ RN0 and y1 ∈ RM−N0. Due to the biorthogonality of (φ0, p), the unique element q0 ∈ NL such 0 ν0(pn) = P−1 L = span{νn}N0 0 pn = en. n=1 ⊂ M(cid:48) L(Rd). 17 that ν0(q0) = P0φ0(q0) = y0 is given by N0(cid:88) n=1 bnpn q0 = with b = (b1, . . . , bN0) = P−1 0 y0. The other ingredient is Theorem 5, which ensures that any f ∈ ML(Rd) has a unique decomposition as f = L w + q with q ∈ NL and w ∈ M(Rd). Now, the crucial property is that the boundary conditions φ0(L w) = 0 for all w ∈ ML(Rd). This allows us rewrite the solution of our generalized interpolation problem as f = L w) = 0 imply that ν0(L w1 + q0, where −1 φ0 −1 φ0 −1 φ0 −1 φ0 w1 = arg min w∈M(Rd) (cid:107)w(cid:107)M s.t. ν1(L −1 φ0 w) = y1 − ν1(q0). The result then follows from the continuity of L N = {0}. −1 φ0 and the reduced version of Theorem 7 with 8 Further Theoretical and Computational Issues The analogy with the finite-dimensional theory of compressed sensing raises the fundamental theoretical question: Is it possible to provide conditions on the measurement operator ν such that a perfect recovery is possible for certain classes of signals; in particular, splines with a given number of knots? This is an open topic that calls for further investigation. Because the problem is formulated in the continuum, we suspect that it is much more difficult-if not impossible-to identify conditions that ensure unicity. While the reconstruction problem in Theorem 1 is formulated in analysis form (i.e., minimiza- tion of (cid:107)Ls(cid:107)M), the interesting outcome is that the solution (11) is given in synthesis form, with the unusual twist that the underlying dictionary {ρL(· − τ )}τ∈Rd of basis functions is infinite- dimensional and not even countable. This interpretation suggests a natural discretization which n=1 with N (cid:29) M and to rely is to select a finite subset of equally-spaced functions {ρL(· − τn)}N on linear programming for  = 0, or quadratic programming for  > 0, or some other convex optimization technique to numerically solve the underlying (cid:96)1-minimization problem. We have preliminary evidence that this approach is feasible. In particular, we have considered the gen- eralized interpolation scenario covered by Theorem 2 and observed that the simplex algorithm performs well in the sense that it always returns a nonuniform L-spline with a number of knots K ≤ (M − N0). The key theoretical question now is to establish the convergence of such a scheme as the sampling step gets smaller. Since the space that is spanned by the null-space components of L and the integer shifts of ρL is the space of cardinal L-splines (see [52] for the generic case of an ordinary differential operator), one may also consider an alternative discretization that uses the corresponding B-spline basis functions, which are much better conditioned than Green's functions. This would bring us back to a numerical problem that is very similar to (3), with the advantage of maintaining a direct control over the discretization error. In the case of a pure denoising problem, another possible option is to run the taut-string algorithm [38, 44] or some appropriate variation thereof. At any rate, we believe that the issue of the proper discretization of the reconstruction problem (12) as well as the development of adequate numerical schemes are important research topics on their own right. For the cases where the solution is not unique, Theorem 1 also suggests a new computational challenge: the design of a minimization algorithm that systematically 18 converges to an extremal point of the problem, the best solution being the spline that exhibits the minimal number of knots K = (cid:107)a(cid:107)0. A Proof of Theorem 3 f (x) = G{w}(x) =(cid:82) Proof. First, we establish the sufficiency of the stability condition by considering the signal Rd g(x, y)w(y)dy, where w ∈ M(Rd), and by constructing the estimate (cid:12)(cid:12)(cid:12)(cid:12)(cid:90) (cid:12)(cid:12)(cid:12)(cid:12) (cid:33) f (x)(1 + (cid:107)x(cid:107)) −α0 = (1 + (cid:107)x(cid:107)) ≤ (1 + (cid:107)x(cid:107)) Rd −α0 −α0 ess sup y∈Rd g(x, y)w(y)dy g(x, y) (cid:107)w(cid:107)M, which implies that (cid:107)f(cid:107)∞,α0 = (cid:107)G{w}(cid:107)∞,α0 ≤ (cid:32) g(x, y) (1 + (cid:107)x(cid:107)) −α0 ess sup x,y∈Rd (cid:107)w(cid:107)M for all w ∈ M(Rd). In doing so, we have shown that (cid:107)G(cid:107) ≤ ess sup x,y∈Rd g(x, y) (1 + (cid:107)x(cid:107)) −α0 < ∞. To prove necessity, we use the property that g(x, y) = G{δ(· − y)}(x), where the shifted Dirac impulse δ(· − y) is included in M(Rd) with (cid:107)δ(· − y)(cid:107)M = 1. We then observe that, for each y ∈ Rd, (cid:107)G{δ(· − y)}(cid:107)∞,α0 = ess sup x∈Rd (1 + (cid:107)x(cid:107)) −α0g(x, y). Moreover, G being bounded, we have that (cid:107)G{δ(· − y)}(cid:107)∞,α0 ≤ (cid:107)δ(· − y)(cid:107)M (cid:107)G(cid:107) = (cid:107)G(cid:107), which means that ess sup x∈Rd (1 + (cid:107)x(cid:107)) −α0g(x, y) ≤ ess sup x,y∈Rd (1 + (cid:107)x(cid:107)) −α0g(x, y) ≤ (cid:107)G(cid:107) < ∞. As we already know that the inequality holds in the other direction as well, we obtain (cid:107)G(cid:107) = ess sup x,y∈Rd g(x, y) (1 + (cid:107)x(cid:107)) −α0, which concludes the proof. B Proof of Theorem 4 Proof. We first establish the properties of the operator on Schwartz' space of smooth and rapidly- decreasing signals S(Rd) to avoid any technical problem related to splitting sums and interchang- ing integrals. We also rely on Schwartz's kernel theorem which states the equivalence between the continuous linear operators G : S(Rd) → S(cid:48)(Rd) and their Schwartz kernels (or general im- pulse response) g ∈ S(cid:48)(Rd × Rd), meaning that two such operators are identical if and only if their kernels are equal-in the sense of distributions. 19 Using the explicit representation of gφ together with L{ρL} = δ (Dirac distribution) and L{pm} = 0 for m = 1, . . . , N0, we then easily show that −1 φ {ϕ} = ϕ LL for all ϕ ∈ S(Rd). Next, we invoke the biorthogonality property (cid:104)φm, pn(cid:105) = δm−n (Kronecker delta) to evaluate the inner product of (19) with φm as φ {ϕ}(cid:105) = (cid:104)φm, ρL ∗ ϕ(cid:105) − N0(cid:88) −1 (cid:104)φm, L (cid:104)φm, pn(cid:105)(cid:104)qn, ϕ(cid:105) n=1 = (cid:104)φm, ρL ∗ ϕ(cid:105) − (cid:104)qm, ϕ(cid:105) = (cid:104)φm, ρL ∗ ϕ(cid:105) − (cid:104)φm, ρL ∗ ϕ(cid:105) = 0, which proves that the boundary conditions are satisfied. Let us now consider another operator L−1 that is also a right inverse of L. Clearly, the result of the action of the two operators can only differ by a component that is in the null space of −1 φ ϕ) = q ∈ NL. Since (φ, p) forms a biorthogonal system, q is uniquely L so that (L−1ϕ − L determined by φ(q), which implies that the right-inverse operator that imposes the condition φ(L−1ϕ) = 0 is unique. Next, we define C = supx,y∈Rd (gφ(x, y) (1 + (cid:107)x(cid:107))−n0) < ∞ and extract the generic conti- nuity bound (cid:107)L −1 φ ϕ(cid:107)∞,n0 ≤ C(cid:107)ϕ(cid:107)M from the the proof of Theorem 3. This allows us to extend the domain of the operator from S(Rd) to M(Rd), by the Hahn-Banach theorem. Since S(Rd) is dense in M(Rd), we can do likewise for the right-inverse property and the boundary conditions by invoking the continuity of L space ML,φ(Rd) and then by showing that it is the bijective image of M(Rd) by L of Statement 1 in Theorem 5, which also nicely settles the issue of stability). −1 φ and φ. Alternatively, one can establish this extension indirectly by identifying a specific Banach −1 φ (see proof C Proof of Theorem 6 Proof. First, we prove that CL,p(Rd) is isometrically isomorphic to C0(Rd). For any ϕ ∈ C0(Rd), L∗ϕ = 0 implies that ϕ ∈ NL∗ ∩ C0(Rd) = {0} (since the basis functions of the null space do not vanish at infinity); i.e., ϕ = 0. L∗ is therefore injective, and hence bijective since it is −1∗ surjective C0(Rd) → CL,p(Rd) by definition. In particular, this implies that the adjoint L φ −1 φ defined by (19) is the inverse of L∗ from CL,p(Rd) to C0(Rd). Therefore, of the operator L CL,p(Rd) inherits the Banach-space structure of C0(Rd) for the norm (cid:107)L If f ∈ CL,p(Rd), then f = L∗ϕ with ϕ ∈ C0(Rd) and (cid:104)f, p(cid:105) = (cid:104)ϕ, Lp(cid:105) = 0 for any p ∈ NL. The unique element of N (cid:48) L is a (finite-dimensional) Banach space for the norm (cid:107)p(f )(cid:107)2, implying that CL(Rd) = CL,p(Rd)⊕N (cid:48) is a Banach space for (24). −1 φ is continuous and bijective from ML,φ(Rd) to M(Rd) (Theorem 5), Next, we recall that L while we have just shown that its adjoint is continuous and bijective from CL,p(Rd) to C0(Rd). = ML,φ(Rd). Finally, we have −1 φ f(cid:107)∞. L orthogonal to NL is 0 so that the sum CL,p(Rd)⊕N (cid:48) = M(Rd), this implies that(cid:0)CL,p(Rd)(cid:1)(cid:48) Knowing that(cid:0)C0(Rd)(cid:1)(cid:48) L is direct. N (cid:48) L (cid:48) (CL(Rd)) = (CL,p(Rd) ⊕ N (cid:48) (cid:48) L) = (CL,p(Rd)) (cid:48) ⊕ (N (cid:48) (cid:48) L) = ML,φ(Rd) ⊕ NL = ML(Rd), 20 as expected. L1,−n0(Rd) and pn ∈ L∞,n0(Rd) imply that (cid:104)f, pn(cid:105) ≤ (cid:107)pn(cid:107)∞,n0 (cid:107)f(cid:107)L1,−n0 equality). Likewise, using the stability bound (21), we get To establish the weighted L1-norm inequality, we first observe that the hypotheses f ∈ (by the Hölder in- (cid:12)(cid:12)(cid:12)(cid:12)(cid:90) (cid:90) Rd (cid:12)(cid:12)(cid:12)(cid:12) −1∗ φ {f}(y) = L ≤ gφ(x, y) f (x)dx Rd Cφ(1 + (cid:107)x(cid:107))n0f (x)dx = Cφ(cid:107)f(cid:107)L1,−n1, which yields (cid:107)L tion of these individual bounds. −1∗ φ {f}(cid:107)∞ ≤ Cφ(cid:107)f(cid:107)L1,−n1. The desired result is then obtained from the summa- D Proof of Theorem 7 As preparation, we recall that the weak*-topology on M(Rd) = (cid:0)C0(Rd)(cid:1)(cid:48) is the locally The proof follows the same steps as the original one of Fisher and Jerome [27, Theorem 1]. Yet, it differs in the assumptions and technicalities (i.e., the consideration of the non-compact domain Rd and the use of explicit bounds). We have done our best to make it self-contained. convex topology associated with the family of semi-norms pϕ(w) = (cid:104)w, ϕ(cid:105) for ϕ ∈ C0(Rd). In particular, a sequence of elements wn ∈ M(Rd) converges to 0 for the weak*-topology if and only if (cid:104)wn, ϕ(cid:105) → 0 for every ϕ ∈ C0(Rd). A subset of M(Rd) is said to be weak*-closed (weak*-compact, respectively) if it is closed (compact, respectively) for the weak*-topology. We shall use Propositions 2 and 3, which are consequences of the Banach-Alaoglu theorem and its variations [43, p.68]. Proposition 2. Compactness in the weak*-topology of M(Rd). • For every α > 0, the set Bα = {w ∈ M(Rd) : (cid:107)w(cid:107)M ≤ α} is weak*-compact in M(Rd). • If (wn) is a sequence in M(Rd), bounded for the TV-norm, then we can extract a subse- quence that converges in M(Rd) for the weak*-topology. properties also carry over to the Banach space H = M(Rd) ⊕ N = (cid:0)C0(Rd) ⊕ N (cid:48)(cid:1)(cid:48), which is The second point of Proposition 2 is valid because the space C0(Rd) is separable. These endowed with the corresponding weak*-topology: A sequence (wn, pn) in H vanishes for the weak*-topology if and only if wn vanishes for the weak*-topology of M(Rd) and (cid:107)pn(cid:107)N → 0. Proposition 3. Compactness in the weak*-topology of H. • For every α1, α2 > 0, the set Bα1,α2 = {(w, p) ∈ H : (cid:107)w(cid:107)M ≤ α1, (cid:107)p(cid:107)N ≤ α2} is weak*-compact in H. • If (wn, pn) is a sequence in H such that (cid:107)wn(cid:107)M + (cid:107)pn(cid:107)N is bounded, then we can extract a subsequence that converges in H for the weak*-topology. Proof of Theorem 7. The proof is divided in two parts. First, we show that V is a nonempty, convex, and weak*-compact subspace of H. This allows us to specify V by means of its extremal points via the Krein-Milman theorem. Second, we show that the extremal points have the announced form. We set β = inf (w,p)∈U(cid:107)w(cid:107)M. 21 Part I: V is nonempty, convex, and weak*-compact Since F is weak*-continuous, it is also continuous H → RM in the topology of H. Hence, there exists a constant A > 0 such that (cid:107)F (w, p)(cid:107)2 ≤ A((cid:107)w(cid:107)M + (cid:107)p(cid:107)NL). (30) Let us consider a sequence (wn, pn)n∈N in U such that (cid:107)wn(cid:107)M decreases to β. In particular, (cid:107)wn(cid:107)M is bounded above by (cid:107)w0(cid:107)M. We set M = maxx∈C(cid:107)x(cid:107). Using respectively (26), (30), and (cid:107)F (wn, pn)(cid:107)2 ≤ M (since (wn, pn) ∈ U), we deduce the inequalities (cid:107)F (wn, pn) − F (wn, 0)(cid:107)2 (cid:107)pn(cid:107)N ≤ 1 B ≤ 1 B ≤ 1 B 1 B (cid:107)F (0, pn)(cid:107)2 = ((cid:107)F (wn, pn)(cid:107)2 + (cid:107)F (wn, 0)(cid:107)2) (M + A(cid:107)wn(cid:107)M) ≤ 1 B (M + A(cid:107)w0(cid:107)M), (31) which shows that pn is bounded. We can then extract a sequence (wsn, psn) from (wn, pn) that converges to (w∞, p∞) ∈ H for the weak*-topology (by Proposition 3). Since (cid:107)wn(cid:107) → β and (cid:107)wsn(cid:107) is a subsequence, it must also converge to β. On the other hand, the set U = F −1(C) is weak*-closed in H, as the preimage of a closed set by a weak*-continuous function F . Consequently, (w∞, p∞) is the weak*-limit of a sequence of elements in U. We therefore deduce that (w∞, p∞) ∈ U, so that (cid:107)w∞(cid:107)M ≥ β. In light of the previous inequality, this yields (cid:107)w∞(cid:107)M = β, which proves that V is not empty. F linear. Likewise, V = U(cid:84){(w, p) : (cid:107)w(cid:107)M ≤ β} is convex, weak*-closed as the intersection In addition to being weak*-closed, the set U = F −1(C) is convex because C is convex and of two sets with the same property. Finally, for (w, p) ∈ V, we show that (cid:107)p(cid:107)N ≤ M +A(cid:107)w(cid:107)M B = M +Aβ B = γ, based on the same inequalities as in (31). Therefore, we have V ⊂ {(w, p) ∈ H : (cid:107)w(cid:107)M ≤ β, (cid:107)p(cid:107)N ≤ γ}, where the set on the right-hand side is weak*-compact, due to Proposition 3. Since any weak*- closed set included in a weak*-compact set is necessarily weak*-compact, this shows that V is weak*-compact. We are now in the position to apply the Krein-Milman theorem [43, p. 75] to the convex weak*-compact set V ⊂ H, which tells us that "V is the closed convex hull of its extreme points in H endowed with the weak*-topology". This leads us to the final part of the proof, which is the characterization of those extreme points. Part II: The extreme points of V are of the form (27) We shall prove that a neces- sary condition for (w, p) to be an extreme point of V is that there are no disjoint Borelian sets E1, . . . , EM +1 ⊂ Rd such that w(Em) (cid:54)= 0 for m = 1, . . . , M + 1. The only elements of M(Rd) satisfying this condition are precisely those described by (27). We shall proceed by contradiction and assume that there exist disjoint sets E1, . . . , EM +1 We denote the restriction of w to Em as wm = w1Em. We also define E = Rd\(cid:83) such that w(Em) (cid:54)= 0 for all m. m Em, and ¯w = w1E with (cid:107)w(cid:107)M = β. For m = 1, . . . , M + 1, we set ym = F (wm, p) ∈ RM. Since any 22 that(cid:80)M +1 Let µ = (cid:80)M +1 m=1 cmym = 0. collection of (M + 1) vectors in RM is linearly dependent, there exists (cm)1≤m≤M +1 (cid:54)= 0 such m=1 cmwm ∈ M(Rd) and  ∈ (−max, max) with max = 1/ maxm cm, so that (1 + cm) > 0 and (1 − cm) > 0 for all m. By construction, we have that M +1(cid:88) M +1(cid:88) F (µ, p) = cmF (wm, p) = cmym = 0 (32) and, therefore, that F (w ± µ, p) = F (w, p). Hence, m=1 m=1 Moreover, w ± µ = ¯w +(cid:80)M +1 supports and (1 ± cm) > 0, we have (w ± µ, p) ∈ U. m=1 (1 ± cm)wm. Since the measures ¯w, w1, . . . , wM +1 have disjoint (cid:107)w ± µ(cid:107)M = (cid:107) ¯w(cid:107)M + = (cid:107) ¯w(cid:107)M + (1 ± cm)(cid:107)wm(cid:107)M M +1(cid:88) (cid:107)wm(cid:107)M ±  cm(cid:107)wm(cid:107)M = (cid:107)w(cid:107)M ±  m=1 cm(cid:107)wm(cid:107)M m=1 M +1(cid:88) M +1(cid:88) M +1(cid:88) m=1 = β ±  m=1 cm(cid:107)wm(cid:107)M. (33) If (cid:80)M +1 m=1 cm(cid:107)wm(cid:107)M (cid:54)= 0, then we either have (cid:107)w + µ(cid:107)M < β or (cid:107)w − µ(cid:107)M < β, which is impossible since the minimum over U is β. Hence, M +1(cid:88) m=1 M +1(cid:88) cm(cid:107)wm(cid:107)M = 0 and (cid:107)w + µ(cid:107)M = (cid:107)w − µ(cid:107)M = β, which translates into (w + µ, p) and (w − µ, p) being 2 (w − µ, p) is not an extreme included in V. This, in turn, implies that (w, p) = 1 2 (w + µ, p) + 1 point of V. . m=1 Acknowledgments The research was partially supported by the Swiss National Science Foundation under Grant 200020-162343, the Center for Biomedical Imaging (CIBM) of the Geneva-Lausanne Universities and EPFL, and the European Research Council under Grant 692726 (H2020-ERC Project Glob- alBioIm). The authors are thankful to H. Gupta for helpful discussions. They are also grateful to the three anonymous reviewers for their thoughtful suggestions which have helped improve the manuscript. 23 References [1] B. Adcock and A. Hansen, Generalized sampling and infinite-dimensional compressed sensing, Foundations of Computional Mathematics, (2015), pp. 1–61. [2] B. Adcock, A. C. Hansen, C. Poon, and B. Roman, Breaking the coherence barrier: A new theory for compressed sensing, Forum of Mathematics, Sigma, (to appear). [3] A. Beck and M. Teboulle, Fast gradient-based algorithms for constrained total varia- tion image denoising and deblurring problems, IEEE Transactions on Image Processing, 18 (2009), pp. 2419–2434. [4] A. Beck and M. Teboulle, A fast iterative shrinkage-thresholding algorithm for linear inverse problems, SIAM Journal on Imaging Sciences, 2 (2009), pp. 183–202. [5] K. Bredies and H. Pikkarainen, Inverse problems in spaces of measures, ESAIM: Con- trol, Optimisation and Calculus of Variations, 19 (2013), pp. 190–218. [6] A. M. Bruckstein, D. L. Donoho, and M. Elad, From sparse solutions of systems of equations to sparse modeling of signals and images, SIAM Review, 51 (2009), pp. 34–81. [7] E. J. Candès, The restricted isometry property and its implications for compressed sensing, Comptes Rendus de l'Académie des Sciences, 346 (2008), pp. 589–592. [8] E. J. Candès and C. Fernandez-Granda, Super-resolution from noisy data, Journal of Fourier Analysis and Applications, 19 (2013), pp. 1229–1254. [9] E. J. Candès and J. Romberg, Sparsity and incoherence in compressive sampling, Inverse Problems, 23 (2007), pp. 969–985. [10] E. J. Candès, J. K. Romberg, and T. Tao, Stable signal recovery from incomplete and inaccurate measurements, Communications on Pure and Applied Mathematics, 59 (2006), pp. 1207–1223. [11] A. Chambolle, An algorithm for total variation minimization and applications, Journal of Mathematical Imaging and Vision, 20 (2004), pp. 89–97. [12] A. Cohen, W. Dahmen, I. Daubechies, and R. DeVore, Harmonic analysis of the space BV, Revista Matematica Iberoamericana, 19 (2003), pp. 235–263. [13] W. Dahmen and C. Micchelli, On theory and application of exponential splines, in Top- ics in Multivariate Approximation, C. Chui, L. Schumaker, and F. Utreras, eds., Academic Press, New York, 1987, pp. 37–46. [14] I. Daubechies, M. Defrise, and C. De Mol, An iterative thresholding algorithm for linear inverse problems with a sparsity constraint, Communications on Pure and Applied Mathematics, 57 (2004), pp. 1413–1457. [15] C. de Boor, A Practical Guide to Splines, Springer-Verlag, New York, 1978. [16] Q. Denoyelle, V. Duval, and G. Peyré, Support recovery for sparse deconvolution of positive measures, arXiv preprint arXiv:1506.08264, (2015). 24 [17] J. Deny and J.-L. Lions, Les espaces du type de Beppo Levi, Annales de l'Institut Fourier, 5 (1954), pp. 305–370. [18] N. Dey, L. Blanc-Féraud, C. Zimmer, P. Roux, Z. Kam, J.-C. Olivo-Marin, and J. Zerubia, Richardson-Lucy algorithm with total variation regularization for 3D confocal microscope deconvolution, Microscopy Research and Technique, 69 (2006), pp. 260–266. [19] D. L. Donoho, Compressed sensing, IEEE Transactions on Information Theory, 52 (2006), pp. 1289–1306. [20] D. L. Donoho and M. Elad, Optimally sparse representation in general (nonorthogonal) dictionaries via (cid:96)1 minimization, Proceedings of the National Academy of Sciences, 100 (2003), pp. 2197–2202. [21] J. Duchon, Splines minimizing rotation-invariant semi-norms in Sobolev spaces, in Con- structive Theory of Functions of Several Variables, W. Schempp and K. Zeller, eds., Springer- Verlag, Berlin, 1977, pp. 85–100. [22] V. Duval and G. Peyré, Exact support recovery for sparse spikes deconvolution, Foun- dations of Computational Mathematics, 15 (2015), pp. 1315–1355. [23] M. Elad, Sparse and Redundant Representations. From Theory to Applications in Signal and Image Processing, Springer, 2010. [24] Y. Eldar, Sampling Theory: Beyond Bandlimited Systems, Cambridge University Press, 2015. [25] C. Fernandez-Granda, Super-resolution of point sources via convex programming, Infor- mation and Inference, 5 (2016), pp. 251–303. [26] M. A. T. Figueiredo and R. D. Nowak, An EM algorithm for wavelet-based image restoration, IEEE Transactions on Image Processing, 12 (2003), pp. 906–916. [27] S. Fisher and J. Jerome, Spline solutions to L1 extremal problems in one and several variables, Journal of Approximation Theory, 13 (1975), pp. 73–83. [28] S. Foucart and H. Rauhut, A Mathematical Introduction to Compressive Sensing, Springer, 2013. [29] I. M. Gelfand and G. Shilov, Generalized Functions. Vol. 1. Properties and Operations, Academic Press, New York, USA, 1964. [30] T. Goldstein and S. Osher, The split Bregman method for L1-regularized problems, SIAM Journal on Imaging Sciences, 2 (2009), pp. 323–343. [31] L. Grafakos, Classical and Modern Fourier Analysis, Prentice Hall, 2004. [32] Y. Hel-Or and P. C. Teo, Canonical decomposition of steerable functions, Journal of Mathematical Imaging and Vision, 9 (1998), pp. 83–95. [33] L. Hörmander, The Analysis of Linear Partial Differential Operators. I. Distribution The- ory and Fourier Analysis, vol. 256, Springer-Verlag, Berlin, 2nd ed., 1990. 25 [34] T. Kurokawa, Riesz potentials, higher Riesz transforms and Beppo Levi spaces, Hiroshima Math. J, 18 (1988), pp. 541–597. [35] J. E. Lavery, Shape-preserving, multiscale fitting of univariate data by cubic L1 smoothing splines, Computer Aided Geometric Design, 17 (2000), pp. 715–727. [36] M. Lustig, D. L. Donoho, and J. M. Pauly, Sparse MRI: The application of compressed sensing for rapid MR imaging, Magnetic Resonance in Medicine, 58 (2007), pp. 1182–1195. [37] W. R. Madych and S. A. Nelson, Polyharmonic cardinal splines, Journal of Approxi- mation Theory, 60 (1990), pp. 141–156. [38] E. Mammen and S. van de Geer, Locally adaptive regression splines, Annals of Statistics, 25 (1997), pp. 387–413. [39] S. Ramani and J. Fessler, Parallel MR image reconstruction using augmented Lagrangian methods, IEEE Transactions on Medical Imaging, 30 (2011), pp. 694 –706. [40] H. Rauhut, K. Schnass, and P. Vandergheynst, Compressed sensing and redundant dictionaries, IEEE Transactions on Information Theory, 54 (2008), pp. 2210–2219. [41] L. I. Rudin, S. Osher, and E. Fatemi, Nonlinear total variation based noise removal algorithms, Physica D, 60 (1992), pp. 259–268. [42] W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 3rd ed., 1987. [43] W. Rudin, Functional Analysis, McGraw-Hill, New York, 1991. [44] O. Scherzer, Taut-string algorithm and regularization programs with g-norm data fit, Jour- nal of Mathematical Imaging and Vision, 23 (2005), pp. 135–143. [45] I. J. Schoenberg, Contributions to the problem of approximation of equidistant data by analytic functions, Quarterly of Applied Mathematics, 4 (1946), pp. 45–99, 112–141. [46] M. H. Schultz and R. S. Varga, L-splines, Numerische Mathematik, 10 (1967), pp. 345– 369. [47] L. L. Schumaker, Spline Functions: Basic Theory, Cambridge University Pressity Press, Cambridge, 3rd edition ed., 2007. [48] G. Steidl, S. Didas, and J. Neumann, Splines in higher order TV regularization, Inter- national Journal of Computer Vision, 70 (2006), pp. 241–255. [49] R. Tibshirani, Regression shrinkage and selection via the Lasso, Journal of the Royal Statistical Society. Series B, 58 (1996), pp. 265–288. [50] M. Unser, Sampling-50 years after Shannon, Proceedings of the IEEE, 88 (2000), pp. 569– 587. [51] M. Unser and T. Blu, Fractional splines and wavelets, SIAM Review, 42 (2000), pp. 43– 67. [52] M. Unser and T. Blu, Cardinal exponential splines: Part I-Theory and filtering algo- rithms, IEEE Transactions on Signal Processing, 53 (2005), pp. 1425–1449. 26 [53] M. Unser and T. Blu, Self-similarity: Part I-Splines and operators, IEEE Transactions on Signal Processing, 55 (2007), pp. 1352–1363. [54] M. Unser, J. Fageot, and H. Gupta, Representer theorems for sparsity-promoting (cid:96)1 regularization, IEEE Transactions on Information Theory, 62 (2016), pp. 5167–5180. [55] M. Unser and P. D. Tafti, An Introduction to Sparse Stochastic Processes, Cambridge University Press, 2014. [56] M. Vetterli, P. Marziliano, and T. Blu, Sampling signals with finite rate of innova- tion, IEEE Transactions on Signal Processing, 50 (2002), pp. 1417–1428. [57] J. Ward and M. Unser, Approximation properties of Sobolev splines and the construction of compactly supported equivalents, SIAM Journal on Mathematical Analysis, 46 (2014), pp. 1843–1858. [58] H. Wendland, Scattered Data Approximations, Cambridge University Press, 2005. 27
1906.00477
1
1906
2019-06-02T20:30:41
Matrix methods for wave equations
[ "math.FA" ]
In analogy to a characterisation of operator matrices generating $C_0$-semigroups due to R. Nagel (\cite{[Na89]}), we give conditions on its entries in order that a $2\times 2$ operator matrix generates a cosine operator function. We apply this to systems of wave equations, to second order initial-boundary value problems, and to overdamped wave equations.
math.FA
math
MATRIX METHODS FOR WAVE EQUATIONS DELIO MUGNOLO Abstract. In analogy to a characterisation of operator matrices generating C0-semigroups due to R. Nagel ([13]), we give conditions on its entries in order that a 2 × 2 operator matrix generates a cosine operator function. We apply this to systems of wave equations, to second order initial-boundary value problems, and to overdamped wave equations. 1. Introduction In [13], R. Nagel started a systematic matrix theory for unbounded operators on Banach spaces. In particular, he described under which assumptions on the entries A, H, and D an operator matrix (cid:18)A H 0 D(cid:19) with diagonal domain D(A) × D(D) generates a C0-semigroup on a suitable product space. However, the theory presented in [13] only accounts for first order problems. In other words, the generation of cosine operator functions is not an issue there. After briefly recalling in Section 2 some known results on cosine operator functions, we state in Section 3 our main results. In analogy to the theory developed in [13, § 3], we characterize when an operator matrix with diagonal domain generates a cosine operator function. In the remainder of our paper, we systematically exploit this abstract result to tackle concrete wave equations of different kinds. The easiest application is to systems of wave equations, possibly on different underlying spaces. In Section 4 we show that the well-posedness of the initial value problem associated with a system of n uncoupled oscillators is not affected by the introduction of coupling terms, provided that the operators modelling such terms are not too unbounded. As a nontrivial application, we consider in Example 4.2 an operator arising in fluid dynamics and already discussed by G. Strohmer in [17] and also, in a slightly modified setting, by Nagel in [13, § 4]. We show that in a L2-setting Strohmer's and Nagel's results can be essentially improved. In Section 5 we consider second order abstract initial-boundary value prob- lems equipped with dynamic boundary conditions. Such a topic has aroused vast interest in recent years: we refer, e.g., to [10], [2], [4], and [11]. The results of Section 3 allow to discuss the well-posedness of a class of wave equations with 2000 Mathematics Subject Classification. 47D09, 35L05, 35L20. 1 2 DELIO MUGNOLO dynamical boundary conditions larger than that considered in [11, § 3]. This in turn allows to improve in Example 5.6 some results obtained by Casarino et al. in [3, § 3]. Finally, in Section 6 we consider a class of second order complete abstract Cauchy problems and give a criterion for their well-posedness. Our assumptions are in fact stronger than those imposed e.g. in [19, § 2.5] and [8, § VI.3b], but in this way we are able to enlarge the space on which such problems are well-posed. 2. Basic facts on cosine operator functions Given a closed operator A on a Banach space X, we denote by [D(A)] the Banach space obtained by endowing the domain of A with its graph norm. We assume the reader to be familiar with the theory of cosine operator functions as presented, e.g., in [9] or [1, § 3.14], and only recall the following, cf. [1, Theorem 3.14.11, Theorem 3.14.17, and Theorem 3.14.18]. (If A generates a cosine operator function (COF), we denote it by (C(t, A))t∈R, and the associated sine operator function (SOF) by (S(t, A))t∈R). Lemma 2.1. Let A be a closed operator on a Banach space X. Then the operator A generates a COF on X if and only if there exists a Banach space V , with [D(A)] ֒→ V ֒→ X, such that the operator matrix IV A :=(cid:18) 0 A 0 (cid:19) , D(A) := D(A) × V, generates a C0-semigroup (etA)t≥0 in V × X. In this case there holds (2.1) etA =(cid:18) C(t, A) AS(t, A) C(t, A)(cid:19) , S(t, A) t ≥ 0. If such a space V exists, then it is unique and is called Kisy´nski space associated with (C(t, A))t∈R. The (unique) product space X = V × X is called phase space associated with (C(t, A))t∈R (or with A). Lemma 2.2. If A generates a COF on a Banach space X, then it also generates an analytic semigroup of angle π 2 on X. Further, the spectrum of A lies inside a parabola. The following similarity and perturbation results have been proved in [1, § 3.14] and [11, § 2]. Lemma 2.3. Let V1, V2, X1, X2 be Banach spaces with V1 ֒→ X1 and V2 ֒→ X2, and let U be an isomorphism from V1 onto V2 and from X1 onto X2. Then an operator A generates a COF with associated phase space V1 × X1 if and only if U AU −1 generates a COF with associated phase space V2 × X2. In this case, there holds U C(t, A)U −1 = C(t, U AU −1), t ∈ R. Lemma 2.4. Let A generate a COF with associated phase space V × X. Then also A + B generates a COF with associated phase space V × X, provided B is an operator that is bounded from [D(A)] to V , or from V to X. MATRIX METHODS FOR WAVE EQUATIONS 3 3. Main results To begin with, we state an analogue of [13, Proposition 3.1] in the context of cosine operator functions. Theorem 3.1. Let A and D be generators of COFs on X and Y , respectively, with associated phase space V × X and W × Y . Consider an operator H that is bounded from [D(D)] to X. Then the operator matrix A :=(cid:18)A H 0 D(cid:19) , D(A) := D(A) × D(D), generates a COF on X × Y , with associated phase space (V × W ) × (X × Y ), if and only if (3.1) C(t − s, A)HS(s, D)ds, t ≥ 0, Z t 0 (3.2) (3.3) can be extended to a family of linear operators from Y to X which is uniformly bounded as t → 0+. In this case, there holds C(t, A) =(cid:18)C(t, A) R t 0 0 C(t − s, A)HS(s, D)ds C(t, D) (cid:19) , t ∈ R, (where we consider the bounded linear extension from Y to X of the upper-right entry) and the associated SOF is S(t, A) =(cid:18)S(t, A) R t 0 0 S(t − s, A)HS(s, D)ds S(t, D) (cid:19) , t ∈ R. Such a SOF is compact if and only if the embeddings [D(A)] ֒→ X and [D(D)] ֒→ Y are both compact. Proof. The operator matrix A generates a COF with associated phase space (V × W ) × (X × Y ) if and only if the reduction matrix A :=(cid:18) 0 A IV ×W 0 (cid:19) , D(A) := (D(A) × D(D)) × (V × W ) , generates a C0-semigroup on (V × W ) × (X × Y ). Define the operator matrix which is an isomorphism from (V × W ) × (X × Y ) onto (V × X) × (W × Y ) with 0 0 0 IX IW 0 0 0 0 0 0 IY 0 0 IX 0 0 0 IW 0 0 0 0 IY ,     . IV 0 0 0 U :=  U−1 :=  A =(cid:18)A H 0 D(cid:19) , IV 0 0 0 D( A) := D(A) × D(D). Then the similar operator matrix A := UAU−1 becomes 4 DELIO MUGNOLO Here A and D are the reduction operator matrices and respectively, while H is given by IV IW A :=(cid:18) 0 D :=(cid:18) 0 A 0 (cid:19) , D 0 (cid:19) , H :=(cid:18) 0 0 H 0(cid:19) , D(A) := D(A) × V, D(D) := D(D) × W, D(H) := D(D). By Lemma 2.1, the operators A and D generate C0-semigroups on V × X and W ×Y , respectively. Moreover H ∈ L([D(D)], V ×X), and a direct computation shows that e(t−s)AHesD =(cid:18)S(t − s, A)HC(s, D) S(t − s, A)HS(s, D) C(t − s, A)HC(s, D) C(t − s, A)HS(s, D)(cid:19) , 0 ≤ s ≤ t. By virtue of [13, Proposition 3.1] we obtain that A generates a C0-semigroup on (W × X) × (W × Y ) if and only if the family of operators e(t−s)AHesDds, t ≥ 0, Z t 0 from W × Y to V × X is uniformly bounded as t → 0+. Hence, if A gen- 0 C(t − erates a C0-semigroup, or equivalently if A generates a COF, then R t s, A)HS(s, D)ds is uniformly bounded as t → 0+. Again by [13, Proposition 3.1] etA =(cid:18)etA R t 0 0 e(t−s)AHesDds etD (cid:19) , t ≥ 0. By similarity, etA = etU−1 AU = U−1etAU, t ≥ 0. Thus, a direct computation shows that the semigroup generated by A on the space (V × W ) × (X × Y ) is given by 0 S(t − s, A)HC(s, D)ds 0 S(t − s, A)HS(s, D)ds 0 C(t, A) R t AS(t, A) R t 0   C(t, D) DS(t, D) 0 S(t, A) R t C(t, A) R t 0 S(t, D) C(t, D)   0 C(t − s, A)HC(s, D)ds 0 C(t − s, A)HS(s, D)ds for t ≥ 0. Since by assumption A generates a C0-semigroup on the space (V × W )×(X × Y ), comparing the above formula with (2.1) yields (3.2) and (3.3). One can also check directly that the lower-right block-entry defines a COF on X × Y . Further, integrating by parts one sees that the upper-right and lower- right block-entries can be obtained by integrating the upper-left and lower- left block-entries, respectively, and moreover that the diagonal blocks coincide. Hence, by definition of SOF, all the blocks are strongly continuous families as soon as the lower-right is strongly continuous. Consequently, if the family MATRIX METHODS FOR WAVE EQUATIONS 5 R t R t 0 C(t − s, A)HS(s, D)ds is uniformly bounded as t → 0+, then the family 0 e(t−s)AHesDds is uniformly bounded as t → 0+. Finally, it is known that a SOF is compact if and only if its generator has (cid:3) compact resolvent, cf. [18, Propositio 2.3], and the claim follows. (cid:3) Observe that the operators defined in (3.1) are in general only bounded from W to X. In Theorem 3.1 we however required that they are bounded from the larger space Y to X. Such an extension can usually be performed whenever the integrated operator provides some kind of regularizing effect. However, usually COFs do not enjoys any regularity property (see [18, Propositio 4.1]). Therefore, in most cases the following analogue of [13, Corollary 3.2] can be applied more easily. Proposition 3.2. Let A and D be closed operators on the Banach spaces X and Y , respectively. Consider further Banach spaces V, W such that [D(A)] ֒→ V ֒→ X and [D(D)] ֒→ W ֒→ Y . Assume moreover - the operator H to be bounded from [D(D)] to V , or from W to X, and - the operator K to be bounded from [D(A)] to W , or from V to Y . Then the operator matrix K D(cid:19) , A :=(cid:18)A H D(A) := D(A) × D(D), generates a COF with associated phase space (V × W ) × (X × Y ) if and only if A and D generate COFs with associated phase space V × X and W × Y , respectively. In this case, (S(t, A))t∈R is compact if and only if the embeddings [D(A)] ֒→ X and [D(D)] ֒→ Y are both compact. Proof. The diagonal matrix A0 :=(cid:18)A 0 0 D(cid:19) , D(A0) := D(A), generates a COF with associated phase space (V × W ) × (X × Y ) if and only if A and D generate COFs with associated phase space V × X and W × Y , respectively. Consider now perturbations of A0 given by the operator matrices H :=(cid:18)0 H 0(cid:19) 0 and 0 K :=(cid:18) 0 K 0(cid:19) . Observe that both H and K are, by assumption, either bounded from [D(A)] × [D(D)] to V × W , or from V × W to X × Y . By Lemma 2.4 also their sum A0 + H + K = A generates a COF with associated phase space (V × W ) × (X × Y ). (cid:3) (cid:3) Remark 3.3. In the special case of A ∈ L(X), Proposition 3.2 reads as follows: Let D be a closed operator on Y , and consider a further Banach space W such that [D(D)] ֒→ W ֒→ Y . Assume moreover that H ∈ L([D(D)], X) and K ∈ L(X, Y ). Then A :=(cid:18)A H K D(cid:19) , D(A) := D(A) × D(D), 6 DELIO MUGNOLO generates a COF with associated phase space (X × W ) × (X × Y ) if and only if D generates a COF with associated phase space W × Y . 4. Systems of abstract wave equations We consider in this section systems of n abstract wave equations and show how they can be solved by means of the results of Section 3. In the trivial case of n uncoupled oscillators modelled by (USn) u1(t) = A1u1(t), u2(t) = A2u2(t), ... t ∈ R, t ∈ R, un(t) = Anun(t), t ∈ R,   it is clear that the initial value problem associated with (USn) is well-posed (in a natural sense) if and only if each operator Ai, i = 1, . . . , n, generates a COF (with suitable associated phase spaces Vi × Xi). If however there is an interplay among the single oscillators given by u1(t) = A1u(t) + B1 u2(t) = A2u(t) + C 2 ... un(t) = Anu(t) + C n (CSn)   assumptions on the operators Bh to obtain well-posedness. 2 u2(t) + B1 1 u1(t) + B2 3 u3(t) . . . + B1 3 u3(t) . . . + B2 nun(t), nun(t), t ∈ R, t ∈ R, 1 u1(t) + C n k and C k 2 u2(t) . . . + C n h, 1 ≤ h < k ≤ n, are needed in order n−1un−1(t), t ∈ R, Theorem 4.1. Let the initial value problem associated with (USn) be well- posed, so that Vi ×Xi is the phase space associated with Ai, 1 ≤ i ≤ n. Consider operators Bh h , 1 ≤ h < k ≤ n, such that k and C k - Bh - C k k ∈ L([D(Ak)], Vh) or Bh h ∈ L([D(Ah)], Vk) or C k k ∈ L(Vk, Xh), and h ∈ L(Vh, Xk). associated phase space (Qn Then the initial value problem associated with (CSn) is governed by a COF with i=1 Xi), and in particular it is well-posed. Proof. For the sake of simplicity we only discuss the case n = 2, since the general case can be proved by induction on n. Consider the system i=1 Vi) × (Qn (CS2) (cid:26) u1(t) = A1u1(t) + B1 u2(t) = A2u2(t) + C 2 2 u2(t), 1 u1(t), t ∈ R, t ∈ R, which can be written as an abstract wave equation on the Banach space X1 × X2, where u(t) = Au(t), t ∈ R, A :=(cid:18)A1 B1 1 A2(cid:19) , C 2 2 and D(A) := D(A1) × D(A2), u :=(cid:18)u1 u2(cid:19) . MATRIX METHODS FOR WAVE EQUATIONS 7 By assumption, A1 and A2 generate COFs with associated phase space V1 × X1 and V2 × X2, respectively. Due to the assumptions on B1 1 the claim follows from Proposition 3.2. (cid:3) 2 and C 2 (cid:3) Example 4.2. Following work of A. Matsumura and T. Nishida ([15], [16]), some linearized equations from fluid dynamics in an open, bounded domain Ω ⊂ Rn lead to consider the operator matrix A1 B1 C 2 1 A2 C 3 0 1 3 , C 2 2 , B1 where the operator entries A1, A2, B1 1 are defined below. Such a setting has been thoroughly discussed in [17] and, in a simplified and slightly different version, in [13, § 4]. Both authors show, by different means, that A A :=  0   , 2 B1 3 0 generates an analytic semigroup on (cid:0)Lp(Ω)(cid:1)n × Lp(Ω) × W 1,p(Ω), 1 < p < ∞; in [17] some description of the spectrum of A is also given, and it is shown that the generated semigroup has angle of analyticity ≥ π 4 . 1 , C 3 Our aim is to show that such an operator matrix, equipped with domain D(A) := D(A1) × D(A2) × H 1(Ω), is in fact the generator of a COF on the Hilbert space(cid:0)L2(Ω)(cid:1)n×L2(Ω)×H 1(Ω). Consequently, by Lemma 2.2 it also generates an analytic semigroup of angle π 2 on the same space, and moreover its spectrum is contained inside a parabola. Here A1 := µ1∆n + µ2grad·div, with (µ1, µ2) ∈ R2 + \ {0, 0}. If ∂Ω is smooth enough, then integrating by parts a direct computation shows that A1 is the operator associated with the sesquilin- D(A1) :=(cid:0)H 2(Ω) ∩ H 1 0 (Ω)(cid:1)n , ear form a on the Hilbert space (cid:0)L2(Ω)(cid:1)n defined by n a(f, g) := µ1 ∂fi ∂xj ∂gi ∂xj + µ2 ∂fi ∂xi ∂gj ∂xj dx One sees that a is symmetric, closed, and densely defined. Moreover, a is positive, since a(f, f ) := µ1 n Xi=1ZΩ divfi2dx + µ2ZΩ gradf 2 dx. It is then well-known (see, e.g., [6, Theorem 1.2.1]) that the operator A1 asso- ciated with a is self-adjoint and dissipative, hence the generator of a COF with associated phase space V1 × X1 :=(cid:0)H 1 Further, for µ3 > 0, we define A2 := µ3∆ on Ω equipped with either (in [17]) Robin, or (in [13, § 4]) Dirichlet boundary conditions. In both cases A2 gen- erates a COF, and it is well-known (see [9, Chapter 4]) that the associated 0 (Ω)(cid:1)n ×(cid:0)L2(Ω)(cid:1)n. for all Xi,j=1ZΩ  , g :=   g1 ... gn   f1 ... fn f :=  ∈ D(a) :=(cid:0)H 1 0 (Ω)(cid:1)n . 8 DELIO MUGNOLO 0 (Ω) × L2(Ω), respectively. Also phase space V2 × X2 is H 1(Ω) × L2(Ω) or H 1 any bounded operator on H 1(Ω) generates a COF with associated phase space H 1(Ω) × H 1(Ω). Define finally B1 2 := p1grad, B1 3 := p2grad, and C 2 1 := p3div, C 3 1 := p4div, where p1, p2, p3, p4 are constants. Since the operator grad is bounded from 3 ∈ L(V3, X1). H 1(Ω) to (cid:0)L2(Ω)(cid:1)n, it follows that B1 Similarly, the operator div is bounded from (cid:0)H 1(Ω)(cid:1)n to L2(Ω) as well as from (cid:0)H 2(Ω)(cid:1)n to H 1(Ω), and accordingly C 2 1 ∈ L(V1, X2) and also C 3 We conclude by Theorem 4.1 that the whole operator matrix A generates a 2 ∈ L(V2, X1) and also B1 L([D(A1)], V3). 1 ∈ COF The associated phase space is if A2 is equipped with Robin boundary conditions, or rather (cid:16)(cid:0)H 1 (cid:16)(cid:0)H 1 0 (Ω)(cid:1)n × H 1(Ω) × H 1(Ω)(cid:17) ×(cid:16)(cid:0)L2(Ω)(cid:1)n × L2(Ω) × H 1(Ω)(cid:17) 0 (Ω) × H 1(Ω)(cid:17) ×(cid:16)(cid:0)L2(Ω)(cid:1)n × L2(Ω) × H 1(Ω)(cid:17) 0 (Ω)(cid:1)n × H 1 if A2 is equipped with Dirichlet boundary conditions. 5. Abstract initial -- boundary value problems We impose the following assumptions throughout this section and refer to [3] and [11] for motivation. Assumption 5.1. (1) X and Y are Banach spaces such that Y ֒→ X. (2) ∂X and ∂Y are Banach spaces such that ∂Y ֒→ ∂X. (3) A : D(A) → X is linear with D(A) ⊂ Y . (4) L : D(A) → ∂X is linear and surjective. (5) A0 := A ker(L) is densely defined and has nonempty resolvent set. (6) (cid:18)A L(cid:19) : D(A) → X × ∂X is closed1 as an operator from X to X × ∂X. (7) B : [D(A)L] → ∂X is linear and bounded; further, B is bounded either from [D(A0)] to ∂Y , or from Y to ∂X. (8) B : D( B) ⊂ ∂X → ∂X is linear and closed, with D( B) ⊂ ∂Y . Under the Assumptions 5.1 one can define a solution operator DA,L of the λ abstract (eigenvalue) Dirichlet problem (ADP) (cid:26) Au = λu, Lu = w, 1Observe that, under the Assumption 5.1.(6), we obtain a Banach space by endowing D(A) with the graph norm of (cid:0)A L(cid:1), i.e., kuk(A L) := kukX + kAukX + kLuk∂X . We denote this Banach space by [D(A)L]. MATRIX METHODS FOR WAVE EQUATIONS 9 for all λ ∈ ρ(A0). More precisely, the following holds, cf. [3, Lemma 2.3] and [11, Lemma 3.2]. Lemma 5.2. The problem (ADP) admits a unique solution u := DA,L λ w for all w ∈ ∂X and λ ∈ ρ(A0). Moreover, the solution operator DA,L is bounded from ∂X to Z for every Banach space Z satisfying D(A∞) ⊂ Z ֒→ X. In particular, DA,L λ λ ∈ L(∂X, Y ). λ ∈ L(∂X, [D(A)L]) as well as DA,L Observe that, by Lemma 5.2, BDA,L λ ∈ L(∂X) for all λ ∈ ρ(A0). We want to discuss well-posedness for a second order abstract initial-boundary value problem of the form (AIBPV2) u(t) = Au(t), w(t) = Bu(t) + Bw(t), w(t) = Lu(t), u(0) = f, w(0) = h, u(0) = g, w(0) = j. t ∈ R, t ∈ R, t ∈ R,   Observe that the equations on the first and the fourth line take place on the Banach space X, while the remainders on the Banach space ∂X. We begin by re-writing (AIBVP2) as a more standard second order abstract Cauchy problem (ACP 2) (cid:26) u(t) = Au(t), u(0) = f, u(0) = g, t ∈ R, on the product space X := X × ∂X, where (5.1) B B(cid:19) , A :=(cid:18)A 0 D( A) :=(cid:26)(cid:18)u w(cid:19) ∈ D(A) × D( B) : Lu = w(cid:27) , is an operator matrix with coupled domain on X . Here the new variable u(·) and the inital data f, g are to be understood as u(t) :=(cid:18) u(t) Lu(t)(cid:19) for t ∈ R, f :=(cid:18)f h(cid:19) , g :=(cid:18)g j(cid:19) . Taking the components of (ACP 2) in the factor spaces of X yields the first two equations in (AIBVP2), while the coupling relation Lu(t) = w(t), t ∈ R, is incorporated in the domain of the operator matrix A. We can thus equivalently investigate (ACP 2) instead of (AIBVP2). In particular, we are interested in characterizing whether A generates a COF in terms of analogue properties of A0 and B. Taking into account Lemma 5.2 and [12, Lemma 3.10] (see also [7, § 2]), a direct matrix computation yields the following. Lemma 5.3. Let λ ∈ ρ(A0). Then A − λ is similar to the operator matrix (5.2) Aλ := A0 − DA,L λ B − λ −DA,L B ( B + BDA,L λ − λ λ B + BDA,L λ − λ) ! , 10 DELIO MUGNOLO with diagonal domain D(Aλ) := D(A0) × D( B). The similarity transformation is given by the operator Mλ :=(cid:18)IX −DA,L I∂X (cid:19) , 0 λ which is an isomorphism on Y := Y × ∂Y as well as on X . Theorem 5.4. Under the Assumptions 5.1, the operator matrix A defined in (5.1) generates a COF with associated phase space Y × X if and only if A0 and B generate COFs with associated phase space Y × X and ∂Y × ∂X, respectively. In this case, (S(t, A))t∈R is compact if and only if the embeddings [D(A0)] ֒→ X and [D( B)] ֒→ ∂X are both compact. Proof. Take λ ∈ ρ(A0). By Lemma 5.3 the operator matrix A − λ is similar to Aλ defined in (5.2), and the similarity transformation is performed by Mλ, which is an isomorphism on X as well as on the candidate Kisy´nski space Y. It follows by Lemma 2.3 that A−λ, and hence A generates a COF with associated phase space Y × X if and only if the similar operator Aλ generates a COF with same associated phase space. We decompose Aλ =(cid:18)A0 −DA,L λ B 0 B (cid:19) +(cid:18)−DA,L λ B 0 B 0(cid:19) + −λ DA,L 0 λ (λ − BDA,L λ − λ ! . BDA,L ) λ Since the operator matrix Aλ has diagonal domain D(Aλ) = D(A0) × D( B), we are now in the position to apply the results of Section 3. One sees that the second operator on the right hand side is bounded from [D(Aλ)] to Y or from Y to X , while the third one is bounded on X . Thus, by Lemma 2.4 we conclude that A generates a COF with associated phase space Y × X if and only if (cid:18)A0 −DA,L λ B 0 B (cid:19) with domain D(A0) × D( B) generates a COF with phase space Y × X . Since DA,L claim follows by Proposition 3.2. (cid:3) λ B ∈ L([D( B)], Y ), the (cid:3) By Lemma 2.2 we hence obtain the following. Corollary 5.5. Under the Assumptions 5.1, let A0 and B generate COFs with associated phase space Y × X and ∂Y × ∂X, respectively. Then the operator matrix A defined in (5.1) generates an analytic semigroup of angle π 2 in X ×∂X. Further, such an analytic semigroup is compact if and only if the embeddings [D(A0)] ֒→ X and [D( B)] ֒→ ∂X are both compact. We can now revisit a problem considered in [3] and improve the result ob- tained therein. MATRIX METHODS FOR WAVE EQUATIONS 11 Example 5.6. Let Ω be a bounded open domain of Rn with boundary ∂Ω smooth enough, and consider the second order initial-boundary value problem u(t, x) = ∆u(t, x), w(t, z) = Bu(t, z) + ∆w(t, z), w(t, z) = ∂u u(0, x) = f (x), w(0, z) = h(z), ∂ν (t, z), u(0, x) = g(x), x ∈ Ω, w(0, z) = j(z), z ∈ ∂Ω. t ∈ R, x ∈ Ω, t ∈ R, z ∈ ∂Ω, t ∈ R, z ∈ ∂Ω,   (5.3) Set X := L2(Ω), Y := H 1(Ω), ∂X := L2(∂Ω), and ∂Y := H 1(∂Ω). Define the operators A := ∆, D(A) :=nu ∈ H 3 2 (Ω) : ∆u ∈ L2(Ω)o , L := ∂ ∂ν B := ∆, , D(L) := D(A), D( B) := H 2(∂Ω), i.e., B is the Laplace -- Beltrami operator on ∂Ω. It has been shown in [3, § 3] that A, L, and B satisfy the Assumptions 5.1. In particular, the restriction A0 of A to ker(L) is the Neumann Laplacian, which generates a COF with associated phase space H 1(Ω) × L2(Ω) by [9, Theorem IV.5.1]. Further, the Laplace -- Beltrami operator is by definition self- adjoint and dissipative on L2(∂Ω), and its form domain is H 1(∂Ω). Hence B generates a COF with associated phase space H 1(∂Ω) × L2(∂Ω). By Theorem 5.4 we conclude that the problem (5.3) is governed by a COF with associated phase space (H 1(Ω) × H 1(∂Ω)) × (L2(Ω) × L2(∂Ω)) whenever B is a bounded operator from H 1(Ω) to L2(∂Ω). In other words, (5.3) admits a unique classical solution if, in particular, f ∈ H 2(Ω), g ∈ H 1(Ω), h ∈ H 2(∂Ω), and j ∈ H 1(∂Ω). Finally, due to the boundednes of Ω, and hence to the com- pactness of the embeddings H 1(Ω) ֒→ L2(Ω) and H 1(∂Ω) ֒→ L2(∂Ω), we can conclude that the SOF associated with the COF that governs (5.3) is compact. This also improves the result obtained for the first order case in [3, § 3]. We consider a complete second order abstract Cauchy problem 6. Damped problems (cACP2) (cid:26) u(t) = Au(t) + C u(t), u(0) = f, u(0) = g. t ≥ 0, In the case of C "subordinated" to A (i.e., when −C is somehow related to a fractional power of −A) the well-posedness of (cACP2) has been discussed, among others, by Fattorini in [9, Chapter VIII], by Chen -- Triggiani in [5], and by Xiao -- Liang in [19, Chapters 4 -- 6]. In the overdamped case (i.e., when C is "more unbounded" than A) the treatment is easier and several well-posedness results have been obtained, under the essential assumption that C generates a C0-semigroup, in [14], [19], and [8]. 12 DELIO MUGNOLO A natural step is to introduce the reduction matrix (6.1) D(A) = D(A) × D(C). A :=(cid:18) 0 A ID(C) C (cid:19) , Its generator property has already been studied, under appropriate assump- tions: we refer, e.g., to [14, § 5 -- 6] and [8, § VI.3]. The prototype result is in fact the following (see [19, Theorem 2.5.2]): Let A ∈ L(X). Then (cACP2) is well-posed if and only if C generates a C0-semigroup on X. An analogue of this can be proved in the context of COFs taking into account the results of Section 3. Proposition 6.1. Let C be a closed operator on a Banach space X and let V be a Banach space such that [D(C)] ֒→ V ֒→ X and A ∈ L(V, X). Then A (with domain V × D(C)) generates a COF on V × X if and only if C generates a COF with associated phase space V × X. Proof. We can regard 0 as a generates a COF with associated phase space V × V . Since A ∈ L(V, X) and of course ID(C) ∈ L([D(C)], V ), it follows from Remark 3.3 that C generate a COF with associated phase space V × X if and only if A generates a COF with associated phase space (V × V ) × (V × X). (cid:3) (cid:3) We can now state the following result on very strongly damped wave equa- tions. It generalizes the above mentioned [19, Theorem 2.5.2] because we do not assume A to be bounded on X. Theorem 6.2. Let C generate a COF with associated phase space V × X. If A ∈ L(V, X), then the operator matrix A (with domain V × D(C)) defined in (6.1) generates an analytic semigroup of angle π 2 on V × X. In particular, (cACP2) is well-posed, and in fact it admits a unique classical solution u for all initial data f ∈ V , g ∈ X. Example 6.3. Consider the initial value problem   (6.2) u(t, x) = ∆u(t, x) − ∆2 u(t, x), u(t, z) = u(t, z) = ∆ u(t, z) = 0, u(0, x) = f (x), u(0, x) = g(x), t ≥ 0, x ∈ Ω, t ≥ 0, z ∈ ∂Ω, x ∈ Ω. for an overdamped wave equation on an open, bounded domain Ω ⊂ Rn with Lipschitz boundary. Set Define V := H 2(Ω) ∩ H 1 0 (Ω) and X := L2(Ω). A := ∆, D(A) := V, C := −∆2, D(C) :=(cid:8)u ∈ H 4(Ω) ∩ H 1 0 (Ω) : ∆u∂Ω = 0(cid:9) , and observe that −C is the square of A. Then (6.2) can be written in the ab- stract form (cACP2). Since the operator C is self-adjoint and strictly negative, 2 ] × X = (H 2(Ω) ∩ it generates a COF with associated phase space [D(−C) 0 (Ω)) × L2(Ω). Moreover, the Laplacian A is bounded from H 2(Ω) to L2(Ω), H 1 1 MATRIX METHODS FOR WAVE EQUATIONS 13 hence we conclude by Theorem 6.2 that the operator matrix A defined as in (6.1) generates an analytic semigroup of angle π In particular, the problem (6.2) admits a unique classical solution for all ini- tial data g ∈ L2(Ω) and f ∈ H 2(Ω)∩H 1 0 (Ω): while applying [8, Corollary VI.3.4] to the same problem yields existence and uniqueness of a classical solution only for f ∈ D(C), i.e., for u(0, ·) in a class of H 4(Ω)-functions. 2 on (cid:0)H 2(Ω) ∩ H 1 0 (Ω)(cid:1) × L2(Ω). References [1] W. Arendt, C.J.K. Batty, M. Hieber, and F. Neubrander, Vector-valued Laplace Trans- forms and Cauchy Problems, Monographs in Mathematics 96, Birkhauser, Basel, 2001. [2] A. B´atkai and K.-J. Engel, Abstract wave equations with generalized Wentzell boundary conditions. J. Diff. Equations 204 (2004), 1 -- 20. [3] V. Casarino, K.-J. Engel, R. Nagel, and G. Nickel, A semigroup approach to boundary feedback systems, Integral Equations Oper. Theory 47 (2003), 289 -- 306. [4] V. Casarino, K.-J. Engel, G. Nickel, and S. Piazzera, Decoupling techniques for wave equations with dynamic boundary conditions, Discrete Contin. Dyn. Syst. 12 (2005), 761 -- 772. [5] S.P. Chen and R. Triggiani, Characterization of domains of fractional powers of certain operators arising in elastic systems, and applications, J. Diff. Equations 88 (1990), 279 -- 293. [6] E.B. Davies, Heat Kernels and Spectral Theory, Cambridge Tracts in Mathematics 92, Cambridge University Press, Cambridge, 1990. [7] K.-J. Engel, Spectral theory and generator property for one-sided coupled operator matrices, Semigroup Forum 58 (1999), 267 -- 295. [8] K.-J. Engel and R. Nagel, One-Parameter Semigroups for Linear Evolution Equations, Graduate Texts in Mathematics 194, Springer-Verlag, New York, 2000. [9] H.O. Fattorini, Second Order Linear Differential Equations in Banach Spaces, Mathemat- ics Studies 108, North-Holland , Amsterdam, 1985. [10] A. Favini, G.R. Goldstein, J.A. Goldstein, and S. Romanelli, The one dimensional wave equation with Wentzell boundary conditions, in: S. Aicovici and N. Pavel (eds.), "Differen- tial Equations and Control Theory" (Proceedings Athens 2000), Lecture Notes in Pure and Applied Mathematics 225, Marcel Dekker, New York, 2001, 139 -- 145. [11] D. Mugnolo, Operator matrices as generators of cosine operator functions. Integral Equa- tions Oper. Theory. Published online: 1st October 2005. [12] D. Mugnolo, Abstract wave equations with acoustic boundary conditions, Math. Nachr. 279 (2006), 293 -- 318. [13] R. Nagel, Towards a "matrix theory" for unbounded operator matrices, Math. Z. 201 (1989), 57 -- 68. [14] F. Neubrander, Well-posedness of higher order abstract Cauchy problems, Trans. Am. Math. Soc. 295 (1986), 257-290. [15] T. Nishida and A. Matsumura, The initial value problem for the equations of motion of compressible, viscous, and heat conducting fluids, Proc. Jap. Acad. Ser. A 55 (1979), 337 -- 342. [16] T. Nishida and A. Matsumura, The initial value problem for the equations of motion of viscous and heat conductive gases, J. Math. Kyoto Univ. 20 (1980), 67 -- 104. [17] G. Strohmer, About the resolvent of an operator from fluid dynamics, Math. Z. 194 (1987), 183 -- 191. [18] C.C. Travis and G.F. Webb, Compactness, regularity, and uniform continuity properties of strongly continuous cosine families, Houston J. Math. 3 (1977), 555 -- 567. [19] T.-J. Xiao and J. Liang, The Cauchy Problem for Higher-Order Abstract Differential Equations, Lecture Notes in Mathematics 1701, Springer-Verlag, 1998. 14 DELIO MUGNOLO Abteilung Angewandte Analysis der Universitat, Helmholtzstrasse 18, D-89081 Ulm, Germany and Dipartimento di Matematica dell'Universit`a degli Studi, Via Orabona 4, I-70125 Bari, Italy E-mail address: [email protected]
1708.07416
1
1708
2017-08-22T18:58:10
Sobolev, Besov and Paley-Wiener vectors in Banach and Hilbert spaces
[ "math.FA" ]
We consider Banach spaces equipped with a set of strongly continuous bounded semigroups satisfying certain conditions. Using these semigroups we introduce an analog of a modulus of continuity and define analogs of Besov norms. A generalization of a classical interpolation theorem is proven in which the role of Sobolev spaces is played by subspaces defined in terms of infinitesimal operators of these semigroups. We show that our assumptions about a given set of semigroups are satisfied in the case of a strongly continuous bounded representation of a Lie group. In the case of a unitary representation in a Hilbert space we consider an analog of the Laplace operator and use it to define Paley-Wiener vectors. It allows us to develop a generalization of the Shannon-type sampling in Paley-Wiener subspaces and to construct Paley-Wiener nearly Parseval frames in the entire Hilbert space. It is shown that Besov spaces defined previously in terms of the modulus of continuity can be described in terms of approximation by Paley-Wiener vectors and also in terms of the frame coefficients. Throughout the paper we extensively use theory of interpolation and approximation spaces. The paper ends with applications of our results to function spaces on homogeneous manifolds.
math.FA
math
SOBOLEV, BESOV AND PALEY-WIENER VECTORS IN BANACH AND HILBERT SPACES Dedicated to 100s Birthday of my teacher S.G. Krein Isaac Z. Pesenson 1 Keywords: One-parameter semigroups, representations of Lie groups in Banach and Hilbert spaces, moduli of continuity, sampling theorem, frames Subjclass: 43A85, 41A17. Abstract. We consider Banach spaces equipped with a set of strongly con- tinuous bounded semigroups satisfying certain conditions. Using these semi- groups we introduce an analog of a modulus of continuity and define analogs of Besov norms. A generalization of a classical interpolation theorem is proven in which the role of Sobolev spaces is played by subspaces defined in terms of infinitesimal operators of these semigroups. We show that our assumptions about a given set of semigroups are satisfied in the case of a strongly continuous bounded representation of a Lie group. In the case of a unitary representation in a Hilbert space we consider an analog of the Laplace operator and use it to define Paley-Wiener vectors. It allows us to develop a generalization of the Shannon-type sampling in Paley-Wiener subspaces and to construct Paley- Wiener nearly Parseval frames in the entire Hilbert space. It is shown that Besov spaces defined previously in terms of the modulus of continuity can be described in terms of approximation by Paley-Wiener vectors and also in terms of the frame coefficients. Throughout the paper we extensively use theory of interpolation and approximation spaces. The paper ends with applications of our results to function spaces on homogeneous manifolds. 1. Introduction and Main Results I am honored to have had Selim Grigorievich Krein as my academic advisor and co-author. Without his steady, strong interest in my research and his support I would not have been able to have the privilege of becoming a mathematician. His inspiring encouragement shaped not only my professional trajectory but my life in general. The first five sections of this paper are devoted to Sobolev and Besov subspaces and to relevant moduli of continuity in Banach spaces. This theory is rooted in my results obtained in 70s: [19]-[21], [23], [9]. It is a far going generalization of the one-dimensional theory by J. Lions [11] and J. Lions-J. Peetre [12] which I learned from the nice book by P. Butzer and H. Berens [3]. The Paley-Wiener vectors and corresponding approximation theory in abstract Hilbert spaces were introduced in 80s in my papers [22], [24], [25], [9] (see sections 6 and 7 below). In papers [26]- [34] I used the notion of Paley-Wiener vectors to prove Shannon-type sampling theorems on manifolds and in general Hilbert spaces. The construction of frames 1Department of Mathematics, Temple University, Philadelphia, PA 19122; [email protected] 1 2 SOBOLEV, BESOV AND PALEY-WIENER VECTORS and description of Besov subspaces on manifolds and in Hilbert spaces in terms of frame coefficients is rather recent development and can be found in articles with my collaborators [7], [38], [6]. We consider a Banach space E and operators D1, D2, ..., Dd which generate strongly continuous uniformly bounded semigroups T1(t), T2(t), ..., Td(t), kT (t)k ≤ 1, t ≥ 0. An analog of a Sobolev space is introduced as the space Er of vectors in E for which the following norm is finite fEr = kfkE + rXk=1 X1≤j1,...jk≤d kDj1...Djk fkE, where r ∈ N, f ∈ E. By using the closed graph theorem and the fact that each Di is a closed operator in E, one can show that this norm is equivalent to the norm (1.1) kfkr = kfkE + X1≤i1,...,ir≤d kDi1...Dik fkE, r ∈ N. The mixed modulus of continuity is introduced as Ωr(s, f ) = (1.2) X1≤j1,...,jr≤d sup ... 0≤τj1≤s sup 0≤τjr ≤sk (Tj1(τj1 ) − I) ... (Tjr (τjr ) − I) fkE, where f ∈ E, r ∈ N, and I is the identity operator in E. Let D(Di) be the domain of the operator Di. For every f ∈ E we introduce a vector-valued function T f : Rd 7−→ E defined as T f (t1, t2, ..., td) = T1(t1)T2(t2)...Td(td)f. Assumption 1. We assume that the following properties hold. (1) The set E1 =Td i=1 D(Di) is dense in E. (2) The set E1 is invariant with respect to all Ti(t), 1 ≤ i ≤ d, t ≥ 0. (3) For every 1 ≤ i ≤ d, every f ∈ E1 and all t = (t1, ..., td) in the standart open unit ball U in Rd (1.3) DiT f (t1, ..., td) = ζk i (t) (∂kT f ) (t1, ..., td), dXk=1 where ζk i (t) belong to C∞(U ), ∂k = ∂ ∂tk . Remark 1.1. Since E1 is invariant with respect to all bounded operators Ti(ti), 1 ≤ i ≤ d, we obtain for every f ∈ E1 T1(t1)T2(t2)...DkTk(tk)...Td(td)f = T1(t1)T2(t2)...Tk−1(tk−1) lim s→0 1 s (Tk(tk + s) − Tk(tk)) Tk+1(tk+1)...Td(td)f = 1 s lim s→0 (1.4) T1(t1)T2(t2)...Tk−1(tk−1) (Tk(tk + s) − Tk(tk)) Tk+1(tk+1)...Td(td)f = ∂ ∂tk T1(t1)T2(t2)...Td(td)f =(cid:18) ∂ ∂tk T f(cid:19) (t1, ..., td). SOBOLEV, BESOV AND PALEY-WIENER VECTORS 3 Thus the formula (1.3) can be rewritten as (1.5) DiT1(t1)T2(t2)...Td(td)f = ζk i (t)T1(t1)T2(t2)...DkTk(tk)...Td(td)f, dXk=1 where ζk i (t) belong to C∞(U ), t = (t1, ..., td) ∈ U . Remark 1.2. When semigroups commute with each other Ti(ti)Tj(tj) = Tj(tj )Ti(ti), 1 ≤ i, j ≤ d, ti, tj ≥ 0, then ζk i = δk i . It is well known that most of remarkable properties of the so-called Besov func- tional spaces follow from the fact that they are interpolation spaces (see section 4 below) between two Sobolev spaces [3], [10]. For this reason we define Besov spaces by the formula (1.6) Eα,q = (E, Er)K α/r, q , 0 < α < r ∈ N, 1 ≤ p, q ≤ ∞, where K is the so-called Peetre's interpolation functor (see section 4 below). The main result is the following. Theorem 1.3. If Assumption 1 is satisfied then the following holds true. (1) The functionals Ωr(s, f ) and K(sr, f, E, Er) are equivalent. Namely, there exist constants c > 0, C > 0, such that for all f ∈ E, t ≥ 0 c Ωr(s, f ) ≤ K(sr, f, E, Er) ≤ C (Ωr(s, f ) + min(sr, 1)kfk) . (2) The norm of the Besov space Eα,q = (E, Er)K α/r,q , 0 < α < r ∈ N, 1 ≤ p, q ≤ ∞, is equivalent to the norm kfkE +(cid:18)Z ∞ 0 (s−αΩr(s, f ))q ds s (cid:19)1/q , 1 ≤ q < ∞, with the usual modifications for q = ∞. (3) The following isomorphism holds true (E, Er)K where 0 ≤ k1 < α < k2 ≤ r ∈ N, 1 ≤ q ≤ ∞. (4) If α is not integer then the norm (1.8) is equivalent to the norm (α−k1)/(k2−k1),q , α/r,q =(cid:0)Ek1, Ek2(cid:1)K s (cid:19)1/q (1.7) (1.8) (1.9) kfkE[α] + X1≤j1,...,j[α]≤d(cid:18)Z ∞ 0 (cid:16)s[α]−αΩ1(s, Dj1 ...Dj[α] f )(cid:17)q ds where [α] is the integer part of α. (5) If α = k ∈ N is an integer then the norm (1.8) is equivalent to the norm (Zygmund condition) (1.10) kfkEk−1 + X1≤j1,...,jk−1≤d(cid:18)Z ∞ 0 (cid:0)s−1Ω2(s, Dj1...Djk−1 f )(cid:1)q ds s (cid:19)1/q . Next we make another assumption. Assumption 2. In addition to Assumption 1 we assume that E = H is a Hilbert space and the following properties hold. 1 + ... + D2 d is a non-negative self-adjoint operator in (1) The operator L = D2 H. 4 SOBOLEV, BESOV AND PALEY-WIENER VECTORS (2) The domain D(Lk/2), k ∈ N, of the non-negative square root Lk/2 coincides with the space Hk and the norms (1.1) and kfkH+kLk/2fkH are equivalent. This assumption allows us to introduce notion of Paley-Wiener vectors (bandlim- ited vectors) (Definition 1) and to prove an abstract version of the Paley-Wiener Theorem (Theorem 6.3). To formulate an analog of the Shannon-type sampling in abstract Paley-Wiener spaces (Theorem 6.5) we consider the following assumptions. Assumption 3. We assume that there exist C, c > 0 and m0 ≥ 0 such that for , defined on Hm0, for which any 0 < ρ < 1 there exists a set of functionals A(ρ) =nA(ρ) H ≤ C Xk∈K(cid:12)(cid:12)(cid:12)A(ρ) k (f )(cid:12)(cid:12)(cid:12) cXk (cid:12)(cid:12)(cid:12)A(ρ) k (f )(cid:12)(cid:12)(cid:12) ≤ kfk2 for all f ∈ Hm, m > m0. (1.11) 2 k ok∈K H! , + ρ2mkLm/2fk2 2 This Abstract Sampling Theorem 6.5 is used to construct bandlimited frames in E (Theorem 6.6). By exploring relations between Interpolation and Approximation spaces we develop approximation theory by Paley-Wiener vectors in Theorems 7.4 and 7.5. Finally, in Theorem 7.6 we obtain description of Besov spaces in terms of frame coefficients. In section 2 we show that our Assumptions 1 are satisfied for strongly continuous representations of Lie groups in Banach spaces and Assumptions 2 are satisfied for unitary representations of Lie groups in Hilbert spaces. Concerning Assumptions 3 we note that our results in [26]-[33] imply that at least when one is considering a so-called regular or quasi-regular representation of a Lie group in a function space on a homogeneous manifold M , some specific sets of Dirac measures (or even more general functionals [29]) "uniformly" distributed over M can serve as functionals , where ρ ∈ R+ represents a specific "spacing" of these Dirac A(ρ) =nA(ρ) measures. k ok∈Kρ It should be noted that due to importance of the theory of function spaces there is constant interest in extending classical constructions and results from Euclidean to non-Euclidean settings. It is impossible to list even the most significant publications on this subject which appeared during the last years. Here we mention just a very few papers which are relevant to our work [5], [15], [17]. 2. Lie groups and their representations 2.1. Lie groups and their representations in Banach spaces. Lie algebra g of a Lie group G can be identified with the tangent space Te(G) of G at the identity e ∈ G. Let exp(tX) : Te(G) → G, t ∈ R, X ∈ Te(G), be the exponential geodesic map i. e. exp(tX) = γ(1), where γ(t) is a geodesic of a fixed left-invariant metric on G which starts at e with the initial vector tX ∈ Te(G): γ(0) = e, dγ(0) dt = tX. It is known that exp is an analytic homomorphism of R onto one parameter subgroup exp tX of G: exp ((s + t)X) = exp(sX) exp(tX), s, t ∈ R. Let X1, ..., Xd, d = dimG form a basis in the Lie algebra of G, then one can consider the following coordinate system in a neighborhood of identity e (2.1) (t1, ..., td) 7→ exp(t1X1 + ... + tdXd). SOBOLEV, BESOV AND PALEY-WIENER VECTORS 5 If Y1 = s1X1 + ... + sdXd and Y2 = t1X1 + ... + tdXd then exp Y1 exp Y2 = exp Z, where Z is given by the Campbell-Hausdorff formula (2.2) Z = Y1 +Y2 + [Y1, Y2]+ [Y2, [Y1, [Y1, Y2]]]+... . 1 2 1 12 [Y1, [Y1, Y2]]− [Y2, [Y1, Y2]]− 1 12 1 24 It implies that Z = ζ1X1 + ... + ζdXd, where (2.3) ζj = sj + tj + O(ǫ2), tj,sj ≤ ǫ, 1 ≤ j ≤ d. One can also consider another local coordinate system around e which is given by the formula (2.4) (t1, ..., td) 7→ ϕ dXj=1 tjXj = exp(t1X1)... exp(tdXd). ′ ′ ′ E ≤ kfk E = supg∈G kT (g)fkE, f ∈ E, in which kT (g)fk Let us remind that a strongly continuous representation of a Lie group G in a Banach space E is a homomorphism g 7→ T (g), g ∈ G, T (g) ∈ GL(E), of G into the group GL(E) of linear bounded invertible operators in E such that trajectory T (g)f, g ∈ G, f ∈ E, is continuous with respect to g for every f ∈ E. We will consider only uniformly bounded representations. In this case one can introduce a new norm kfk E. Thus, without any restriction we will assume that the last inequality is satisfied in the original norm k · kE. Let X1, ..., Xd be a basis in g. With every Xj, 1 ≤ j ≤ d, one associates a strongly continuous one-parameter group of isometries t 7→ T (exp tXj), t ∈ R, whose generator is denoted as Dj, 1 ≤ j ≤ d. Lemma 2.1. If T : G 7→ GL(E) is a strongly continuous bounded representation of G in a Banach space E and Tj(t) = T (exp tXj), where {X1, ..., Xd} is a basis in g then the Assumption 1 is satisfied for groups Tj and their infinitesimal operators Dj, 1 ≤ j ≤ d. Proof. The fact that E1 =Td j=1 D(Dj) is dense in E and invariant with respect to T is well known [13], [14]. Since exp and ϕ are diffeomorphisms in a neighborhood of zero in g the map exp−1 ◦ϕ : g 7→ g is also a diffeomorphism. The formulas (2.2) and (2.3) give connection between (2.1) and (2.4) ϕ dXj=1 tjXj = exp dXk=1 αk(t)Xk! , t = (t1, , , td), (2.5) where (2.6) αj (t) = tj + O(ǫ2), tj ≤ ǫ, 1 ≤ j ≤ d. In particular, (2.2) implies exp τ Xj expPd (t1, ..., td), where i=1 tiXi = expPd k=1 γj k(t, τ )Xk, t = (2.7) γj k(t, τ ) = tk + τ ζj k(t) + τ 2Rj k(t, τ ), k(t) = δj and ζj are convergent series in t1, ..., td and t1, ..., td, τ respectively. k is the Kronecker symbol and Qj k(t), where δj k + Qj k(t) and Rj k(t, τ ) 6 SOBOLEV, BESOV AND PALEY-WIENER VECTORS f ∈ E1 the following Since for f ∈ E1 one has DjT (g)f = d dτ T (exp τ Xj) T (g)fτ =0 we obtain for DjT1(t1)...Td(td)f = DjT ϕ dXi=1 T (exp τ Xj) T ϕ dXi=1 tiXi!! fτ =0 = tiXi!! = DjT exp T exp dXi=1 d dτ d dτ αi(t)Xi! f = dXi=1 i (α, τ )Xi! fτ =0, γj where α = (α1(t), ..., αd(t)) and according to (2.7) γj τ 2Rj i (α(t)) + i (α, τ ). By using the Chain Rule and (2.5) we finally obtain the formula (1.3) i (α, τ ) = αi(t) + τ ζj DjT1(t1)...Td(td)f = d dτ dXk=1(cid:18) d dτ k(α, τ )τ =0(cid:19) ∂kT exp γj dXi=1 Lemma is proved. i (α, τ )Xi! fτ =0 = γj dXi=1 T exp i (α, τ )Xi! fτ =0 = γj dXk=1 ζj k(t)∂kT1(t1)...Td(td)f. (cid:3) 2.2. Unitary representations in Hilbert spaces. A strongly continuous uni- tary representation of a Lie group G in a Hilbert space H is a homomorphism T : G 7→ U (H) where U (E) is the group of unitary operators of H such that T (g)f, g ∈ G, is continuous on G for any f ∈ H. The Garding space G is defined as f ∈ H, ϕ ∈ C∞0 (G), dg is a left-invariant measure on G. If X ∈ g is identified with a right-invariant vector field the set of vectors h in H that have the representation h =RG ϕ(g)T (g)f dg, where Xϕ(g) = lim t→0 ϕ (exp tX · g) − ϕ(g) t , D(X)h = −RG Xϕ(g)T (g)f dg. It is known that G ⊂Tr∈N Hr = H∞ is invariant a basis in g and Di = D(Xi), 1 ≤ i ≤ d, we consider the operator LG = −Pd Since LG is symmetric and the differential operator −Pd then one has a representation D(X) of g by operators which act on G by the formula with respect to all operators D(X), X ∈ g, and dense in every Hr. If X1, ..., Xd is i=1 D2 i defined on G. i is elliptic on the group G the Theorem 2.2 in [14] implies that LG is essentially self-adjoint, which means LG = L∗ . In other words, the closure LG = L of LG from G is a self-adjoint G operator. Obviously, L ≥ 0. We introduce the self-adjoint operator Λ = I + L ≥ 0. Theorem 2.2. The space Hr with the norm (1.1) is isomorphic to the domain of Λr/2 with the norm kΛr/2fkH. Proof. In the case r = 2k, the inequality i=1 X 2 (2.8) kfkH2k ≤ C(k)kΛkfkH is shown in [13], Lemma 6.3. The reverse inequality is obvious. We consider now the case r = 2k + 1. If f ∈ H2 = D(Λ), then since D(Λ) ⊂ D(Λ1/2) we have H +Xj kfk2 kDjfk2 H = hf, fi +Xj hDjf, Djfi = hf, fi +*−Xj j f, f+ = D2 SOBOLEV, BESOV AND PALEY-WIENER VECTORS 7 (2.9) *f −Xj j f, f+ = hΛf, fi = kΛ1/2fk2 H. D2 j1 ...D2 . j1 ...D2 jmf. Thus if f ∈ H4k+2 then These equalities imply that H1 is isomorphic to D(Λ1/2). Our goal is to to prove existence of an isomorphism between H2k+1 and D(Λk+1/2). It is enough to es- tablish equivalence of the corresponding norms on the set H4k+2 = D(Λ2k+1) since the latest is dense in H2k+1. If f ∈ H4k+2 ⊂ H2k then Djf ∈ H4k+1 ⊂ H2k and is a polynomial in D1, ..., Dd whose degree ≤ 2k. According to (2.8) and (2.2) we have that Λkf =Pm≤kP D2 jmDj2k+1 f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H (cid:13)(cid:13)Dj1 ...Dj2k+1 f(cid:13)(cid:13)H ≤ C(cid:13)(cid:13)ΛkDj2k+1 f(cid:13)(cid:13)H =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xm≤kX D2 Multiple applications of the identity DiDj − DjDi =Pk ck lead to the inequality(cid:13)(cid:13)Dj1 ...Dj2k+1 f(cid:13)(cid:13)H ≤ C(cid:0)kDj2k+1ΛkfkH + kRfkH(cid:1) , where R (cid:13)(cid:13)Dj2k+1 Λkf(cid:13)(cid:13)H ≤(cid:13)(cid:13)(cid:13)Λ1/2Λkf(cid:13)(cid:13)(cid:13)H =(cid:13)(cid:13)(cid:13)Λk+1/2f(cid:13)(cid:13)(cid:13)H and also kRfkH ≤ kfkH2k ≤ C(k)(cid:13)(cid:13)Λkf(cid:13)(cid:13)H . Since kΛkfkH is not decreasing with equality Λkf =Pm≤kP D2 Now, since for f ∈ H4k+2 we have Dj1...Dj2k f ∈ H2k+2 ⊂ H1 = D(Λ1/2), and the kDj1...Dj2k+1 fkH ≤ C(k)kΛk+1/2fkH, f ∈ H4k+2. j1 ...D2 jmf, holds we obtain, by using (2.2) i,jDk which holds on H2 k we get the following estimate j1...D2 jm fkH ≤ CkfkH2k+1, C = C(k). kΛk+1/2fkH = kΛ1/2Xm≤kX D2 Theorem is proved. (cid:3) Corollary 2.1. If T is a strongly continuous unitary representation of a Lie group in a Hilbert space H and X1, ..., Xd is a basis in the corresponding algebra Lie g then for Tj(t) = T (exp tXj), 1 ≤ j ≤ d, and their generators Dj, 1 ≤ j ≤ d, the Assumption 2 is satisfied. 3. Hardy-Steklov operator associated with operators D1, D2, ..., Dd. We return to the general set up. Namely, we consider a Banach space E and op- erators D1, D2, ..., Dd which generate strongly continuous uniformly bounded semi- groups T1(t), T2(t), ..., Td(t), kT (t)k ≤ 1, t ≥ 0. It is assumed that the Assumption 1 holds. 3.1. Preliminaries. If D generates in E a strongly continuous bounded semigroup TD(t), t ≥ 0, and Ωr(s, f ) = sup 0≤τ≤sk (TD(τ ) − I)r fkE, where I is the identity operator, then one can easily prove the following two in- equalities (3.1) and (3.2) Ωm (f, s) ≤ skΩm−k(Dkf, s), Ωm (f, as) ≤ (1 + a)m Ωm(f, s), a ∈ R. SOBOLEV, BESOV AND PALEY-WIENER VECTORS Let F (x1, x2, ..., xN ) be a function on RN that takes values in the Banach space 8 E For 1 ≤ i ≤ N we introduce difference operator by the formula (∆i(s)F )(x1, x2, ..., xN ) = F (x1, x2, ..., xi−1, s, xi+1, ..., xN )− F : RN 7−→ E. (3.3) For a scalar differentiable function θ : RN 7−→ R and for 1 ≤ i1, ..., il ≤ N, im 6= ik, the following inequality holds F (x1, x2, ..., xi−1, 0, xi+1, ..., xN ). max 0≤xj≤sj ∆i1 (si1 )...∆il (sil )θ(x1, ..., xN ) ≤ si1 ...sil max One has for 1 ≤ i ≤ N ∂k ∂i1 ...∂ik 0≤xj≤sj(cid:12)(cid:12)(cid:12)(cid:12) θ(x1, ..., xN )(cid:12)(cid:12)(cid:12)(cid:12) . ∆i(s)(θF )(x1, ..., xN ) = ∆i(s)θ(x1, ..., xN )F (x1, ..., xi−1, s, xi+1, ..., xN )+ (3.4) and then if im 6= ik one has θ(x1, ..., xi−1, 0, xi+1, ..., xN )∆i(s)F (x1, ..., xN ), ∆i1 (si1 )...∆il (sil )(θF )(x1, ..., xN ) = (3.5)X ∆i(1) 1 (si(1) 1 )...∆i(1) l (si(1) l )θ(x(0) 1 ,...,i(2) i(2) l )∆i(2) 1 (si(2) 1 )...∆i(2) l (si(2) l )F (x(s) 1 ,...,i(1) i(1) l ), where the sum is taken over all possible partitions of the natural vector (i1, ..., il) in the sum of nonnegative integer vectors (i(1) ). The x(0) denotes a vector obtained from the vector (x1, ..., xN ) by replacing the 1 ,...,i(2) i(2) coordinates xi(2) 1 , ..., i(1) i(1) If F is a differentiable function and m 6= i then we have by 0 and x(s) 1 ,...,i(1) i(1) l , respectively. , denotes the vector obtained by replacing ) and (i(2) 1 , ..., i(1) 1 , ..., i(2) , ..., si(1) by si(1) 1 ,...,i(2) l l l 1 l l l (3.6) ∆i(s) F (x1, x2, ..., xN ) = ∂ ∂xm ∂ ∂xm ∆i(s)F (x1, x2, ..., xN ), m 6= i. In particular, for any r ≥ 2 for τj = τj,1 + ... + τj,r, j = 1, ..., d, 1 ≤ i, i k, k , k ′ ′ ′ ′ ≤ d, 1 ≤ ) 6= (i, k) we have ∆i,k(s) Tj (τj) f = ∆i,k(s) ∂ ∂τi′ ,k′ Tj (τj) f, f ∈ E1, ∂ ∂τi′ ∆i,k(s) Tj (τj) f = ∆i,k(s) ∂ ∂τi′ Tj (τj) f, f ∈ E1, where the right-hand side and ∆i,k(s)Qd j=1 Tj (τj ) f are defined according to (3.1). In the same notations τj = τj,1 + ... + τj,r, 1 ≤ m, i ≤ d, 1 ≤ k ≤ r, we clearly have the identity Dm∆i,k(s) dYj=1 Tj (τj) f = ∆i,k(s)Dm dYj=1 Tj (τj) f, f ∈ E1, ≤ r, if (i ∂ ∂τi′ ,k′ or, equvalently, dYj=1 dYj=1 dYj=1 dYj=1 SOBOLEV, BESOV AND PALEY-WIENER VECTORS 9 which implies for every f ∈ E1 the next formula (3.7) Dm∆i1,k1(s)..∆il ,kl(s) Tj (τj) f = ∆i1,k1(s)..∆il,kl (s)Dm Tj (τj) f. dYj=1 dYj=1 Using (3.7), (1.3), (3.1) and (3.6) we obtain for f ∈ E1, τj = τj,1 + ... + τj,r, j = 1, ..., d, r ≥ l + 1, 1 ≤ m ≤ d, Tj (τj) f = dYj=1 Dm∆i1,k1 (s)...∆il,kl(s) (3.8) 1 ∆i(1) 1 ,k(1) (s)...∆i(1) X dXil+1=1 where outer summationP and arguments of ζil+1 as in (3.1). (s)ζil+1 m (...) ∂τil+1 ,k(1) ∂ l l ∆i(2) 1 ,k(2) 1 (s)...∆i(2) l ,k(2) l (s) Tj (...) f, dYj=1 m (...) andQd j=1 Tj (...) f the same 3.2. The Hardy-Steklov operator. We introduce a generalization of the classi- cal Hardy-Steklov operator. For a positive small s, natural r and 1 ≤ j ≤ d we set ...Z s/r 0 rXk=1 (−1)kCk r Tj(k(τj,1 + ... + τj,r)f dτj,1...dτj,r, Hj.r(s)f = (s/r)−rZ s/r Hr(s)f =Qd where Ck 0 r are the binomial coefficients and then define the Hardy-Steklov operator: j=1 Hj,r(s)f = H1,r(s)H2,r(s)...Hd,r(s)f, f ∈ E. For every fixed f ∈ E the function Hr(s)f is an abstract valued function from R to E and it is a linear combination of some abstract valued functions of the form (3.9) T f (τ )dτ1,1...dτd,r, (s/r)−rdZ s/r 0 ...Z s/r 0 where τ = (k1τ1, k2τ2, ..., kdτd), 1 ≤ kj ≤ r, τj = (τj,1 + τj,2 + ... + τj,r), 1 ≤ j ≤ d, and (3.10) T f (τ ) = T1(k1τ1)T2(k2τ2)... Td(kdτd)f. Lemma 3.1. The following holds: (1) For every f ∈ E the function Hr(s)f maps R to Er. (2) For every 0 ≤ q ≤ r the "mixed derivative" Dj1 ...Djq Hr(s)f, 1 ≤ jk ≤ d is another abstract valued function with values in Er−q and it is a linear combination of abstract valued functions (with values in Er) of the form (3.11) (3.12) 0 0 .....Z s/r {z } (s/r)−rdZ s/r 0≤τi,j≤s∂pµi1,...,il;m rd−m j1,...,jq max where (·) ≤ csm−l, p ∈ N ∪ {0}, µi1,...,il;m j1,...,jq (·)∆i1,k1(s/r)....∆il ,kl(s/r)T f (·)d·, 10 SOBOLEV, BESOV AND PALEY-WIENER VECTORS and 0 < s < 1, 0 ≤ m ≤ rd, 0 ≤ l ≤ m, where l = 0 corresponds to the case when the set of indices {i1, ..., il} is empty. Proof. The proof proceeds by induction on q. We shall show that Hr(s)f is in E1 for every f ∈ E. Let's assume first that f ∈ E1. Since every Dj, 1 ≤ j ≤ d, is a closed operator to show that the term (3.9) belongs to D(Dj) it is sufficient to show existence of the integral (see notations (5.1), (5.2)) According to (1.3) the last integral equals to DjT f (τ )dτ1,1...dτd,r. (3.13) (3.14) dXi=1(cid:16) s ζi j (τ )∂iT f (k1τ1, ..., kdτd)dτ1,1...dτd,r, where τ and τj described in (5.1) and (5.2). Since derivative ∂i is the same as the the integration by parts formula and (3.1) allow to continue (3.14) derivative as follows ∂τi1,1 ∂ 0 0 0 0 dXi=1 ...Z s/r ...Z s/r r(cid:17)−rdZ s/r (cid:16) s r(cid:17)−rdZ s/r  ...Z s/r Z s/r {z }  ...Z s/r Z s/r {z } Z s/r ...Z s/r dXi=1 } {z dXi=1 rd−1 rd−1 rd 0 0 0 0 0 (s/r)−rd j(τ (s/r) ζi i (s/r)−rd (∆i(s/r)ζi + )∆i(s/r)T f (τ )(dτ )i )(dτ )i − j)(τ )T f (τ )dτ = Bj(s)f, j (τ ))T f (τ (0) i (3.15) (s/r)−rd (∂iζi 0 the term dτi,r is missing. where τ = (τ1,1 + ... + τ1,r, ... , τn,1 + ... + τn,r), τ s/r τi,r−1 +s/r, ..., τn,1 +...+τn,r), τ 0 i = (τ1,1 + ... + τ1,r, ..., τi,1 + ... + i = (τ1,1 +...+τ1,r, ..., τi,1 +...+τi,r−1 +0, ..., τn,1 + Thus if f belongs to E1 then Hr(s)f takes values in E1 and the formula DjHr(s)f = ... + τn,r), dτ = dτ1,1....dτn,r, and (dτ )i = dτ1,1....dτi,r−1ddτi,rdτi+1,1...dτn,r, where Since integrand of each of the three integrals is bounded it implies existence of (3.13) which, in turn, shows that (3.9) is an element of D(Dj) for every 1 ≤ j ≤ d. Bj(s)f holds where the operator Bm(s) is bounded. This fact along with the fact that E1 is dense in E implies the formula DjHr(s)f = Bj(s)f for every f ∈ E. Thus we proved the first part of the Lemma for q = 1. To verify the second claim of the Lemma it is sufficient to note that in the first j , in the second line m = 1, l = j1 = ∂iζi j. line of (3.2) one has m = l = q = 1, j = j1, µ1,1 0, q = 1, j = j1, µ0,1 It is easy to verify that in all of these cases the estimates (3.12) hold. j , in the third line m = l = 0, q = 1, µ0,0 j1 = ∆i(s/r)ζi j1 = ζi We now assume that theorem is proved for all q < r and we shall prove it for q + 1 ≤ r. Suppose f ∈ E1. We consider SOBOLEV, BESOV AND PALEY-WIENER VECTORS 11 (3.16) Z s/r 0 .....Z s/r 0 µi1,...,il;m j1,...,jq (·)Djq+1 ∆i1,k1(s/r)....∆il ,kl(s/r)T (·)f d · . By (3.1) this expression is decomposed into a sum of terms of the form (3.17) Z s/r 0 ...Z s/r νi1,...,il+1;m j1,...,jq+1 (·)∂il+1 ∆i(2) 1 ,k(2) 1 (s/r)...∆i(2) l (s/r)T (τ )f dτ, ,k(2) l 0 are constructed from µi1,...,il;m j1,...,jq+1 where νi1,...,il+1;m In- tegrating by parts as in the first step of induction, we obtain that the integral (3.17) exists and is equal to the value of some bounded operator at an element f ∈ E1. As a result we obtain that Hr(s) belongs to the space Eq+1 and also that Dj1...Djq+1 Hr(s)f is a linear combination of terms of the form (3.11) . The rest of the theorem fellows from the induction hypothesis. Theorem is proved. (cid:3) in accordance with (3.1). j1,...,jq 4. Interpolation spaces The goal of the section is to introduce basic notions of the theory of interpola- tion spaces [1], [3], [10]. Later in section 7 we will also introduce the so-called approximation spaces [18], [4]. It is important to realize that the relations be- tween interpolation and approximation spaces cannot be described in the language of normed spaces. We have to make use of quasi-normed linear spaces in order to treat them simultaneously. A quasi-norm k·kE on linear space E is a real-valued function on E such that for any f, f1, f2 ∈ E the following holds true: (1)kfkE ≥ 0; (2) kfkE = 0 ⇐⇒ f = 0; (3)k − fkE = kfkE; (4) there exists some CE ≥ 1 such that kf1 + f2kE ≤ CE(kf1kE + kf2kE). Two quasi-normed linear spaces E and F form a pair if they are linear subspaces of a common linear space A and the conditions kfk − gkE → 0, and kfk − hkF → 0 imply equality g = h (in A). For any such pair E, F one can construct the space E ∩ F with quasi-norm kfkE∩F = max (kfkE,kfkF) and the sum of the spaces, E + F consisting of all sums f0 + f1 with f0 ∈ E, f1 ∈ F, and endowed with the quasi-norm kfkE+F = inf f =f0+f1,f0∈E,f1∈F (kf0kE + kf1kF) . Quasi-normed spaces H with E ∩ F ⊂ H ⊂ E + F are called intermediate If both E and F are complete the inclusion mappings are between E and F. automatically continuous. An additive homomorphism T : E → F is called bounded if kTk = supf∈E,f6=0 kT fkF/kfkE < ∞. An intermediate quasi-normed linear space H interpolates between E and F if every bounded homomorphism T : E+F → E+F which is a bounded homomorphism of E into E and a bounded homomorphism of F into F is also a bounded homomorphism of H into H. On E + F one considers the so-called Peetre's K-functional K(f, t) = K(f, t, E, F) = inf The quasi-normed linear space (E, F)K or 0 ≤ θ ≤ 1, q = ∞, is introduced as the set of elements f in E + F for which f =f0+f1,f0∈E,f1∈F θ,q, with parameters 0 < θ < 1, 0 < q ≤ ∞, (kf0kE + tkf1kF) . (4.1) kfkθ,q =(cid:18)Z ∞ 0 (cid:0)t−θK(f, t)(cid:1)q dt t (cid:19)1/q < ∞. SOBOLEV, BESOV AND PALEY-WIENER VECTORS It turns out that (E, F)K θ,q with the quasi-norm (4.1) interpolates between E and 12 F. 5. Approximation by Hardy-Steklov averages, K-functor and modulus of continuity We are going to prove items (1) and (2) of Theorem 1.3. Proof. First we prove the right-hand side of the inequality (1.7). The following simple lemma plays an important role in the roof. Lemma 5.1. In any ring R with multiplicative identity 1 the following formulas hold for a1, a2, ...., an ∈ R, (5.1) a1a2...an − 1 = a1(a2 − 1) + ... + a1a2...an−1(an − 1), (5.2) (a1 − 1)a2...an = (a1 − 1) + a1(a2 − 1) + ... + a1a2...an−1(an − 1). By using the first formula we obtain (cid:13)(cid:13)(cid:13)(−1)n(r+1)Hr(s)f − f(cid:13)(cid:13)(cid:13)E =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nYj=1 (−1)n+1Hj,r(s)f − f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)E ≤ (5.3) c sup 0≤τj≤s/r k(Tj(τj) − I)rfkE ≤ CΩr(s, f ). nXj=1 sup where T (·) = Qn j=1Qr s−l To estimate srkHr(s)fkE we note that kHr(s)fkE ≤ CkfkE. According to Lemma 3.1 the quantity kDj1 ...Djr Hr(s)fkE is estimated for 0 ≤ s ≤ 1 by (5.4) 0≤τj,k≤sk∆j1,k1(s/r)...∆jl ,kl(s/r)T (·)fkE, k=1 Tj(τj,k). By the definition of ∆j,k(s/r) the expression ∆j,k(s/r)T (·) differs from T (·) only in that in place of the factor Tj(τj,k) the factor Tj(s/r) − I appears. Multiple applications of the identity (5.2) to the operator ∆j1,k1(s/r)...∆jl ,kl(s/r)T (·) allow its expansion into a sum of operators each of which is a product of not less than l ≤ r of operators Ti(σi)− I, σi ∈ (0, s/r), 1 ≤ i ≤ n. Consequently, (5.4) is dominated by a multiple of s−lΩl(s, f ). By summing the estimates obtained above we arrive at the inequality (5.5) K(sr, f, E, Er) ≤ C nXl=1 sr−lΩl(s, f ) + srkfkE! , 0 ≤ s ≤ 1. Note, that by repeating the known proof for the classical modulus of continuity one can prove the inequality Ωl(s, f ) ≤ C(cid:18)slkfkE + slZ 1 s σ−1−lΩk+r(σ, f )dσ(cid:19) , which implies sr−lΩl(s, f ) ≤ C (srkfkE + Ωr(s, f )) . By applying this inequality to (5.5) and taking into account the inequality K(sr, f, E, Er) ≤ kfk, we obtain the right-hand side of the estimate (1.7). To prove the left-hand side of (1.7) we first notice that the following inequality holds (5.6) Ωr(s, g) ≤ Csk X1≤j1,...jk≤n Ωr−k (s, Dj1...Djk g) , g ∈ Ek, C = C(k, r), k ≤ r, SOBOLEV, BESOV AND PALEY-WIENER VECTORS 13 which is an easy consequence of the identity (5.2) and the identity (Tj(t) − I) g = 0 Tj(τ )Djgdτ, g ∈ D(Dj). From here, for any f ∈ E, g ∈ Er we obtain Ωr(s, f ) ≤ Ωr(s, f − g) + Ωr(s, g) ≤ C (kf − gkE + srkgkEr). The first item of Theorem 1.3 is proven and it obviously implies second item of the same Theorem. R t (cid:3) Remark 5.2. The proof shows that the left-hand side of (1.7) holds true for any finite set of one-parameter strongly continuous bounded semigroups. Below is the proof of items (3)-(5) of Theorem 1.3. Proof. We will need the following lemma. Lemma 5.3. The following inequalities hold (5.7) (5.8) E kfkEk ≤ Ckfk1−k/r Er , f ∈ Er, C = C(k, r), K(sr, f, E, Er) ≤ CskkfkEk, f ∈ Ek, C = C(k, r). kfkk/r Proof. The first inequality follows from its well-known one-dimensional version (see also [33]). The second one follows from the right-hand estimate of (1.7) and (5.6). (cid:3) This lemma shows that one can use the Reiteration Theorem (see [3], [10]), which immediately implies item (3) of Theorem 1.3. Next, let α > 0, [α] be a non-integer and its integer part respectively. According to item (3) of The- orem 1.3 we have (E, Er)K α−[α],q = is a continuous map from (α−[α])/(r−[α]),q. All together it shows that (α−[α])/(r−[α]),q and (cid:0)E, E1(cid:1)K (α−[α])/(r−[α]),q. Note, that Dj1 Dj2 ...Dj[α] α/r,q = (cid:0)E[α], Er(cid:1)K (cid:0)E, Er−[α](cid:1)K (cid:0)E[α], Er(cid:1)K if f ∈ (E, Er)K (5.9) (α−[α])/(r−[α]),q to(cid:0)E, Er−[α](cid:1)K α/r,q then Dj1 Dj2...Dj[α] f ∈(cid:0)E, E1(cid:1)K (cid:13)(cid:13)Dj1Dj2 ...Dj[α] f(cid:13)(cid:13)(E,E1)K Conversely, let Dj1 Dj2...Dj[α] f ∈(cid:0)E, E1(cid:1)K the right-hand estimate of (1.7) and (5.6) imply α−[α],q and α−[α],q ≤ Ckfk(E,Er)K . α/r,q (α−[α])/(r−[α]),q . Then α−[α],q =(cid:0)E[α], Er(cid:1)K nXj1,...,j[α]=1(cid:13)(cid:13)Dj1 Dj2 ...Dj[α] f(cid:13)(cid:13)(E,E1)K . α−[α],q (5.10) kfk(E,Er)K α/r,q ≤ C Inequalities (5.9) and (5.10) imply item (4) of Theorem 1.3. Proof of item (5) is similar. Theorem 1.3 is completely proved. (cid:3) 6. Shannon sampling, Paley-Wiener frames and abstract Besov subspaces 6.1. Paley-Wiener vectors in Hilbert spaces. Consider a self-adjoint positive definite operator L in a Hilbert space H. Let √L be the positive square root of L. According to the spectral theory for such operators [2] there exists a direct integral of Hilbert spaces X = R X(λ)dm(λ) and a unitary operator F from H onto X, 14 SOBOLEV, BESOV AND PALEY-WIENER VECTORS which transforms the domains of Lk/2, k ∈ N, onto the sets Xk = {x ∈ Xλkx ∈ X} with the norm (6.1) kx(λ)kXk = hx(λ), x(λ)i1/2 X(λ) =(cid:18)Z ∞ 0 λ2kkx(λ)k2 X(λ)dm(λ)(cid:19)1/2 . and satisfies the identity F (Lk/2f )(λ) = λk(F f )(λ), if f belongs to the domain of Lk/2. We call the operator F the Spectral Fourier Transform [25], [26]. As known, X is the set of all m-measurable functions λ 7→ x(λ) ∈ X(λ), for which the following norm is finite: kxkX =(cid:18)Z ∞ 0 kx(λ)k2 X(λ)dm(λ)(cid:19)1/2 For a function F on [0,∞) which is bounded and measurable with respect to dm one can introduce the operator F (√L) by using the formula F (√L)f = F−1F (λ)F f, f ∈ H. (6.2) If F is real-valued the operator F (√L) is self-adjoint. Remark 6.1. In many applications L is a second-order differential operator and then √L is a first-order pseudo-differential operator. Definition 1. For √L as above we will say that a vector f ∈ H belongs to the Paley-Wiener space PWω(√L) if the support of the Spectral Fourier Transform F f is contained in [0, ω]. The next two facts are obvious. Theorem 6.2. The spaces PWω(√L) have the following properties: (1) the space PWω(√L) is a linear closed subspace in H. (2) the spaceSω>0 PWω(√L) is dense in H; Next we denote by Hk the domain of Lk/2. It is a Banach space, equipped with the graph norm kfkk = kfkH + kLk/2fkH. The next theorem contains gen- eralizations of several results from classical harmonic analysis (in particular the Paley-Wiener theorem). It follows from our results in [26]. Theorem 6.3. The following statements hold: (6.3) and the following Bernstein inequalities holds true kLs/2fkH ≤ ωskfkH for all s ∈ R+; (1) (Bernstein inequality) f ∈ PWω(√L) if and only if f ∈ H∞ =T∞k=1 Hk, (2) (Paley-Wiener theorem) f ∈ PWω(√L) if and only if for every g ∈ H the scalar-valued function of the real variable t 7→ heit√Lf, gi is bounded on the (3) (Riesz-Boas interpolation formula) f ∈ PWω(√L) if and only if f ∈ H∞ real line and has an extension to the complex plane as an entire function of the exponential type ω; and the following Riesz-Boas interpolation formula holds for all ω > 0: (6.4) i√Lf = ω π2Xk∈Z (−1)k−1 (k − 1/2)2 ei( π ω (k−1/2))√Lf. SOBOLEV, BESOV AND PALEY-WIENER VECTORS 15 Proof. (1) follows immediately from the definition and representation (6.1). To prove (2) it is sufficient to apply the classical Bernstein inequality [16] in the uniform norm on R to every function heit√Lf, gi, g ∈ H. To prove (3) one has to apply the classical Riesz-Boas interpolation formula on R, [36], [16] to a function heit√Lf, gi. (cid:3) 6.2. Frames in Hilbert spaces. A family of vectors {θv} in a Hilbert space H is called a frame if there exist constants A, B > 0 such that (6.5) hf, θvi2 ≤ Bkfk2 H for all f ∈ H. Akfk2 H ≤Xv The largest A and smallest B are called lower and upper frame bounds. is f =Pv hf, θvi Θv. Dual frames are not unique in general. Moreover it may be The family of scalars {hf, θvi} represents a set of measurements of a vector f . In order to resynthesize the vector f from this collection of measurements in a linear way one has to find another (dual) frame {Θv}. Then a reconstruction formula difficult to find a dual frame in concrete situations. If A = B = 1 the frame is said to be tight or Parseval. Parseval frames are similar in many respects to orthonormal wavelet bases. For example, if in addition all vectors θv are unit vectors, then the frame is an orthonormal basis. The main feature of Parseval frames is that decomposing and synthesizing a vector from known data are tasks carried out with the same family of functions, i.e., the Parseval frame is its own dual frame. 6.3. Sampling in abstract Paley-Wiener spaces. We now assume that the Assumption 3 is satisfied. It meant that there exists a C > 0 and m0 ≥ 0 such that for any 0 < ρ < 1 there exists a set of functionals A(ρ) =nA(ρ) k o , defined on Hm0, for which the inequalties (1.11) hold true. Remark 6.4. Following [27], [29] we call inequality (1.11) a Poincar´e-type inequal- ity since it is an estimate of the norm of f through the norm of its "derivative" Lm/2f . Let us introduce vectors µk ∈ H such that hf, µki = A(ρ) Let Pω be the orthogonal projection of H onto PWω(√L) and put k (f ), f ∈ Hm, m > m0. (6.6) φω k = Pωµk. Using the Bernstein inequality (6.3) we obtain the following statement. Theorem 6.5. (Sampling Theorem) Assume that inequality (1.11) holds and for a given ω > 0 and δ ∈ (0, 1) pick a ρ such that (6.7) Then the family of vectors {φω (6.8) k} is a frame for the Hilbert space PWω(√L) and ρ2m = C−1ω−2mδ. (1 − δ)kfk2 hf, φω ki2 ≤ kfk2 H ≤Xk H, f ∈ PWω(√L). k ∈ PWω(√L) and provides the k , f ∈ PWω(√L). The canonical dual frame {Θω following reconstruction formulas ki Θω hf, φω (6.9) f =Xk k =Xk k} has the property Θω hf, Θω ki φω 16 SOBOLEV, BESOV AND PALEY-WIENER VECTORS 6.4. Partitions of unity on the frequency side. The construction of frequency- localized frames is achieved via spectral calculus. The idea is to start from a partition of unity on the positive real axis. In the following, we will be considering two different types of such partitions, whose construction we now describe in some detail. j=0 F 2 Let g ∈ C∞(R+) be a non-increasing function such that supp(g) ⊂ [0, 2], and g(λ) = 1 for λ ∈ [0, 1], 0 ≤ g(λ) ≤ 1, λ > 0. We now let h(λ) = g(λ) − g(2λ) , which entails supp(h) ⊂ [2−1, 2], and use this to define F0(λ) =pg(λ) , Fj(λ) = ph(2−jλ) , j ≥ 1 , as well as Gj(λ) = [Fj(λ)]2 = F 2 the definitions, we get for all λ ≥ 0 the equations Pn g(2−nλ), and as a consequence Pj≥0 Gj (λ) = Pj≥0 F 2 j (λ) , j ≥ 0 . As a result of j (λ) = j (λ) = 1 , λ ≥ 0, with finitely many nonzero terms occurring in the sums for each fixed λ. We call the sequence (Gj )j≥0 a (dyadic) partition of unity, and (Fj)j≥0 a quadratic (dyadic) partition of unity. As will become soon apparent, quadratic parti- tions are useful for the construction of frames. Using the spectral theorem one has F 2 j=0 Gj(λ) = Pn j (λ)F f (λ)(cid:1) , f = F−1F f (λ) = F−1Xj≥0 j (√L)f, fi =Pj≥0 kFj(√L)fk2 Taking inner product with f gives kFj(√L)fk2 Pj≥0hF 2 elements Fj (√L)f and Gj (√L)f are bandlimited to [2j−1, 2j+1], whenever j ≥ 1, f . Moreover, since the functions Gj , Fj, have their supports in [2j−1, 2j+1], the j (λ)F f (λ) =Xj≥0 H. Similarly, we get the identityPj≥0 Gj(√L)f = j (√L)f, fi and kfk2 j (√L)f = F−1(cid:0)F 2 j ≥ 1, and thus j (√L)f F 2 H = hF 2 (6.10) H = F 2 and to [0, 2] for j = 0. 6.5. Paley-Wiener frames in Hilbert spaces. Using the notation from above and Theorem 6.5, one can describe the following Paley-Wiener frame in an abstract Hilbert space H. Theorem 6.6. (Paley-Wiener nearly Parseval frame in H) For a fixed δ ∈ (0, 1) and j ∈ N let {φj k} be a set of vectors described in Theorem 6.5 that correspond to ω = 2j+1. Then for functions Fj the family of Paley-Wiener vectors Φj k has the following properties: k = Fj(√L)φj (1) Each vector Φj (2) The family nΦj (1 − δ)kfk2 (6.11) k belongs to PW[2j−1, 2j+1](√L), j ∈ N, k = 1, .... ko is a frame in H with constants 1 − δ and 1: H ≤Xj≥0Xk (cid:12)(cid:12)(cid:12)Df, Φj kE(cid:12)(cid:12)(cid:12) H, f ∈ H. ≤ kfk2 2 (3) The canonical dual frame {Ψj k} also consists of bandlimited vectors Ψj k ∈ PW[2j−1, 2j+1](√L), j ∈ [0, ∞), k = 1, ..., and has frame bounds A = 1, B = (1 − δ)−1. f =PjPkDf, Φj (4) The reconstruction formulas hold for every f ∈ H k. kE Φj k =PjPkDf, Ψj The last two items here follow from the first two and general properties of frames. We also note that for reconstruction of a Paley-Wiener vector from a set of samples kE Ψj SOBOLEV, BESOV AND PALEY-WIENER VECTORS 17 one can use, besides dual frames, the variational (polyharmonic) splines in Hilbert spaces developed in [27]. 7. Besov subspaces in Hilbert spaces 7.1. Approximation spaces. Let us introduce another functional on E + F, where E and F form a pair of quasi-normed linear spaces E(f, t) = E(f, t, E, F) = inf g∈F,kgkF≤t kf − gkE. Definition 2. The approximation space Eα,q(E, F), 0 < α < ∞, 0 < q ≤ ∞ is the quasi-normed linear spaces of all f ∈ E + F for which the quasi-norm (7.1) is finite. kfkEα,q(E,F) =(cid:18)Z ∞ 0 (tαE(f, t))q dt t (cid:19)1/q The next theorem represents a very abstract version of what is known as an Equivalence Approximation Theorem [18], [2]. In the form it is stated below it was proved in [9]. Theorem 7.1. Suppose that T ⊂ F ⊂ E are quasi-normed linear spaces and E and F are complete. If there exist C > 0 and β > 0 such that the following Jackson-type inequality is satisfied tβE(t, f,T , E) ≤ CkfkF, t > 0, f ∈ F, then the following embedding holds true (7.2) (E, F)K θ,q ⊂ Eθβ,q(E,T ), 0 < θ < 1, 0 < q ≤ ∞. If there exist C > 0 and β > 0 such that the following Bernstein-type inequality holds kfkF ≤ Ckfkβ (7.3) T kfkE, f ∈ T , then the following embedding holds true Eθβ,q(E,T ) ⊂ (E, F)K θ,q, 0 < θ < 1, 0 < q ≤ ∞. H,q(√L) = ral abelian group as the additive group of a vector space, with the quasi-norm 7.2. Besov subspaces in Hilbert spaces. According to (1.6) we introduce Bα (H, Hr)K θ,q, 0 < θ = α/r < 1, 1 ≤ q ≤ ∞. We also introduce a notion of best approximation: E(f, ω) = inf g∈PWω (√L) kf − gkH. Our goal is to apply Theorem 7.1 in the situation where E = H, F = Hr and T = PWω(√L) is a natu- kfkT = infnω′ > 0 : f ∈ PWω′(cid:16)√L(cid:17)o . To be more precise it is the space of finite sequences of Fourier coefficients c = (c1, ...cm) ∈ PWω(√L) where m is the greatest index such that the eigenvalue λm ≤ ω. For a c = (c1, ...cm) ∈ PWω(√L) the quasi-norm is defined as kckEω(L) = max(cid:8)pλj : cj 6= 0, cj+1 = ... = cm = 0(cid:9) . normed spaces is because k·kT is clearly not a norm, only a quasi-norm on PWω(√L). inequality for bandlimited functions in f ∈ PWω(√L). One can prove the following statement (see [22], [26]). Lemma 7.3. A vector f belongs to the space PWω(√L) if and only if the following Bernstein inequality holds kLr/2fkH ≤ ωrkfkH, r ∈ R+. Remark 7.2. Let us emphasize that the reason we need the language of quasi- The Plancherel Theorem allows us to verify a generalization of the Bernstein 18 SOBOLEV, BESOV AND PALEY-WIENER VECTORS One also has an analogue of the Jackson inequality (see [22], [26]) E(f, ω) ≤ ω−rkfkHr, f ∈ Hr. These two inequalities and Theorem 7.1 imply the following result (compare to [31], [34]). Theorem 7.4. For α > 0, 1 ≤ q ≤ ∞ the norm of Bα H,q(√L), is equivalent to (7.4) kfkH + ∞Xj=0(cid:0)2jαE(f, 2j)(cid:1)q 1/q . Let the functions Fj be as in Subsection 6.4. Theorem 7.5. For α > 0, 1 ≤ q ≤ ∞ the norm of Bα H,q(√L), is equivalent to 1/q (7.5) f 7→ ∞Xj=0(cid:16)2jα(cid:13)(cid:13)(cid:13)Fj(√L)f(cid:13)(cid:13)(cid:13)H(cid:17)q Proof. We obviously have E(f, 2l) ≤Pj>l(cid:13)(cid:13)(cid:13)Fj(√L)f(cid:13)(cid:13)(cid:13)H with the standard modifications for q = ∞. sion of Hardy's inequality [3] we obtain the estimate , . By using a discrete ver- (7.6) kfk + ∞Xl=0(cid:0)2lαE(f, 2l)(cid:1)q!1/q ∞Xj=0(cid:16)2jα(cid:13)(cid:13)(cid:13)Fj (√L)f(cid:13)(cid:13)(cid:13)H(cid:17)q ≤ C 1/q . Conversely, for any g ∈ PW2j−1 (√L) we have(cid:13)(cid:13)(cid:13)Fj(√L)f(cid:13)(cid:13)(cid:13)H =(cid:13)(cid:13)(cid:13)Fj(√L)(f − g)(cid:13)(cid:13)(cid:13)H ≤ kf − gkH. This implies the estimate(cid:13)(cid:13)(cid:13)Fj (√L)f(cid:13)(cid:13)(cid:13)H ≤ E(f, 2j−1), which shows that the inequality opposite to (7.6) holds. The proof is complete. (cid:3) Theorem 7.6. For α > 0, 1 ≤ q ≤ ∞ the norm of Bα H,q(√L) is equivalent to (7.7)  ∞Xj=0 2!q/2 2jαq Xk (cid:12)(cid:12)(cid:12)Df, Φj kE(cid:12)(cid:12)(cid:12) 1/q ≍ kfkBα q , with the standard modifications for q = ∞. Proof. For f ∈ H and operator Fj(√L) we apply (6.8) to Fj(√L)f ∈ PW2j+1 (√L) to obtain (7.8) Since Φj 2 2 k = Fj(√L)φj (1 − δ)(cid:13)(cid:13)(cid:13)Fj(√L)f(cid:13)(cid:13)(cid:13) Xk (cid:12)(cid:12)(cid:12)Df, Φj kE(cid:12)(cid:12)(cid:12) ≤(cid:13)(cid:13)(cid:13)Fj (√L)f(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)Fj (√L)f(cid:13)(cid:13)(cid:13) kE(cid:12)(cid:12)(cid:12) H ≤Xk (cid:12)(cid:12)(cid:12)DFj(√L)f, φj 1 − δXk (cid:12)(cid:12)(cid:12)Df, Φj kE(cid:12)(cid:12)(cid:12) H ≤ k we obtain the following inequality 1 2 2 2 Our statement follows now from Theorem 7.5. 2 H . for all f ∈ H. (cid:3) SOBOLEV, BESOV AND PALEY-WIENER VECTORS 19 8. Applications l = 1, ..., ln be an orthonormal basis in Pn. 8.1. Analysis on Sd. We will specify the general setup in the case of standard harmonics of degree n, which are restrictions to Sd of harmonic homogeneous poly- nomials of degree n in Rd. The Laplace-Beltrami operator ∆Sd on Sd is a restriction unit sphere. Let Sd =(cid:8)x ∈ Rd+1 : kxk = 1(cid:9) . Let Pn denote the space of spherical of the regular Laplace operator ∆ in Rd. Namely, ∆Sd f (x) = ∆ef (x), x ∈ Sd, where ef (x) is the homogeneous extension of f : ef (x) = f (x/kxk). Another way to com- pute ∆Sd f (x) is to express both ∆Sd and f in a spherical coordinate system. Each Pn is the eigenspace of ∆Sd that corresponds to the eigenvalue −n(n + d − 1). Let Yn,l, Let e1, ..., ed+1 be the standard orthonormal basis in Rd+1. If SO(d + 1) and SO(d) are the groups of rotations of Rd+1 and Rd respectively then Sd = SO(d + 1)/SO(d). On Sd we consider vector fields Xi,j = xj∂xi−xi∂xj which are generators of one-parameter groups of rotations exp tXi,j ∈ SO(d + 1) in the plane (xi, xj). These groups are defined by the formulas for τ ∈ R, exp τ Xi,j · (x1, ..., xd+1) = (x1, ..., xi cos τ − xj sin τ, ..., xi sin τ + xj cos τ, ..., xd+1) Let Ti,j(τ ) be a one-parameter group which is a representation of exp τ Xi,j in the space Lp(Sd). It acts on f ∈ Lp(Sd) by the following formula Ti,j(τ )f (x1, ..., xd+1) = f (x1, ..., xi cos τ − xj sin τ, ..., xi sin τ + xj cos τ, ..., xd+1). Let Di,j be a generator of Ti,j in Lp(Sd). The Laplace-Beltrami operator ∆Sd can i,j. One can easily illustrate our results by describing norms in Sobolev and Besov spaces on Sd in terms of operators Ti,j and Di,j. In this situation role of Paley-Wiener subspaces is played by subspaces described in (1.11), (6.7) can be represented be identified with the operator L =Pi<j D2 k ok∈Kρ Pn. A set of functionalsnA(ρ) by a set of Dirac measures at nodes {x(ρ) sphere Sd. k } "nearly uniformly" distributed over the 8.2. Compact homogeneous manifolds. It should be noted, that a similar sit- uation holds on any compact homogeneous manifolds M = G/K where G is a compact Lie group and K is its closed subgroup. Moreover, in this case description of Besov spaces in terms of approximation by Paley-Wiener vectors (eigenfunctions of a corresponding Laplace-Beltrami operator) and in terms of frame coefficients can be extended to any 1 ≤ p ≤ ∞ [7], [37], [6]. A set of functionalsnA(ρ) a set of nodes {x(ρ) can be represented by a set of Dirac measures (or some other functionals [29]) at k } "nearly uniformly" distributed over the manifold M with the spacing comparable to ρ > 0. The Weyl's asymptotic formula [8] implies [28] that a rate of sampling which is given by (6.7) is essentially optimal. k ok∈Kρ 8.3. Non-compact symmetric spaces. Our framework also holds on non-compact symmetric spaces [30]-[35]. Besov spaces can be characterized either using corre- sponding modulus of continuity [32] or by approximation by Paley-Wiener vectors [33] or in terms of frames [35]. In this situation Paley-Wiener functions which admit explicit description in terms of the Helgason-Fourier transform [33] can be represented by a set of Dirac measures (or even more general functionals [29]) at 20 SOBOLEV, BESOV AND PALEY-WIENER VECTORS a set of nodes {x(ρ) spacing comparable to ρ > 0. k } "nearly uniformly" distributed over the manifold M with the Remark 8.1. It should be noted that in the case of non-compact symmetric spaces our approach leads to Sobolev and Besov spaces which are different from the conven- tional ones generated by the Laplace-Beltrami operator associated with the natural metric. Acknowledgement: I would like to thank Dr. Meyer Pesenson for a number of useful suggestions. References 1. J. Bergh, J. Lofstrom, Interpolation spaces, Springer-Verlag, 1976. 2. M. Birman and M. Solomyak, Spectral theory of selfadjoint operators in Hilbert space, D.Reidel Publishing Co., Dordrecht, 1987. 3. P. Butzer, H. Berens, Semi-Groups of operators and approximation, Springer, Berlin, 1967. 4. P. L. Butzer, K. Scherer, Jackson and Bernstein-type inequalities for families of commutative operators in Banach spaces, J. Approx. Theory 5 (1972), 308-342. 5. Dai, Feng; Xu, Yuan, Approximation theory and harmonic analysis on spheres and balls, Springer Monographs in Mathematics. Springer, New York, 2013. xviii+440 pp. ISBN: 978-1- 4614-6659-8; 978-1-4614-6660-4 6. H. G. Feichtinger, H. Fuhr, I. Z. Pesenson, Geometric SpaceFrequency Analysis on Manifolds, Journal of Fourier Analysis and Applications, December 2016, Volume 22, Issue 6, pp 1294- 1355. 7. D. Geller and I. Pesenson, Band-limited localized Parseval frames and Besov spaces on compact homogeneous manifolds, J. Geom. Anal. 21/2 (2011), 334-371. 8. L. Hormander, The Analysis of Linear Partial Differential Operators. III. Pseudo-differential Operators, Springer, Berlin, 2007. 9. S. Krein, I. Pesenson, Interpolation Spaces and Approximation on Lie Groups, The Voronezh State University, Voronezh, 1990, 10. S. Krein, Y. Petunin, E. Semenov, Interpolation of linear operators, Translations of Mathe- matical Monographs, 54. AMS, Providence, R.I., 1982. 11. J. Lions, Theorems de trace et d'interpolation, Ann. Scuola Norm.Sup.Pisa (3)13, 389-403 (1959). 12. J. Lions, J. Peetre, Sur one classe d'espaces d'interpolation, Inst. Hautes Etudes Sci. Publ. Math. 19, 5-68 (1964). 13. E. Nelson, Analytic vectors, Ann. of Math., 70(3), (1959), 572-615. 14. E. Nelson, W. Stinespring, Representation of elliptic operators in an enveloping algebra, Amer. J. Math. 81 (1959), 547-560. 15. Muller, Detlef; Yang, Dachun A difference characterization of Besov and Triebel-Lizorkin spaces on RD-spaces. Forum Math. 21 (2009), no. 2, 259-298. 16. S. M. Nikolskii, Approximation of functions of several variables and imbedding theorems, Springer, Berlin, 1975. 17. Nursultanov, Erlan; Ruzhansky, Michael; Tikhonov, Sergey, Nikolskii inequality and Besov, Triebel-Lizorkin, Wiener and Beurling spaces on compact homogeneous manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 16 (2016), no. 3, 981-1017. 18. J. Peetre, G. Sparr, Interpolation on normed Abelian groups, Ann. Mat. Pura Appl. 92 (1972), 217-262. 19. Pesenson, I. Z., Interpolation of noncommuting operators, (Russian) Uspehi Mat. Nauk 33 (1978), no. 3(201), 183-184. 20. I. Pesenson, Interpolation spaces on Lie groups, (Russian) Dokl. Akad. Nauk SSSR 246 (1979), no. 6, 1298 -- 1303. 21. I. Pesenson, Nikolskii-Besov spaces connected with representations of Lie groups, (Russian) Dokl. Akad. Nauk SSSR 273 (1983), no. 1, 45 -- 49. SOBOLEV, BESOV AND PALEY-WIENER VECTORS 21 22. I. Pesenson, The Best Approximation in a Representation Space of a Lie Group, Dokl. Acad. Nauk USSR, v. 302, No 5, pp. 1055-1059, (1988) (Engl. Transl. in Soviet Math. Dokl., v.38, No 2, pp. 384-388, 1989.) 23. I. Pesenson, On the abstract theory of Nikolskii-Besov spaces, (Russian) Izv. Vyssh. Uchebn. Zaved. Mat. 1988, no. 6, 59 -- 68; translation in Soviet Math. (Iz. VUZ) 32 (1988), no. 6, 8092 24. I. Pesenson, Approximations in the representation space of a Lie group, (Russian) Izv. Vyssh. Uchebn. Zaved. Mat. 1990, no. 7, 43 -- 50; translation in Soviet Math. (Iz. VUZ) 34 (1990), no. 7, 4957. 25. I. Pesenson, The Bernstein Inequality in the Space of Representation of Lie group, Dokl. Acad. Nauk USSR 313 (1990), 86 -- 90; English transl. in Soviet Math. Dokl. 42 (1991). 26. I. Pesenson, A sampling theorem on homogeneous manifolds, Trans. Amer. Math. Soc. 352/9 (2000), 4257 -- 4269. 27. I. Pesenson, Sampling of band limited vectors, J. Fourier Anal. Appl. 7/1 (2001), 93-100. 28. I. Pesenson, An approach to spectral problems on Riemannian manifolds, Pacific J. Math. 215/1 (2004), 183-199. 29. I. Pesenson, Poincar´e-type inequalities and reconstruction of Paley-Wiener functions on man- ifolds, J. Geometric Anal. 4/1 (2004), 101-121. 30. I. Pesenson, Bernstein-Nikolski inequality and Riesz interpolation formula on compact homo- geneous manifolds, J. Approx. Theory, 150/2 (2008), 175-198. 31. I. Pesenson, Paley-Wiener approximations and multiscale approximations in Sobolev and Besov spaces on manifolds, J. Geom. Anal. 19 (2009), no. 2, 390419. 32. I. Z. Pesenson, Bernstein-Nikolskii and Plancherel-Polya inequalities in Lp-norms on non- compact symmetric spaces, Math. Nachr. 282/2 (2009), 253-269. 33. I. Pesenson, A Discrete Helgason-Fourier transform for Sobolev and Besov functions on non- compact symmetric spaces, Contemp. Math. 464, Amer. Math. Soc. (2008), 231-249. 34. I. Z. Pesenson, M. Z. Pesenson, Approximation of Besov vectors by Paley-Wiener vectors in Hilbert spaces, Approximation Theory XIII: San Antonio 2010 (Springer Proceedings in Mathematics), by Marian Neamtu and Larry Schumaker, 249-263. 35. I. Z. Pesenson, Paley-Wiener-Schwartz nearly Parseval frames on noncompact symmetric spaces, Commutative and Noncommutative Harmonic Analysis and Applications, 55-71, Con- temp. Math. 603, Amer. Math. Soc., Providence, RI, 2013. 36. I. Z. Pesenson, Boas-type formulas and sampling in Banach spaces with applications to anal- ysis on manifolds , in New Perspectives on Approximation and Sampling Theory, Springer International Publishing, Switzerland (2014), 39 -61. 37. I. Z. Pesenson, Approximations in L p-norms and Besov spaces on compact manifolds, Con- temporary Mathematics, 650 (2015), 199-210. 38. Pesenson, Isaac Z. Sampling, splines and frames on compact manifolds. GEM Int. J. Geomath. 6 (2015), no. 1, 43-81.
1507.00049
1
1507
2015-06-30T21:52:29
Functional calculus estimates for Tadmor-Ritt operators
[ "math.FA", "math.NA" ]
We show $H^{\infty}$-functional calculus estimates for Tadmor-Ritt operators (also known as Ritt operators), which generalize and improve results by Vitse. These estimates are in conformity with the best known power-bounds for Tadmor-Ritt operators in terms of the constant dependence. Furthermore, it is shown how discrete square function estimates influence the estimates.
math.FA
math
FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS FELIX L. SCHWENNINGER ABSTRACT. We show H¥ -functional calculus estimates for Tadmor–Ritt operators (also known as Ritt operators), which generalize and improve results by Vitse. These estimates are in conformity with the best known power-bounds for Tadmor–Ritt operators in terms of the constant depen- dence. Furthermore, it is shown how discrete square function estimates influence the estimates. 1. INTRODUCTION When studying numerical stability of a difference equation of the form (1.1) xn = T xn−1 + rn, n > 0, the notion of power-boundedness emerges naturally. Here, T is a square matrix or, more general, a linear operator and stability w.r.t. to the initial value x0 can be measured by E = supn∈Nkxn − xnk, (for all x0, x0) reduces to asking if supnkT nk < ¥ where xn denotes the solution to (1.1) with initial value x0. Since kxn − xnk = kT n(x0 − x0)k, the question whether E < ¥ Although the characterization of power-bounded matrices in terms of the eigenvalues is well-known, one aims for different conditions implying power-boundedness, like con- ditions on the resolvent (zI − T )−1. The most famous characterization for matrices is probably given by the Kreiss Matrix Theorem [22, 43]. As the Kreiss Matrix Theorem fails for infinite dimensions, one has to strengthen the con- ditions on the resolvent in order to guarantee power-boundedness. This leads to the notion of Tadmor–Ritt operators. This paper deals with general estimates for Tadmor–Ritt operators, which particularly im- ply power-boundedness. . In the following, let D denote the open unit disc in the complex plane, D its closure and ¶ D its boundary. The spectrum of a linear (bounded) operator T : X → X on a Banach space X will be denoted by s (T ) and the resolvent set by r (T ). For z ∈ r (T ), we define the resolvent R(z,T ) := (zI − T )−1. In the following, all considered operators T are assumed to be linear and bounded. For an open set W ⊂ C, H¥ , equipped with the supremum norm k·k¥ 1.1. Tadmor–Ritt and Kreiss operators. In the following we will give a brief introduc- tion about Tadmor–Ritt operators, and explain their relation to more general Kreiss opera- tors. Unless stated otherwise, X will denote a general Banach space. (W ) denotes the space of bounded analytic functions on W . ,W 2010 Mathematics Subject Classification. 47A60, 47A99, 65M12. Key words and phrases. Functional calculus; Tadmor–Ritt operator; Ritt operator; Square function estimates; Power-bounded operator; Kreiss Matrix Theorem. The author has been supported by the Netherlands Organisation for Scientific Research (NWO), grant no. 613.001.004. 1 2 FELIX L. SCHWENNINGER Definition 1.1. An operator T on X is called a Tadmor–Ritt operator if s (T ) ⊂ D and if (1.2) C(T ) := sup z>1k(z− 1)R(z,T )k < ¥ . Let T R(X) denote the set of all Tadmor–Ritt operators on X. Tadmor–Ritt operators, in the literature also sometimes referred to as Ritt operators, were, with a slightly different but equivalent definition, first studied in [41]. See [5, 9, 12, 49] for a detailed discussion of these two definitions. Tadmor–Ritt operators form a class consisting of operators satisfying Kreiss’ resolvent condition, (1.3) s (T ) ⊂ D, and CKreiss(T ) = sup z>1k(z− 1)R(z,T )k < ¥ . We will call operators satisfying (1.3) Kreiss operators and denote the set of all such oper- ators on X by KR(X). Obviously, T R(X) ⊂ KR(X). The most prominent question related to these operators is the one of power-boundedness, i.e. whether Pb(T ) := sup n∈NkT nk < ¥ . In 1962, O. Kreiss studied the question for finite-dimensional spaces X, [22]. He showed that in this case the answer is positive and that for all T ∈ KR(X), (1.4) Pb(T ) ≤ g(CKreiss(T ),N), for a function g depending on CKreiss(T ) and the dimension N of the space X. Kreiss’ originial estimate (of the function g) was improved steadily in the following decades ending up with the final result proved by Spijker in 1991, [43], (1.5) ∀T ∈ KR(X) : Pb(T ) ≤ eCKreiss(T )N. For the detailed history of the result we refer to the monograph [47] and the recent work [37]. By [29], estimate (1.5) is sharp in the sense that there exists a sequence of matrices TN ∈ KR(CN×N) such that Pb(TN) CKreiss(TN)N = e. lim N→¥ However, for this sequence, CKreiss(TN) → ¥ , hence, for CKreiss(T ) ≤ C with a fixed con- stant C, the behavior could theoretically be better. Indeed, a recent result by Nikolski shows that for T having unimodular spectrum, i.e. s (T ) ⊂ ¶ D, and a basis of eigenvectors, one gets a sublinear growth in the dimension. Theorem 1.2 (N. Nikolski 2013 [37]). Let X be a Hilbert space of dimension N < ¥ be a Kreiss operator on X such that s (T ) ⊂ ¶ D and such that T has a basis of eigenvectors XN = (x j)N . Let T j=1. Then where e = 0.32 b(XN )2 and b(XN) denotes the basis constant of XN, i.e. Pb(T ) ≤ 2p CKreiss(T )N1−e , b(XN) = sup k≤l kP(x j)k j=lk, denotes the projection onto the span of the vectors (x j)k j=l. where P(x j)k j=l The proof of Theorem 1.2 is based on a classic theorem by McCarthy and Schwartz [32]. We remark that Nikolski also shows a corresponding result on more general Banach spaces using a generalization of McCarthy and Schwartz’ result by Gurari and Gurari [15]. By using well-known techniques from Spijker, Tracogna, Welfert [44], he further proves that the sublinear behavior is sharp. As indicated by Nikolski, in order to get an estimate in the spirit of the Kreiss Matrix Theorem, one has to close the loop by estimating b(XN) in terms of CKreiss(T ). This still remains open. FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 3 If we turn to general infinite-dimensional spaces X, the power-boundedness of general Kreiss operators, even on Hilbert spaces, is no longer true. We refer to [13] and [17] for counterexamples. In the conference paper [45], E. Tadmor states that the growth of kT nk can at most be logarithmically in n under the additional assumption that the spectrum of T ‘is not too dense in the neighbourhood of the unit circle’. This condition is in particu- lar ensured if (1.2) holds. Moreover, the existence of an example is stated confirming the sharpness of the growth. As both the proof and the example are unfortunately not pub- lished, we are indebted to E. Tadmor for sharing them with us, [46]. Knowing that general Kreiss operators are not power-bounded, the same question for Tadmor–Ritt operators remained open until 1999 when Lyubich, [31], and Nagy & Ze- manek, [34] used a preceding result of O. Nevanlinna, [35], to prove that they are indeed power-bounded. We remark that in 1993, C. Palencia [38] and, independently, Crouzeix, Larsson, Piskarev and Thom´ee [11] showed that the Crank–Nicolson-scheme is stable for sectorial operators. In particular, this shows that the Cayley transform Cay(A) := (I−A)(A+I)−1 of a sectorial operator A is power-bounded. As it well-known that the map- ping A 7→ Cay(A) establishes a one-to-one correspondence between sectorial operators A with 0 ∈ r (A) and Tadmor–Ritt operators, the result already shows the power-boundedness of Tadmor–Ritt operators. This fact seems to be unnoticed in the literature. Moreover, Pa- lencia’s result shows that any bounded operator S with s (S) ⊂ D and such that there exists a constant M(S) > 0 and z > 1, is power-bounded. Note that Tadmor–Ritt operators are of this form. kR(z,S)k ≤ M(S)(z + 1−1 +z− 1−1), In 2002, El-Fallah and Ransford, [12] showed that for a Tadmor–Ritt operator T , Pb(T )≤ C(T )2, which was subsequently improved by Bakaev [5] to (1.6) ∀T ∈ T R(X) : Pb(T ) ≤ aC(T )log(aC(T )), for some absolute constant a > 0 (which was not determined). The latter result seems to be not so well-known. In [50, Remark 2.2] an alternative proof for the quadratic dependence on C(T ) is sketched. A careful study of this sketch reveals that it is based on a similar approach as in Bakaev’s proof, which, with a sharper estimation and some additional work, actually yields (1.6). We will encounter a similar approach in the proof of Theorem 2.3, which was actually motivated by a result of the author for analytic semigroups, [42]. In [48, 50] Vitse investigated the more general setting of a functional calculus for j=m a jz j Tadmor–Ritt operators and proved that for 1 ≤ m ≤ n and any polynomial p(z) = (cid:229) n (1.7) kp(T )k ≤ c(C(T ),m,n)· sup z∈Dp(z), with c(C(T ),m,n) = 191C(T )5 log(cid:16) e(n+1) m (cid:17). We also remark that Le Merdy showed in [25] that a Tadmor–Ritt operator on a Hilbert space has bounded polynomially calculus, i.e. (1.8) sup{kp(T )k : p is polynomial,kpk¥ ,D ≤ 1} < ¥ , if and only if T is similar to a contraction. Obviously, (1.7) implies power-boundedness of T , however, yet with a C(T )-dependence worse than in (1.6). By functional calculus, more general functions f in H¥ (D) can be considered instead of polynomials p in (1.7). This leads to the study of the H¥ -calculus for Tadmor–Ritt operators [2, 3, 24]. We will show that the constant c(C(T ),m,n) in (1.7) can be improved significantly, cou- pling it to the, so-far known, optimal constant for the power-bound of T in (1.6). Precisely, 4 FELIX L. SCHWENNINGER in Theorem 2.5 we will show that for p(z) = (cid:229) n kp(T )k ≤ aC(T )log(cid:18)C(T ) + b + log j=m a jz j, 0 ≤ m ≤ n, m + 1(cid:19)·kpk¥ n + 1 ,D, (1.9) with absolute constants a,b > 0. The proof is shorter and more direct than the one for (1.7) in [51]. Note also that we allow for m = 0. Moreover, the result is actually a consequence of a more general functional calculus result for Tadmor–Ritt operators, see Theorem 2.3. Finally, motivated by the result for analytic semigroup generators (i.e. sectorial oper- ators) [42], which can be seen as the continuous counterparts of Tadmor–Ritt operators, we discuss the influence of square function estimates on the the calculus estimates, see also [27]. For Hilbert spaces, it is known that if a Tadmor–Ritt operator and its dual op- erator satisfy square function estimates, then the corresponding H¥ -functional calculus is bounded. As for the more known continuous counterpart of sectorial operators, here, it is essential to have square function estimates for both T and T ∗. We show that having only T (or alternatively T∗) satisfying square function estimates however improves the functional calculus estimate (1.9), see Theorem 3.6. In Section 4, we generalize the result about square function estimates to general Banach spaces. This involves a refined definition of square function estimates using Rademacher means and R-boundedness. These abstract square function estimates are the discrete counterpart to the ones for sectorial operators, which were introduced by Kalton and Weis [21] and have proved very useful in the study of Lp-maximal regularity for parabolic evolution equations since then. In Section 6, we discuss sharpness of the derived estimates. We conclude by a result about a Besov-space calculus for Tadmor–Ritt operators, which is a refinement of [50, Theorem 2.5]. 1.2. Properties of Tadmor–Ritt operators. Unless stated otherwise, X will always de- note a, in general infinite-dimensional, Banach space. From (1.2) it follows that for Tadmor–Ritt operators the only possible spectral point on T is 1. Moreover, it is well-known that the spectrum is contained in the Stolz type domain Bq , p which is the interior of the convex hull of {{1} ,Bsinq (0)} for some q ∈ (0, 2 ), see Figure 2. Here, Br(z0) denotes the open ball centred at z0 with radius r. For this and a proof of the following lemma we refer to Vitse [49, 50] and Le Merdy [27], which improves earlier results in [31, 34] and [35]. Lemma 1.3. Let T be a Tadmor–Ritt operator on a Banach space X. Then, there exists q ∈ [0, p 2 ) such that (i) s (T ) ⊂ Bq , and p (ii) for all h ∈ (q , 2 ], (1.10) Ch (T ) = supz∈C\Bh k(z− 1)R(z,T )k ≤ We say that T is of type q . Moreover, q can always chosen to be q = arccos 1 C(T ) . C(T ) cosq 1− cosh . Note that Ba ⊂ Bb for a < b . The previous lemma tells us that for h going to q , the p 2 it becomes C(T ). We further remark 2 ) such that s (T ) ⊂ Bq right-hand-side of (1.10) explodes whereas for h = that the converse of Lemma 1.3 also holds: If there exists q ∈ (0, and Ch < ¥ We further need the following well-known characterization, which can be found e.g., in [27, 31, 34, 50]. p 2 ), then T is Tadmor–Ritt, see [27, Lemma 2.1]. for all h ∈ (q , p Lemma 1.4. Let T be an operator on a Banach space X. The following assertions are equivalent. FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 5 i i · q Bq 0 1 ¶ Bq 0 h r 1 ¶ W h ,r FIGURE 1. The sets Bq and W h ,r with h ∈(cid:0)q , p 2(cid:1). (i) T is Tadmor–Ritt. (1.11) Pb(T ) = sup (ii) The sets {T n : n ∈ N} and(cid:8)n(T n − T n−1) : n ∈ N(cid:9) are bounded, i.e. n∈Nkn(T n − T n−1)k < ¥ , and c1,T := sup Let us emphasize that supn∈Nkn(T n − T n−1)k < ¥ does not imply power-bounded- ness of T in general, see [18]. Hence, in (ii) of Lemma 1.4, the assumption of power- boundedness cannot be dropped. See also [36] for a discussion on power-boundedness related to estimates on the resolvent. n∈NkT nk < ¥ . 2. A FUNCTIONAL CALCLULUS RESULT FOR TADMOR–RITT OPERATORS By Lemma 1.3 we know that the spectrum of a Tadmor–Ritt operator is contained in the Stolz type domain Bq , with q = arccos simply connected subset of C. Then for any function holomorphic on W can be defined via the Riesz-Dunford integral C(T ) . Let W ⊃ Bq be an open, bounded and , the operator f (T ) 1 1 2p iZG f (T ) = f (z) R(z,T ) dz, is a rectifiable, positively orientated, simple contour inside W which encircles Bq . (W ) denote the bounded holomorphic functions on W (2.1) where G Let H¥ Remark 2.1. Let H¥ (Bd ) for which exist constants c,s > 0 such that f (z) ≤ c1 − zs for all z ∈ Bd . For d ∈ (q , 0 (Bd ), f (T ) can still be defined by (2.1) with G equal to the boundary of ¶ Bd ′ of Bd ′ with d ′ ∈ (q ,d ). Analogously to the situation for sectorial operators, see e.g., [16], it can be shown that the mapping f 7→ f (A) becomes an algebra homomorphism from H¥ 0 (Bd ) to B(X), see [27, 0 (Bd ) be the functions f in H¥ 2 ) and f ∈ H¥ . p Section 2] for more details. For 0 < r < 1 and h ∈ (0, (2.2) see Figure 2. The function p 2 ], we define the ‘keyhole-shaped’ set, h ,r := Bh ∪ Br(1), (2.3) e−x x is known as the Exponential integral. It holds that (2.4) dx Ei(s) =Z ¥ s(cid:1) < Ei(s) < e−s log(1 + 1 s ), s ∈ (0, 1 2 ], Ei(s) < log( 1 (2.5) 1 2 e−s log(cid:0)1 + 2 s s ), s > 0, W 6 FELIX L. SCHWENNINGER (2.6) where ¶ W see [14] and [42] for more details. The following lemma outsources technicalities in the proof of the results to come, Theorem 2.3. Estimates of this kind for deriving functional calculus estimates can already be found in [5, 38, 50], see also [28, 30] for a slightly different setting. Here, the focus is laid on deriving estimates explicitly in the used constants. p 2 ), we have that G(m,h ,r) :=Z¶ W Lemma 2.2. For 0 < r < 1 , m ≥ 0 and h ∈ (0, zm z− 1 dz ≤ C (r,m,h ) cosh + 4Ei(cid:0)r m+1 h ,r denotes the boundary of the set defined in (2.2) and C (r,m,h ) := 4(sinh )m+1 log 4 2 cosh (cid:1) + 2p (1 + r)m, C (r,m,h ) ≤ −8 logcosh − 4 log(r(m + 1)) + 2p (1 + r)m + 12 log2. where Ei is defined in (2.3). If r ≤ 1 m+1 , by (2.5), Proof. Let us first assume that r < cosh . We split up the path ¶ W where G 2 denotes the union of the two straight line segments of ¶ W Figure 2), whereas G 1, G 3 denote the part of ¶ W Br(1), respectively (dotted lines in Figure 2). Precisely, h ,r = G 1 ∪ G 2 ∪ G 3, h ,r (dashed lines in h ,r that lies on the circles Bsinh (0) and h ,r ,t ∈ (r,cosh ](cid:9) , i p For G 1, we see that 2 − h ,p ]o,G 2 =(cid:8)1− te±ih G 1 =n(sinh )eid G 3 =n1 + reid Next we estimate Gi :=RG ,d ∈ ( ,d ∈ [0,p − h )o, zm z−1dz for i = 1,2,3. G1 = 2(sinh )m+1Z p eix sinh − 1 2 −h Since 2Reix − 1 ≥ eix − 1 for all R,x ≥ 0, = 2√2(sinh )m+1Z p eix − 1 2 −h 1√1−cosx for x ∈ (0,p ), we derive G1 ≤ 4(sinh )m+1Z p 2 −h 4 is a primitive of G1 ≤ −4(sinh )m+1 logtan Since √2 logtan x dx dx . p p p p 2 − h 4 ≤ 4(sinh )m+1 log 4p 2 −h , dx √1− cosx . p where in the last step we used that tan x ≥ x for x ∈ [0, 4 ], which follows from the Taylor series of tan. Since sin x ≤ x for all x ≥ 0, we finally get 4 (2.7) cosh G1 ≤ 4(sinh )m+1 log . To estimate G2, note that 1− teih G2 = ZG 2 zmdz z− 1 2 e 1 Since 1− x ≤ e−x and e− x 2Z cos2 h G2 ≤ 2e r cosh 1 e−x m Finally, G3 can be estimated by r = 2Z cosh 1− teih 2 ≥ 1 for x ∈ [0,1], 2Z ¥ 2 − x 1 + reid dx x ≤ 2e 1 2 G3 = 2Z p −h 0 (2.8) 2 ≤ (1− t cosh ) for t ∈ [0,cosh ] and thus, 1− t cosh t ≤ 2Z cosh m dt m 2 r dt t r m+1 2 cosh e−x x dx = 4Ei(cid:0)r m+1 2 cosh (cid:1) . m dd ≤ 2p (1 + r)m. FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 7 This shows (2.6) for r < cosh . If r ≥ cosh , then ¶ W the convenient parts of the circles ¶ Bsinh (0), ¶ Br(1) such that G 1 ∪ G 3 = ¶ W G 1 =(cid:8)sinh eid b depending on h . Since r ≥ cosh h ,r = ¶ (Bsinh (0)∪ Br(1)). Hence, we choose G 1 and G 3 to be h ,r, i.e. ,d ∈ [0,p − b )(cid:9) for certain angles a , 2 − h and hence, we can cosh + 2p (1 + r)m, ,d ∈ (a ,p ](cid:9) and G 3 =(cid:8)1 + reid +ZG 3 ≤ G1 + 2p (1 + r)m ≤ 4(sinh )m+1 log 4 G(m,h ,r) =ZG 1 estimate similarly as in (2.7) and (2.8), it is easy to see that a > p which concludes the proof as the right hand side is smaller than C (r,m,h ). Theorem 2.3. Let T be a Tadmor–Ritt operator on X. Let q = arccos m ∈ N0, r ∈ (0,1) and h ∈ (q , p 2 ) we have, with t m(z) = zm, that h ,r k( f · t m)(T )k ≤ c(T,m,r,h )·k fk¥ (2.9) ,W (cid:3) 1 C(T ) . Then, for for f ∈ H¥ (W h ,r). Here, c(T,m,r,h ) ≤ Ch (T ) 2p C (r,m,h ) where Ch (T ) = supz∈C\Bh k(z− 1)R(z,T )k, and C as in Lemma 2.2. Proof. Let h ∈ (q , H(W 2 ) and r > 0. By Lemma 1.3 we know that s (T ) ⊂ W p h ,r. Let f ∈ h ,r, h ,r). Since f t m is holomorphic on W 1 ( f t m)(T ) = where h ∈ (q ,h ) and r ∈ (0,r). Since W 2p k( f t m)(T )k ≤ 2p iZ¶ W h ,r ⊂ W k fk¥ C h (T ) h ,r C h (T ) ≤ 2p k fk¥ f (z)zmR(z,T ) dz, h ,r, h ,r Z¶ W zm z− 1 h ,r · C (r,m, h ). h ,r ,W ,W dz The last inequality followed by Lemma 2.2. Letting ( h , r) → (h ,r) yields that C h (T ) → Ch (T ) by the maximum principle (applied to z 7→ hx′,zR(z,T )xi for x ∈ X, x′ ∈ X′) and that C (r,m, h ) → C (r,m,h ) since C is continuous (see Lemma 2.2). Together, this gives the assertion. (cid:3) The following inequality is a direct consequence of the maximum principle. The disc p 2 ) can be traced back to S. Bernstein, and can be found in [40, p. 346], or [39, case (h = Problem III. 269, p.137]. Lemma 2.4. Let Ba , a ∈ (0, ing assertions hold. p 2 ], be the Stolz type domain defined in Sec. 1.2. The follow- (2.10) (i) For a polynomial p of degree n, and r ≥ 1, sina (cid:17)n ,rBa ≤(cid:16) r kpk¥ ·kpk¥ ,Ba . (ii) For f ∈ H(Ba ) and continuous on Ba , m ∈ N and t m(z) = zm, k f · t mk¥ ,Ba ≤ k fk¥ ,Ba ≤ 1 (sina )m k f · t mk¥ ,Ba . (2.11) (2.12) Proof. The assertion is a consequence of the maximum principle applied to p(z)z−n. In fact, let z ∈ C\ Ba . Then, since z 7→ p(z)z−n is analytic at ¥ , by the maximum principle, p(z)z−n ≤ max z∈¶ Ba p(z)z−n ≤ max z∈¶ Ba z−n·kpk¥ ,Ba . (2.13) with absolute constants a,b, that can be chosen as kp(T )k ≤ aC(T )(cid:18)2 logC(T ) + b + log b = −2 log(s) + 6, a = 2e p (1−s), ,D, n + 1 m + 1(cid:19)·kpk¥ s ∈ (0,1). 8 FELIX L. SCHWENNINGER It is easy to see that maxz∈¶ Ba z−1 = 1 that z ≤ r for z ∈ ¶ (rBa ) ⊂ C\ Ba yields sina (cid:17)n p(z) ≤(cid:16) r kpk¥ Therefore, (2.10) follows by the maximum principle. ,Ba , z ∈ ¶ (rBa ). sina . Hence, multiplying (2.12) by zn and noting It is easy to see that sina ≤ z for z ∈ ¶ Ba . Therefore, by the maximum principle, k fk¥ ,Ba = sup z∈¶ Ba f (z) ≤ (sina )m sup The other inequality of (2.11) is clear as Ba ⊂ D. Theorem 2.5. Let T be a Tadmor–Ritt operator on X and let m,n ∈ N such that 0 ≤ m ≤ n. Then, for any p(z) = (cid:229) n z∈¶ Ba zm f (z) = k=m akzk, we have that (sina )m k f t mk¥ ,Ba (cid:3) 1 1 Proof. Let p(z) = (cid:229) n n− m. For s ∈ (0,1] let h (s) = arccos k=m akzk = zm p0(z) with 0 ≤ m ≤ n and p0 is a polynomial of degree C(T ) . By Theorem 2.3 we have for s,r ∈ (0,1) that s (2.14) where p(z) = zm p0. Since W (2.15) kp0k¥ By the maximum principle, kp0k¥ (0,1), Eq. (2.14) becomes ,W kp(T )k ≤ c(T,m,r,h (s))·kp0k¥ ,W h (s),r , h (s),r ⊂ (1 + r)D, Lemma 2.4 (i) (with a = h (s),r ≤ kp0k¥ ,(1+r)D ≤ (1 + r)n−mkp0k¥ ,D. p 2 ) yields ,D = kpk¥ ,D. Hence, by choosing r = t n+1 with t ∈ (2.16) kp(T )k ≤ c(T,m, t n ,h (s))· (1 + t n+1 )n−mkpk¥ ,D. It remains to estimate the right hand side. Clearly, (1 + that t n+1 )n−m ≤ e. Theorem 2.3 yields We can further estimate C using Lemma 2.2. Since r = (−2 logcosh (s) + log n+1 ,h (s)) ≤ c(T,m, 2Ch (s) p t c(T,m, t t Ch (s) 2p C ( n+1 ,h (s)) ≤ t n+1 ,m,h (s)). n+1 ≤ 1 m+1 , n + 1 m + 1 − logt + 1−s for s ∈ (0,1). Since cosh (s) = s C(T ) , n + 1 m + 1 − logt + (2 logC(T )− 2 logs + log p 2 p 2 By Lemma 1.3, Ch (s)(T ) ≤ C(T ) t c(T,m, n ,h (s)) ≤ As mint ∈(0,1) log 1 Corollary 2.6. Let T be a Tadmor–Ritt operator. Then T is power-bounded, + 3 log2 < 6, together with (2.16), this yields (2.13). 2C(T ) p (1− s) p 2 et t + e t + 3 log2). (cid:3) t e + 3 log2). with absolute constants a,b > 0 as in Theorem 2.5. supn∈NkT nk ≤ aC(T ) (2 logC(T ) + b) Remark 2.7. FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 9 (1) Theorem 2.5 shows that a Tadmor–Ritt operator has a bounded H¥ [m,n]-calculus, where H [m,n] = {p(z) = (cid:229) n k=m akzk : ak ∈ C} and m≤ n. With different techniques, such a result was proved by Vitse in [50], see also (1.7). However, in [50] the bound of the calculus depends on a factor C(T )5, whereas in our Theorem 2.5, this gets improved to a behavior of C(T )(logC(T ) + 1). Moreover, Corollary 2.6 shows that the same dependence holds true for the power-bound of a Tadmor–Ritt operator. This confirms the result by Bakaev [5], which seems not so well-known, and improves the better known quadratic depen- dence C(T )2, see [12], [50]. ,Bh -norm for some h < (2) It is a natural question to ask if the k·k¥ ,D-norm in Theorem 2.5 can be replaced p by the sharper k·k¥ 2 . Indeed, Lemma 2.4 allows us to do this, see also (2.15). However, this leads to an additional factor (sinh )−n, which therefore destroys the logarithmic behavior in n+1 m+1 . Let us further remark that a polynomially bounded Tadmor–Ritt operator T (see (1.8)) on a Hilbert space implies an estimate of the form kp(T )k . kpk¥ ,Bh , p 2 . In other words, T allows for a bounded H¥ for some h < (Bh )-calculus. However, this is not true for general Banach spaces, see [24]. More generally, including the Hilbert space case, if one assumes that T is R-Ritt (see Section 4), then polynomial-boundedness does indeed imply a bounded H¥ (Bh )-calculus, see [27, Proposition 7.6] on arbitrary Banach spaces. 3. THE EFFECT OF DISCRETE SQUARE FUNCTION ESTIMATES - HILBERT SPACE In the following we will show that discrete square function estimates improve the de- pendence in the way that log n+1 m+1 in (2.13) gets replaced by its square root. Definition 3.1 (Hilbert space square function estimate). Let T be a bounded operator on a Hilbert space X. We say that T satisfies square function estimates if there exists a K > 0 such that (3.1) kxk2 T := k=1 kkT kx− T k−1xk2 ≤ K2kxk2, ∀x ∈ X. Square function estimates are a well-known tool characterizing bounded H¥ -calculi for sectorial operators, going back to McIntosh’s seminal work in the 80ties [33]. From the 90ties on, H¥ -calculus has proved very useful in the study of maximal regularity. In [10] a suitable Lp-version of square function estimates was introduced which then got further adapted to general Banach spaces by Kalton and Weis in the unpublished note [21], see also [23] and the references therein. Maximal regularity for discrete-time difference equations were investigated in [8, 7]. Discrete square function estimates for Tadmor–Ritt operators were studied in [19]. We mention that in the literature there exists a whole scale of square functions, see [27, Section 3], whereas we only use the specific form in Definition 3.1. As for sectorial operators, for non-Hilbert (typically, Lp-) spaces suitable square func- tion estimates have to be redefined for Tadmor–Ritt operators using Rademacher means. For the moment we will restrict ourselves to the Hilbert space case and leave the general Banach space case for Section 4. The following characterization of bounded H¥ calculus for Tadmor–Ritt operators was re- cently proved in [27]. For the rest of the section we want to emphasize that on Hilbert spaces the notions of R-Ritt and Tadmor operator coincide, whereas on general Banach spaces R-Ritt is stronger than Tadmor–Ritt. For a definition of R-Ritt operators and square function estimates on general Banach spaces, we refer to Section 4. ¥ ¥ (cid:229) 10 FELIX L. SCHWENNINGER Theorem 3.2 (Le Merdy 2014, [27, Corollary 7.5]). Let T be a Tadmor–Ritt operator on a Banach space X. Consider the assertions (i) T is R-Ritt and both T and T ∗ satisfy square function estimates. (ii) For some h ∈ (0, p 2 ), k f (T )k . k fk¥ ,Bh ∀ f ∈ H 0 (Bh ), (3.2) 0 (Bh ) is defined in Remark 2.1. where H¥ Then, (i) ⇒ (ii). If X is UMD space (in particular, a Hilbert space), then (ii) ⇒ (i). The assumption on (geometry of) the Banach space for the direction (i) to (ii) can be further generalized to X having property (D ), see [27, 20]. In [27, Proposition 8.1] it is further shown that there exist Tadmor–Ritt operators (even on Hilbert spaces) such that (only) T satisfies square function estimates, but (3.2) does not hold. However, we will see that having square function estimates for T (or T ∗) does improve the functional calculus estimate in Theorem 2.5. Note that for a Tadmor–Ritt operator T of type q and r ∈ (0,1), rT is again Tadmor–Ritt with C(rT ) = sup r − T(cid:17)−1(cid:13)(cid:13)(cid:13)(cid:13) ≤ C(T ) sup l >1(cid:13)(cid:13)(cid:13)(cid:13)(l − 1) 1 l >1(cid:12)(cid:12)(cid:12)(cid:12) r(cid:16) l l − 1 l − r(cid:12)(cid:12)(cid:12)(cid:12) = 2C(T ) 1 + r 0 (Bh ) with h ∈ (q , We remark that moreover limrր1 f (rT ) = f (T ) for f ∈ H¥ . Lemma 2.3]. Lemma 3.3. Let T be a Tadmor–Ritt operator on a Hilbert space X. For m ∈ N∪{0}, r ∈ (0,1), p 2 ), see [27, (3.3) k(rT )mxkrT ≤ armsb + log(cid:18)1− 1 2(m + 1)logr(cid:19) kxk ∀x ∈ X, with a = √2c1,T and b = 1 + Pb(T )2 c2 1,T , where c1,T and Pb(T ) are defined in (1.11). Proof. Clearly, rT is a Tadmor–Ritt operator. By definition, ≤ r2m k(rT )mxk2 rT = r2m k=1 kkrkT k+mx− rk−1T k−1+mxk2 kr2(k−1)(cid:16)2kT k+mx− T k−1+mxk2 + 2k(1− r)T k+mxk2(cid:17) kr2(k−1) 2c2 (k + m)2 + 2(1− r)2Pb(T )2!kxk2, where c1,T = supn∈Nkn(T n − T n−1)k which is finite by Lemma 1.4. Since ≤ r2m (3.4) k=1 k=1 1,T k (k+m)2 ≤ 1 k+m , r2k k + 1 + m kr2(k−1) (k + m)2 ≤ k=1 ≤ = ≤ (3.5) k=0 1 m + 1 1 m + 1 1 m + 1 0 dx e2xlogr x + 1 + m +Z ¥ + r−2(m+1)Ei(−2(m + 1)logr) 2(m + 1)logr(cid:19) , + log(cid:18)1− 1 ¥ ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 11 where the last step follows by (2.4). Using this and the fact that (cid:229) we can conclude in (3.4) that k(rT )mxk2 1,T(cid:18) 1 m + 1 1,T r2m(cid:18)b + log(cid:18)1− 1 + log(cid:18)1− 2(m + 1)logr(cid:19)(cid:19) + 2(m + 1)logr(cid:19)(cid:19)kxk2, 1 rT ≤ r2m(cid:20)2c2 ≤ 2c2 . for b = 1 + Pb(T )2 c2 1,T k=1 kr2(k−1) = 1 (1−r2)2 , (1 + r)2(cid:21)kxk2 2Pb(T )2 (cid:3) Another lemma, we will need, is the following result relating square function estimates for T and rT as r ր 1. This can be seen as a discrete analog of [26, Proposition 3.4]. Lemma 3.4. Let T be a Tadmor–Ritt operator on a Hilbert space. Then, the following are equivalent (i) T satisfies square function estimates. (ii) rT satisfies square function estimates uniform in r ∈ (0,1), i.e., kxkrT ≤ K kxk. Proof. This follows from the more general Lemma 4.6 in Section 4. ∃K > 0 ∀r ∈ (0,1)∀x ∈ X : (cid:3) The following theorem is essentially Le Merdy’s key argument to prove that (i) implies (ii) in Theorem 3.2. As we need its precise form, we state it explicitly. For a proof we refer to [27, Proof of Theorem 7.3]. For a definition of R-Ritt operator of R-type q we refer to Section 4. For the moment it suffices to remark that on Hilbert spaces this notion is equivalent to the of one a Tadmor–Ritt operator of type q , see Section 4. Theorem 3.5 (Le Merdy 2014). Let T be a R-Ritt operator of R-type q on a Banach space X. Let 0 < q < h < p 2 . Then, there exists c = c(h ,C(T )) > 0 such that hy, p(T )xi ≤ c·kpk¥ for any polynomial p, x ∈ X and y ∈ X∗. (Note that the right-hand-side is allowed to be ¥ ,Bh ·kxkT ·kykT∗, ). Combining Theorem 3.5 and Lemma 3.3 yields the following refinement of Theorem 2.5. Theorem 3.6. Let T be a Tadmor–Ritt operator on a Hilbert space X. Assume that either T or T ∗ satisfies square function estimates. Then, for integers 0 ≤ m ≤ n and p(z) = (cid:229) n j=m a jz j, kp(T )k ≤ acKe 1 2 ·rb + log n + 2 m + 1 ·kpk¥ ,D, r 1 (z) = p( z p 2 ). Define p 1 with K,a,b,c defined in (3.1), Lemma 3.3 and Theorem 3.5, respectively. Proof. Since X is a Hilbert space, T is R-Ritt of type q = arccos C(T ). Let r ∈ (0,1) and choose h ∈ (q , (rT ) = p(T ) since p is r ). It is easy to see that p 1 a polynomial. Furthermore, we write p(z) = zmq(z) for q having degree n− m. Therefore, for all x ∈ X, (3.6) W.l.o.g. let T ∗ satisfy square function estimates. Hence, by Lemma 3.4, kykrT∗ ≤ Kkyk for all y ∈ X∗ and all r ∈ (0,1). Applying Theorem 3.5 for rT and p = q 1 ,D ·k(rT )mxkrT ·kyk, (3.7) (rT )(rT )mxi ≤ cK ·kq 1 r k¥ p(T )x = q 1 r (rT )(rT )mx. hy,q 1 yields r r r ¥ 12 FELIX L. SCHWENNINGER for x ∈ X,y ∈ X∗ where we used that Bh ⊂ D. By Lemma 2.4 (i) and the maximum ,D. Therefore, and by Lemma 3.3, Eq. (3.7) principle, kq 1 r k¥ yields ,D = rm−nkpk¥ ,D ≤ rm−nkqk¥ r kq 1 Choose r = e− (3.6), 1 (rT )(rT )mk ≤ acKrb + log(cid:16)1− 2rb + log 2(n−m+1) . Then 1− kp(T )k ≤ acKe 2(m+1) logr = n+2 1 1 1 2(m+1) logr(cid:17)· r2m−n ·kpk¥ ,D. m+1 and r2m−n = e n−2m 2(n−m+1) < e 1 2 . Thus, by n + 2 m + 1 ·kpk¥ ,D. (cid:3) Remark 3.7. (1) The proof idea of Theorem 3.6 can also be used for an alternative proof of the logarithmic behavior in Theorem 2.5, if we do a similar computation for kT mykrT∗ (instead of assuming square function estimates kykT . kyk). This finally yields another factor of the formqb + log n+2 in the k·k¥ the estimate. (2) As explained in Remark 2.7, in Theorem 3.6 we can also derive ‘sharper’ estimates ,Bh -norm at the price that additional factors of the form (sinh )−n enter m+1 . 4. DISCRETE SQUARE FUNCTION ESTIMATES ON GENERAL BANACH SPACES (4.1) As indicated in Section 3, for non-Hilbert spaces, Definition 3.1 is not suitable for characterizing boundedness of the H¥ -calculus. For Lp-spaces the proper replacement is given by kxkT :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)(cid:229) 2(cid:13)(cid:13)(cid:13)(cid:13)Lp k=1 kT kx− T k−1x2(cid:17) 1 . kxk , m ), where T is a Tadmor–Ritt operator on Lp(W ), p ∈ [1,¥ ), for some measure space (W see [19], [27] and the references therein. By Fubini’s theorem, this definition coincides with Definition 3.1 if p = 2. However, to cover general Banach spaces, we need the following generalization using Rademacher averages. This approach (for sectorial operators), paving the way for a lot of research in this field, was introduced by Kalton and Weis in their ‘famous’ unpublished note, see the new preprint [21]. For an excellent overview on the topic we refer to [23]. The discrete version of these general square function estimates for Tadmor–Ritt operators recently appeared in [27]. We briefly recap the definition of the needed Rademacher norms. For more details, we refer to [27, 23]. For k ≥ 1, we define the Rademacher function e k(t) = sgn(sin(2kp t)). It is easy to see that (e k)k≥1 forms an orthonormal basis in L2(I) with I = [0,1]. For a Banach space X let us consider the linear span of elements e k ⊗ x = (t 7→ e k(t)x), k ≥ 0, x ∈ X, in the Bochner space L2(I,X). Denote the closure of this set, w.r.t. the norm in L2(I,X), by Rad(X). Hence, Rad(X) becomes a Banach space with the norm for elements x = (cid:229) Rademacher functions it follows that (4.2) Now we can define a general square function by Rad(X) =(cid:8)(cid:229) e k(t)xk(cid:13)(cid:13)2 dt(cid:19) 1 k xkRad(X) =(cid:18)ZI(cid:13)(cid:13)(cid:229) e k ⊗ xk : xk ∈ X,the sum converges in L2(I,X)(cid:9) . k e k ⊗ xk with (xk)k being a finite family in X. By orthonormality of the k=1 , k 2 kxkT =(cid:13)(cid:13)(cid:13) k=1 e k ⊗ k(T kx− T k−1x)(cid:13)(cid:13)(cid:13)Rad(X) , ¥ ¥ (cid:229) ¥ FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 13 if (cid:229) k e k ⊗ k(T kx− T k−1x) /∈ Rad(X). where we set kxkT = ¥ Definition 4.1 (Square function estimates for Tadmor–Ritt operators). Let T be a Tadmor– Ritt operator on a Banach space X. We say that T satisfies (abstract) square function estimates, if there exists KT > 0 such that for all x ∈ X, (4.3) kxkT =(cid:13)(cid:13)(cid:13) e k ⊗ k k=1 1 2 (T kx− T k−1x)(cid:13)(cid:13)(cid:13)Rad(X) ≤ KTkxk. Note that if X is a Hilbert space, as a consequence of Parseval’s identity, this definition of square function estimates coincides with the one given in Definition 3.1. Precisely, for any finite sequence (xk)k ∈ X, (4.4) 1 2 , which shows that both definitions of square functions estimates coincide. Further, it can be shown that for X = Lp = Lp(W (cid:13)(cid:13)(cid:229) k = ((cid:229) e k ⊗ xk(cid:13)(cid:13)Rad(X) , m ) (p ∈ [1,¥ ) and (W e k ⊗ xk(cid:13)(cid:13)Rad(Lp) ∼(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)(cid:229) k (cid:13)(cid:13)(cid:229) k kxkk2) , m ) being s -additive), , 2(cid:13)(cid:13)(cid:13)(cid:13)Lp k x2(cid:1) 1 see [23, Remark 2.9]. Hence, (4.1) is equivalent to having square function estimates using Rademacher averages. The notion of R-boundedness emerges naturally in the framework of the space Rad(X). After being introduced in [6], it has been proved very useful in the study of maximal regularity, see [23] for a detailed introduction. Definition 4.2. Let X be a Banach space and T ⊂ B(X) a set of bounded operators. Then, T is called R-bounded if there exists a constant M such that for any finite family (Tk))k ∈ T , and finite sequence (xk)k ⊂ X, (4.5) . (cid:13)(cid:13)(cid:229) k e k ⊗ Tkxk(cid:13)(cid:13)Rad(X) ≤ M(cid:13)(cid:13)(cid:229) k e k ⊗ xk(cid:13)(cid:13)Rad(X) The smallest possible constant C is called the R-bound. By (4.4), it follows that for Hilbert spaces the notion of R-boundedness of T coin- cides with (uniform) boundedness of T in the operator norm. However, in general, R- boundedness only implies boundedness, see [1]. Now we are able to introduce R-Ritt operators, which first appeared in [8, 7]. Nonetheless the notion R-Tadmor–Ritt would be more consistent in this Chapter, we use the name R-Ritt following Le Merdy [27]. For Hilbert spaces, the following notion is equivalent to the one of a Tadmor–Ritt operator, see Lemma 1.4. Definition 4.3. An operator T on a Banach space X is called R-Ritt if the sets are R-bounded. We denote the bounds by PbR(T ) and cR {T n : n ∈ N} and (cid:8)n(T n − T n−1) : n ∈ N(cid:9) 1,T , respectively. By Lemma 1.4, an R-Ritt operator is always a Tadmor–Ritt operator and the notions coincide on Hilbert spaces. Moreover, the following R-Ritt version of Lemmata 1.3 and 1.4 holds. For a proof, see [27, Lemma 5.2] and [8]. Lemma 4.4. Let T be a bounded operator on a Banach space X. The following assertions are equivalent. (i) T is R-Ritt. (ii) s (T ) ⊂ Bq (4.6) In this case, we say that T is of R-Ritt type q . 2 ) and for all h ∈ (q , for some q ∈ [0, (cid:8)(z− 1)R(z,T ) : z ∈ C\ Bh (cid:9) is R-bounded. p 2 ] p Now we are ready to prove the corresponding R-Ritt version of the results in Section 3 for general Banach spaces. (cid:229) ¥ 14 FELIX L. SCHWENNINGER Lemma 4.5. Let T be a R-Ritt operator on a Banach space X. For m ∈ N∪{0}, r ∈ (0,1), k(rT )mxkrT ≤ armsbR + log(cid:18)1− 1 2(m + 1)logr(cid:19) kxk ∀x ∈ X, with aR = √2cR 1,T and bR = 1 + PbR(T )2 1,T )2 , where cR (cR 1,T ,PbR(T ) are defined in Def. 4.3. Proof. The proof technique is very similar to the proof of Lemma 3.3. Therefore, we will focus on the arguments involving R-boundedness. Since rT is a Tadmor–Ritt operator, we have, see 4.3, where the last step follows since T is R-Ritt. By the definition of the Rad(X)-norm, and Parseval’s identity (for L2[0,1]), the first norm in (4.7) equals 1 2 rk−1x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Rad(X) (4.7) 1 + k=1 k=1 k=1 k=1 ≤ cR 1 e k ⊗ k k 1 2 rk−1 k + m kT mxkrT =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 (rkT k+mx− rk−1T k−1+mx)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Rad(X) e k ⊗ k ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) e k ⊗h(T k+m − T k−1+m) + (1− r)T k+mik x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Rad(X) 1,T(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) e k ⊗ + (1− r)PbR(T )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 rk−1x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Rad(X) k+m x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =Z 1 0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k+m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = kxk2Z 1 k + m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k=1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = kxk2 2(m + 1)logr(cid:19)(cid:19) 1 2(m + 1)logr(cid:19)(cid:19) 1 k(rT )mxkrT ≤ rm"cR 1,T(cid:18) 1 + log(cid:18)1− 1,T r2m(cid:18)bR + log(cid:18)1− √2cR k=1 k 1 2 rk−1 k 1 2 rk−1 k + m x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) e k ⊗ e k(t) k e k(t) k 1 2 rk−1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 2 rk−1 m + 1 1 1 2 . Rad(X) k=1 k=1 2 , 2 X dt 2 dt The remaining series can be estimated as in the Hilbert space proof. Analogously, the second norm in (4.7) can be computed. Therefore, we derive, 2 + PbR(T ) (1 + r)#kxk 2 kxk, (cid:3) ≤ for bR = 1 + PbR(T )2 1,T )2 . (cR We further need the generalization of Lemma 3.4 to (abstract) square function estimates. Lemma 4.6. Let T be a R-Ritt operator on a Banach space X. Then, the following are equivalent. (i) T satisfies (abstract) square function estimates. (ii) rT satisfies (abstract) square function estimates uniform in r ∈ (0,1), ∃K > 0 ∀r ∈ (0,1)∀x ∈ X : kxkrT ≤ K kxk. ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ (cid:229) FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 15 Proof. The proof is similar to one for the continuous time analog [26, Proposition 3.4] and is based on using the identity (I − T )T kx = (I − rT )T kx + (1− r)T k+1x. (4.8) (cid:13)(cid:13)(cid:13) This yields, using that T is R-Ritt, e k ⊗ k k=1 1 2 rk(I − T )T kx(cid:13)(cid:13)(cid:13)Rad(X) ≤(cid:13)(cid:13)(cid:13) e k ⊗ k k=1 1 + 2 (I − rT )(rT )kx(cid:13)(cid:13)(cid:13)Rad(X) 2 rkx(cid:13)(cid:13)(cid:13)Rad(X) e k ⊗ k k=1 1 + PbR(T )(1− r)(cid:13)(cid:13)(cid:13) . It is easy to see that the second term on the right-hand is bounded in r ∈ (0,1), because the Rad(X)-norm equals ((cid:229) 2kxk = kxk(1− r2)−1 by Parseval’s identity. Hence, by Fatou’s lemma, we get that (ii) implies (i). The other direction also follows, with a similar estimation, from (4.8). (cid:3) k=1 kr2k) 1 The Banach space version of Theorem 3.6 now follows completely analogously to the Hilbert space proof with Lemmata 4.5 and 4.6 (instead of Lemmata 3.3 and 3.4). Theorem 4.7. Let T be a R-Ritt operator on a Banach space X. Assume that either T or T ∗ satisfies (abstract) square function estimates. Then, for integers 0 ≤ m ≤ n and p(z) = (cid:229) n j=m a jz j, kp(T )k ≤ aRcKT e 1 2 ·rbR + log n + 2 m + 1 ·kpk¥ ,D, with KT ,aR,bR and c defined in (4.3), Lemma 4.5 and Theorem 3.5, respectively. 5. SHARPNESS OF THE ESTIMATES It is natural to ask whether the deduced functional calculus estimates from Theorems 2.5 and 3.6, (5.1) and (5.2) kp(T )k ≤ aC(T )(cid:18)logC(T ) + b + log 2 ·rb2 + log kp(T )k ≤ a2cKT e 1 n + 1 m + 1(cid:19)kpk¥ ,D, n + 2 m + 1 ·kpk¥ ,D, [m,n], that is p(z) = (cid:229) n for p ∈ H¥ k=m akzk, are sharp. Clearly, here ‘sharpness’ has different aspects depending on the variables C(T ),m,n it is referring to. For a clear discussion, we distinguish between the following questions. (A) Is (5.1) sharp in the variables m,n, with 0 ≤ m ≤ n? (B) Is (5.1) sharp in the variable C(T ) for (some) fixed m,n? (C) Question (A) for (5.2). (D) Question (B) for (5.2). To answer these questions, we introduce the quantity C(T,m,n) = sup{kp(T )k : p ∈ H (5.3) [m,n],kpk¥ ,D ≤ 1} . Question (A) was discussed Vitse in [50, Remark 2.6] using the prior works [49, 51]. In particular, she showed that if X contains a complemented isomorphic copy of ℓ1 or ℓ (e.g., infinite-dimensional L1 or C(K) spaces), then there exists a Tadmor–Ritt operator on X such that C(T,m,n) & log ne m , (cid:229) ¥ (cid:229) ¥ (cid:229) ¥ ¥ ¥ ¥ 16 FELIX L. SCHWENNINGER where the involved constant only depends on X and is thereby linked with constant C(T ). However, the precise dependence on C(T ) is not apparent there. If X is an (infinite- dimensional) Hilbert space (more, generally if the Banach space X contains a comple- mented isomorphic copy of ℓ2), then for any d ∈ (0,1), there exists a Tadmor–Ritt operator such that C(T,m,n) &(cid:16)log ne m(cid:17)d . n (or ℓ2 n respectively). We refer to [48, 50] for details. These statements can be generalized to more general spaces X that uniformly contain uni- form copies of ℓ1 Question (B) can be split up in several cases. If m = 0, hence p is an arbitrary polynomial of degree n, (5.1) implies that C(T,0,n) . C(T )(logC(T ) + log(n + 1)). Hence, we ob- serve ‘linear’ asymptotic behavior in C(T ) as n → ¥ . In fact, in [48, Theorem 2.1] it is shown that it is indeed linear, namely C(T,0,n) ≤ (C(T ) + 1)log(e2n), and there exists a T on some Banach space X such that C(T,0,n) ∼ log(e2n). We point out that the proof technique, [48, Theorem 2.1], requires m = 0. However, for m = n, Question (B) reduces to the prominent question of the optimal power- bound for T . As mentioned in Corollary 2.6, (5.1) yields C(T,n,n) = kT nk . C(T )(logC(T ) + 1), for all n. This is so-far the best known power-bound for Tadmor–Ritt operators, see also [5]. It remains open whether this can be replaced by a linear C(T )-dependence. Furthermore, motivated by the Kreiss Matrix Theorem (1.5), it is not clear whether for N-dimensional spaces X, an estimate of the form (5.4) for some scalar function g can be achieved, where g(N) ∈ o(N). Note that the estimate for g(N) = eN trivially holds by (1.5) and the fact that CKreiss(T ) ≤ C(T ). Let us turn to Question (C) now. We want to show sharpness of Pb(T ) ≤ C(T )g(N) (5.5) ∃c > 0 ∀N ∈ N : m+1(cid:1)d then C(T,m,n) &(cid:0)log n+1 if the basis is Besselian, i.e., ∃cy > 0 (5.6) . k=1 N(cid:229) a kxky k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) : a k ≤ 1,(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) sup((cid:13)(cid:13)(cid:13)(cid:13)(cid:13) cy (cid:0)(cid:229) 2 ≤(cid:13)(cid:13)(cid:229) k xky k(cid:13)(cid:13) , kxk2(cid:1) 1 N(cid:229) k=1 xky k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1) ≥ cN d , As for sectorial Schauder multipliers, it holds that T satisfies square function estimates C(T,m,n) .rlog n + 1 m + 1 under the assumption that T satisfies square function estimates. Therefore, we construct T as a Schauder basis multiplier, which is a well-known technique to construct unbounded calculi, see e.g., [16, Chapter 9] and [4], where it was introduced. Let X be a separable infinite-dimensional Hilbert space with a bounded Schauder basis {y k}. For a sequence (l n) ⊂ [0,1], define the bounded operator T = Ml by for finite sequences (xk) ⊂ C. Let l n = 1−2−n, then T is Tadmor–Ritt, see [27, Proposition 8.2]. With this setting we can use the following argument from [51, Proof of Theorem 2.1]. Let d ∈ (0,1). If for the uniform basis constant ub({y k}N k=1) & Nd , i.e. k=1) it holds that ub({y k}N k xky k(cid:1) = (cid:229) T x =(cid:0)(cid:229) l kxky k, k FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 17 for finite sequences (xk) ⊂ C, see [26, Theorem 5.2] and [27, Theorem 8.2]. Note that (5.6) k=1) ≤ cy m(y )√N, where m(y ) = supkky kk. It remains to already implies that ub({y k}N find a Besselian basis {y k} such that (5.5) is fulfilled for d ∈ (0, 1 Indeed, such an 2 ). example can be constructed for an L2-space on the unit circle with suitable weight, see [26, Thm. 5.2], [42], and [44, Section 4.3]. In fact, the example in [42, Thm. 4.5] gives a basis e−ikt , k ∈ N0, y 2k(t) = t−b y 2k+1(t) = t−b 2 ). Moreover, it is shown that there exist elements eikt , (there, the notation is y ∗) with b ∈ ( 1 3 , 1 x,y ∈ L2 such that xn ∼ n3b −1, where x = (cid:229) n xny n and y ∈ (cid:229) n yny ∗n , and where {y ∗n} denotes the dual basis such that hy ∗k ,y ni = d nk. Choosing a n = 1 such that a nxnyn ∈ R≥0, we deduce n3b −2 ∼ N3b −1. b −1, n ∈ N, a nxny ni = a nxnyn & yn ∼ n hy, N(cid:229) N(cid:229) N(cid:229) n=1 n=1 n=1 n=1 xny nk ≤ b(y )kxk, (5.5) follows for d = 3b − 1 ∈ (0, 1 Since k(cid:229) N proved the following result, which answers (C) for Hilbert spaces. Theorem 5.1. There exists a Hilbert space such that for any d ∈ (0, 1 Tadmor–Ritt operator T which satisfies square function estimates and 2 ). Therefore, we have 2 ) there exists a C(T,m,n) &(cid:18)log n + 1 m + 1(cid:19)d holds, where C(T,m,n) is defined in (5.3). Note that the involved constants depend on d . An open question is whether there exists an R-Ritt operator on a Banach space such that m+1(cid:1) 1 T satisfies square function estimates and C(T,m,n) &(cid:0)log n+1 By c1,T . C(T )3, see [50], and c . Pb(T )3c1,T , see [27, Proof of Theorem 7.3], we can track C(T ) in the constants of the estimate in Theorem 3.6. This yields a C(T )-dependence, which seems far from being sharp. Hence, the answer to (D) is probably ‘no’. 2 . 6. FURTHER RESULTS As a direct corollary of the improvements of Vitse’s result, we get the following result for the Besov space functional calculus of T , which in turn is a slight improvement of [50, Theorem 2.2]. For details of the following notions and facts see [50] and the references ,1(D) is defined by the functions f ∈ H(D) such therein. Recall that the Besov space B¥ that ,D +Z 1 It is well known that there exists an equivalent definition via the dyadic decomposition f = n=0 Wn∗ f , where Wn, n ≥ 1 are shifted Fejer type polynomials, whose Fourier coefficients bWn(k) are the integer values of the triangular-shaped function supported in [2n−1,2n+1] with peak bWn(2n) = 1 and W0(z) = 1 + z. Here (g∗ f )(z) = (cid:229) k=0 g(k) f (k)zk. Then, k fkB := k fk¥ )dr < ¥ f ′(reia maxa 0 . ,D < ¥ ,1(D) ⇐⇒ f ∈ H(D) and k fk∗ = f ∈ B¥ n=0kWn ∗ fk¥ . ,D-estimate of Theorem 2.5 to derive ,1(D)-functional calculus estimates. This follows the same lines as in [50], however, Since Wn ∗ f is a polynomial, we can use the k·k¥ B¥ using the improved constant dependence of our result in Theorem 2.5. (cid:229) ¥ ¥ ¥ (cid:229) 18 FELIX L. SCHWENNINGER k f (T )k . C(T )(log(C(T ) + 1))k fk∗ Theorem 6.1. Let T be a Tadmor–Ritt operator on a Banach space X. Then, (6.1) i.e., for all f ∈ B¥ Proof. Since Wn ∗ f ∈ H¥ the definition of H¥ [2n−1,2n+1] for n ≥ 1 and W0 ∗ f ∈ H¥ [m,n], we can apply Theorem 2.5 to derive 2n+1 + 1 ,1(D), where f (T ) is defined by (cid:229) n=0(Wn ∗ f )(T ). k(Wn ∗ f )(T )k ≤ aC(T )(cid:18)2 logC(T ) + b + log 2n−1 + 1(cid:19)kWn ∗ fk¥ 2n−1+1 ≤ 5. Analogously, k(W0 ∗ n=0k(Wn ∗ f )(T )k . k fk∗, and hence, f (T ) is well- ,D, for n ≥ 1, with absolute constants a,b > 0. Clearly, 2n+1+1 f )(T )k can be estimated. Thus, (cid:229) defined with [0,1], see Remark 2.7 for k f (T )k ≤ aC(T ) (2 logC(T ) + b + log5)k fk∗. (cid:3) ,1(D)-estimate as in (6.1) is derived, but with a In [50, Theorem 2.5] a similar k · kB¥ C(T )-dependence of C(T )5. Acknowledgements. The author would like to express his deepest gratitude to Hans Zwart for numerous discussions, and many very helpful comments on the manuscript. He is very thankful to Eitan Tadmor for an interesting discussion in November 2014, and for sharing with him the proof of a result in [45]. He is grateful to Joseph Ball for being his host at the Department of Mathematics at Virginia Tech in fall 2014, where parts of this manuscript were written. He would also like to thank Mark Embree for inspiring discussions on the Kreiss Matrix Theorem during that time. REFERENCES [1] Wolfgang Arendt and Shangquan Bu. The operator-valued Marcinkiewicz multiplier theorem and maximal regularity. Math. Z., 240(2):311–343, 2002. [2] C´edric Arhancet, Stephan Fackler, and Christian Le Merdy. Isometric dilations and h¥ calculus for bounded analytic semigroups and ritt operators. available at arXiv: 1504.00471, 2015. [3] C´edric Arhancet and Christian Le Merdy. Dilation of Ritt operators on Lp-spaces. Israel J. Math., 201(1):373–414, 2014. [4] J.-B. Baillon and Ph. Cl´ement. Examples of unbounded imaginary powers of operators. J. Funct. Anal., 100(2):419–434, 1991. [5] N. Yu. Bakaev. Constant size control in stability estimates under some resolvent conditions. Num. methods and Prog., 4:348–357, 2003. [6] Earl Berkson and T. A. Gillespie. Spectral decompositions and harmonic analysis on UMD spaces. Studia Math., 112(1):13–49, 1994. [7] Sonke Blunck. Analyticity and discrete maximal regularity on Lp-spaces. J. Funct. Anal., 183(1):211–230, 2001. [8] Sonke Blunck. Maximal regularity of discrete and continuous time evolution equations. Studia Math., 146(2):157–176, 2001. [9] N. Borovykh, D. Drissi, and M. N. Spijker. A note about Ritt’s condition, related resolvent conditions and power bounded operators. Numer. Funct. Anal. Optim., 21(3-4):425–438, 2000. [10] Michael Cowling, Ian Doust, Alan McIntosh, and Atsushi Yagi. Banach space operators with a bounded H¥ functional calculus. J. Austral. Math. Soc. Ser. A, 60(1):51–89, 1996. [11] M. Crouzeix, S. Larsson, S. Piskarev, and V. Thom´ee. The stability of rational approximations of analytic semigroups. BIT, 33(1):74–84, 1993. [12] Omar El-Fallah and Thomas Ransford. Extremal growth of powers of operators satisfying resolvent condi- tions of Kreiss-Ritt type. J. Funct. Anal., 196(1):135–154, 2002. [13] S. R. Foguel. A counterexample to a problem of Sz.-Nagy. Proc. Amer. Math. Soc., 15:788–790, 1964. [14] Walter Gautschi. Some elementary inequalities relating to the gamma and incomplete gamma function. J. Math. and Phys., 38:77–81, 1959/60. [15] V. I. Gurariı and N. I. Gurariı. Bases in uniformly convex and uniformly smooth Banach spaces. Izv. Akad. Nauk SSSR Ser. Mat., 35:210–215, 1971. [16] Markus Haase. The Functional Calculus for Sectorial Operators, volume 169 of Operator Theory: Ad- vances and Applications. Birkhauser Verlag, Basel, 2006. ¥ ¥ FUNCTIONAL CALCULUS ESTIMATES FOR TADMOR–RITT OPERATORS 19 [17] P. R. Halmos. On Foguel’s answer to Nagy’s question. Proc. Amer. Math. Soc., 15:791–793, 1964. [18] N. Kalton, S. Montgomery-Smith, K. Oleszkiewicz, and Y. Tomilov. Power-bounded operators and related norm estimates. J. London Math. Soc. (2), 70(2):463–478, 2004. [19] N. J. Kalton and P. Portal. Remarks on ℓ1 and ℓ¥ -maximal regularity for power-bounded operators. J. Aust. Math. Soc., 84(3):345–365, 2008. [20] N. J. Kalton and L. Weis. The H¥ -calculus and sums of closed operators. Math. Ann., 321(2):319–345, 2001. [21] Nigel Kalton and Lutz Weis. The H¥ -Functional Calculus and Square Function Estimates. unpublished, available at arXiv: 1411.0472, 2001. [22] Heinz-Otto Kreiss. Uber die Stabilitatsdefinition fur Differenzengleichungen die partielle Differentialgle- ichungen approximieren. Nordisk Tidskr. Informations-Behandling, 2:153–181, 1962. [23] Peer C. Kunstmann and Lutz Weis. Maximal Lp-regularity for parabolic equations, Fourier multiplier theo- rems and H¥ -functional calculus. In Functional analytic methods for evolution equations, volume 1855 of Lecture Notes in Math., pages 65–311. Springer, Berlin, 2004. [24] Florence Lancien and Christian Le Merdy. On functional calculus properties of ritt operators. Preprint, available at arXiv: 1301.4875, 2013. [25] Christian Le Merdy. The similarity problem for bounded analytic semigroups on Hilbert space. Semigroup Forum, 56(2):205–224, 1998. [26] Christian Le Merdy. The Weiss conjecture for bounded analytic semigroups. J. London Math. Soc. (2), 67(3):715–738, 2003. [27] Christian Le Merdy. H¥ functional calculus and square function estimates for Ritt operators. Rev. Mat. Iberoam., 30(4):1149–1190, 2014. [28] Marie-Noelle Le Roux. Semidiscretization in time for parabolic problems. Math. Comp., 33(147):919–931, 1979. [29] Randall J. LeVeque and Lloyd N. Trefethen. On the resolvent condition in the Kreiss matrix theorem. BIT, 24(4):584–591, 1984. [30] Christian Lubich and Olavi Nevanlinna. On resolvent conditions and stability estimates. BIT, 31(2):293– 313, 1991. [31] Yu. Lyubich. Spectral localization, power boundedness and invariant subspaces under Ritt’s type condition. Studia Math., 134(2):153–167, 1999. [32] C. A. McCarthy and J. Schwartz. On the norm of a finite Boolean algebra of projections, and applications to theorems of Kreiss and Morton. Comm. Pure Appl. Math., 18:191–201, 1965. [33] Alan McIntosh. Operators which have an H¥ functional calculus. In Miniconference on operator theory and partial differential equations (North Ryde, 1986), volume 14 of Proc. Centre Math. Anal. Austral. Nat. Univ., pages 210–231. Austral. Nat. Univ., Canberra, 1986. [34] B´ela Nagy and Jaroslav Zem´anek. A resolvent condition implying power boundedness. Studia Math., 134(2):143–151, 1999. [35] Olavi Nevanlinna. Convergence of iterations for linear equations. Lectures in Mathematics ETH Zurich. Birkhauser Verlag, Basel, 1993. [36] Olavi Nevanlinna. On the growth of the resolvent operators for power bounded operators. In Linear op- erators (Warsaw, 1994), volume 38 of Banach Center Publ., pages 247–264. Polish Acad. Sci., Warsaw, 1997. [37] N. Nikolski. Sublinear dimension growth in the Kreiss matrix theorem. Algebra i Analiz, 25(3):3–51, 2013. [38] C. Palencia. A stability result for sectorial operators in Banach spaces. SIAM J. Numer. Anal., 30(5):1373– 1384, 1993. [39] G. P´olya and G. Szego. Aufgaben und Lehrsatze aus der Analysis, volume Band I. Springer, Berlin, 1925. [40] Marcel Riesz. Uber einen Satz des Herrn Serge Bernstein. Acta Math., 40(1):337–347, 1916. [41] R. K. Ritt. A condition that limn→¥ n−1T n = 0. Proc. Amer. Math. Soc., 4:898–899, 1953. [42] Felix L. Schwenninger. On measuring unboundedness of the H¥ -calculus for generators of analytic semi- groups. Submitted, 2015. [43] M. N. Spijker. On a conjecture by LeVeque and Trefethen related to the Kreiss matrix theorem. BIT, 31(3):551–555, 1991. [44] M. N. Spijker, S. Tracogna, and B. D. Welfert. About the sharpness of the stability estimates in the Kreiss matrix theorem. Math. Comp., 72(242):697–713 (electronic), 2003. [45] E. Tadmor. The resolvent condition and uniform power boundedness. Linear Algebra Appl., 80:250–252, 1986. [46] E. Tadmor. On Tadmor’s paper “The resolvent condition and uniform boundedness”, November 2014. Pri- vate Communication. [47] Lloyd N. Trefethen and Mark Embree. Spectra and pseudospectra. Princeton University Press, Princeton, NJ, 2005. [48] Pascale Vitse. Functional calculus under the Tadmor-Ritt condition, and free interpolation by polynomials of a given degree. J. Funct. Anal., 210(1):43–72, 2004. 20 FELIX L. SCHWENNINGER [49] Pascale Vitse. The Riesz turndown collar theorem giving an asymptotic estimate of the powers of an operator under the Ritt condition. Rend. Circ. Mat. Palermo (2), 53(2):283–312, 2004. [50] Pascale Vitse. A band limited and Besov class functional calculus for Tadmor-Ritt operators. Arch. Math. (Basel), 85(4):374–385, 2005. [51] Pascale Vitse. A Besov class functional calculus for bounded holomorphic semigroups. J. Funct. Anal., 228(2):245–269, 2005. FELIX L. SCHWENNINGER, UNIVERSITY OF TWENTE,, P.O. BOX 217, 7500 AE ENSCHEDE, THE NETHERLANDS, TEL.: +31-53-4892230 E-mail address: [email protected]
1710.03057
2
1710
2018-02-09T16:02:56
Queer Poisson brackets
[ "math.FA", "math-ph", "math.DG", "math-ph" ]
We give a method to construct Poisson brackets $\{\cdot,\cdot\}$ on Banach manifolds~$M$, for which the value of $\{f,g\}$ at some point $m\in M$ may depend on higher order derivatives of the smooth functions $f,g\colon M\to{\mathbb R}$, and not only on the first-order derivatives, as it is the case on all finite-dimensional manifolds. We discuss specific examples in this connection, as well as the impact on the earlier research on Poisson geometry of Banach manifolds. Those brackets are counterexamples to the claim that the Leibniz property for any Poisson bracket on a Banach manifold would imply the existence of a Poisson tensor for that bracket.
math.FA
math
QUEER POISSON BRACKETS DANIEL BELTITA Institute of Mathematics "S. Stoilow" of the Romanian Academy, 21 Calea Grivitei Street, 010702 Bucharest, Romania TOMASZ GOLIŃSKI University in Białystok, Institute of Mathematics, Ciołkowskiego 1M, 15-245 Białystok, Poland ALICE-BARBARA TUMPACH Université de Lille, Laboratoire Painlevé, CNRS U.M.R. 8524, 59 655 Villeneuve d'Ascq Cedex, France Abstract. We give a method to construct Poisson brackets { · , · } on Banach manifolds M , for which the value of {f, g} at some point m ∈ M may depend on higher order derivatives of the smooth func- tions f, g : M → R, and not only on the first-order derivatives, as it is the case on all finite-dimensional manifolds. We discuss specific examples in this connection, as well as the impact on the earlier re- search on Poisson geometry of Banach manifolds. Those brackets are counterexamples to the claim that the Leibniz property for any Poisson bracket on a Banach manifold would imply the existence of a Poisson tensor for that bracket. 1. Introduction The Poisson brackets in infinite-dimensional setting have played for a long time a significant role in various areas of mathematics including mechanics (both classical and quantum) and integrable systems theory (see e.g. [Fad80, B´00, AMR02, CM74]). However the rigorous approach to the notion of Poisson manifold in the context of Banach space is relatively recent (see [OR03]). It is E-mail addresses: [email protected], [email protected], [email protected]. 1 2 QUEER POISSON BRACKETS known that the Poisson brackets on infinite-dimensional manifolds lack some of the properties known from the finite-dimensional case. It was shown for instance in [OR03] that the existence of Hamiltonian vector fields requires an additional condition on the Poisson tensor in the case of manifolds modelled on a non-reflexive Banach space (i.e. a Banach space E that is not canoni- cally isomorphic to its second dual E E ∗∗, where E ∗ denotes the topo- logical dual of a Banach space). Another example of a new behaviour can be found in [Dit05] -- a Poisson bracket defined only on a certain space of smooth functions might lead to an unbounded Poisson tensor. Moreover on some manifolds, Poisson brackets need not be local although as far as we know a counterexample is not known yet, see a related discussion in [CP12]. The aim of this paper is to prove by example still another phenomenon that is specific to Poisson geometry on an infinite dimensional manifold M, namely the existence of Poisson brackets of higher order. That is, Leibniz property does not ensure that the bracket depends only on the first-order derivatives of functions. The constructed Poisson brackets serve as a counterexample to the statements given in the literature (see [OR03] or subsequently [Ida11]), where it was claimed that the existence of a Poisson tensor Π follows from Leibniz property and skew symmetry of the Poisson bracket {·, ·}, in particular for every m ∈ M one could find a bounded bilinear functional Πm : T ∗ mM × T ∗ mM → R satisfying {f, g}(m) = Πm(f ′ m, g′ m) m, g′ m ∈ T ∗ where f ′ mM are the differentials of f, g ∈ C ∞(M) at point m ∈ M. There is a related fact in [AMR02, Thm. 4.2.16], but we show that it is not applicable here (see Proposition 2.6). We prove that there exist Poisson brackets not given by Poisson tensors on the family of Banach sequence spaces lp for 1 ≤ p ≤ 2 and present an explicit example for p = 2. Such Poisson brackets do not allow to introduce the dynamics by Hamilton equations in the usual way, thus from the point of view of applications in physics one should explicitly assume the existence of Poisson tensor in the definition of a Poisson Banach manifold. In section 2 we investigate "queer operational tangent vectors", that is deriva- tions on spaces of smooth functions on the manifold which are differential op- erators of order higher than 1. This notion was introduced with several results on their existence (including the examples on the Hilbert space) in [KM97]. We explore the case of queer vectors of order 2 on the family of Banach sequence spaces lp for 1 ≤ p < ∞. Section 3 contains our main result, which shows a way to construct higher order Poisson brackets out of queer vector fields, and we illustrate the general result by a specific example on the Hilbert space. We conclude the paper with a version of the definition of Banach Poisson manifold which clarifies QUEER POISSON BRACKETS 3 the one introduced in [OR03]. Some discussion on the problem of localization of Poisson bracket is also included. All Banach and Hilbert spaces considered in this paper are real. By man- ifold we will always mean a smooth real manifold modelled on a Banach space. 2. Queer operational vector fields There are two major approaches to tangent vectors, namely the kinematic one and the operational one. These approaches lead to the same notion for finite-dimensional manifolds, but this is no longer the case in infinite dimen- sions. A kinematic tangent vector to a Banach manifold M at a point m ∈ M is an equivalence class of curves passing through that point (for precise defi- nition see e.g. [AMR02]). On the other hand, an operational tangent vector is defined as a derivation acting in the space of germs of functions (see [KM97], [CP12]). For any m ∈ M consider the set of all functions f : U → R defined on an open neighborhood U of m. One defines an equivalence relation in that set in the following way: two functions f1 : U1 → R and f2 : U2 → R are equivalent if there exists an open neighborhood U ⊂ U1 ∩ U2 of m for which the restrictions of f1 and f2 to U coincide. Any equivalence class defined in this way is called a germ at the point m ∈ M. We denote the set of germs of all smooth functions at m by C ∞ m (M). We note that the value and the derivatives of germs at m ∈ M (that is, jets of germs) are well defined. We denote by Lk(TmM; R) the Banach space of bounded k-linear func- m ∈ Lk(TmM; R) be the k-th tionals on TmM with values in R and let f (k) differential at the point m ∈ M of a germ or a function. Definition 2.1. An operational tangent vector at point m ∈ M is a linear map δ : C ∞ m (M) → R satisfying Leibniz rule : δ(f g) = δf g(m) + f (m) δg. (2.1) For any open subset U ⊆ M with m ∈ U there is a canonical map C ∞(U) → C ∞ m (M) that takes every function on U to its germ at m, hence one has a canonical pull-back of δ to C ∞(U), also denoted by δ. An operational vector field on M is a collection of maps δU : C ∞(U) → C ∞(U) for each open set U ⊂ M, compatible with restrictions to open sub- sets and defining an operational tangent vector δm at every m ∈ M. 4 QUEER POISSON BRACKETS Definition 2.2. The operational tangent vector δ is of order n if it can be expressed in the form δf = n X k=1 ℓk(f (k) m ), (2.2) where ℓk : Lk(TmM; R) → R are continuous and linear. Moreover we re- quire that ℓn does not vanish identically on the subspace of symmetric n-linear maps in Lk(TmM; R). Otherwise the order of δ is infinite. The operational tangent vectors of order at least 2 are called queer. The operational vector field δ is of order at most n if there exists a family (Lk(TmM; R))∗ satisfying (2.2) at of smooth sections ℓk of the bundle F each m ∈ M. m∈M The Leibniz rule (2.1) satisfied by δ implies certain algebraic conditions on functionals ℓk, see [KM97, 28.2]. By definition, operational tangent vectors of order n depend only on the nth jet of functions. The existence of infinite order operational tangent vectors is an open problem as far as we know. Remark 2.3. Any kinematic tangent vector defines an operational tangent vec- tor of order 1. On the other hand in the case of manifolds modelled on non- reflexive Banach spaces, operational tangent vectors of order 1 are given by elements of T ∗∗M which is larger than the (kinematic) tangent bundle T M. Thus in the case of Banach manifolds (even the ones having a global chart, as for instance Banach spaces), the notions of kinematic tangent vector and operational tangent vector do not coincide in general. There are examples of Banach spaces possessing queer operational tangent vectors even in the reflexive case. A construction of second order operational tangent vectors on Hilbert spaces was given in [KM97] and we explore it below for a class of Banach spaces. Let E be a Banach space and consider the natural inclusion of E ∗ × E ∗ into L2(E; R) by : E ∗ × E ∗ → L2(E; R) (f, g) 7→ (f ⊗ g : (v, w) 7→ f (v)g(w)) . (2.3) In general (contrary to the finite-dimensional case) the linear span of its im- age may not be dense. A functional ℓ ∈ (L2(E; R))∗ defines an operational tangent vector of order 2 at any a ∈ E by δℓf = ℓ(f ′′ a ) (2.4) if and only if it vanishes on E ∗ × E ∗ regarded as a subspace of L2(E; R) via (2.3). We also recall here that we can identify L2(E; R) with L(E; E ∗). QUEER POISSON BRACKETS 5 Proposition 2.4. There are no operational tangent vectors of the second order on the Banach space lp of p-summable sequences for 2 < p < ∞. On the other hand, if 1 ≤ p ≤ 2 there are non-trivial operational tangent vectors of the second order. Proof. The proof of existence of operational tangent vectors of the second order has common idea with [KM97, Rem. 28.8]. Namely it is equivalent to the existence of a nonzero continuous linear functional ℓ that vanishes on (lp)∗ × (lp)∗. According to Pitt's theorem, every map from lp to (lp)∗ is compact if 2 < [Pit36], [Rya02, Thm. 4.23], [FHH+01, Prop 6.25]. Moreover p < ∞, see e.g. since all (lp)∗ spaces have the approximation property, the closure of linear span of (lp)∗ × (lp)∗ coincides with the space of compact operators from lp to (lp)∗ [Rya02, Ch. 4]. So, the only continuous functional ℓ which would vanish on E ∗ × E ∗ is the zero functional. Thus there are no non-zero operational tangent vectors of the second order on lp for 2 < p < ∞. In the case 1 ≤ p ≤ 2, the inclusion map ι : lp ֒→ (lp)∗ is not compact, so using Hahn -- Banach theorem it is possible to define a non-zero functional ℓ on L2(E; R) that vanishes on the image of the map (2.3). This implies the existence of non-zero operational tangent vectors of the second order on lp for 1 ≤ p ≤ 2. (cid:3) In particular for p = 2 we obtain an operational tangent vector of the second order on the separable Hilbert space H. We will present this case more explicitly. Example 2.5 (concrete queer operational vector on a Hilbert space). The Banach space L2(H; R) can be identified with the Banach space of bounded operators L∞(H). This identification maps a bilinear map B to the operator A defined by B(v, w) = hAv, wi (2.5) using Riesz theorem. The closure of the linear span of H∗ × H∗ considered as a subspace of L2(H; R) ≃ L∞(H) by inclusion (2.3) is the ideal of com- pact operators on H. One can now obtain the continuous functional ℓ with required properties by putting e.g. ℓ(1) = 1 where 1 denotes the identity map, and ℓ(K) = 0 for any compact operator K ∈ L∞(H) and extending it to the whole L∞(H) by means of Hahn -- Banach theorem. Let us now demonstrate explicitly that the operational tangent vector δℓ given by (2.4) with ℓ defined as above is not a kinematic tangent vector. With- out loss of generality we fix the point a = 0. Taking for example the function ρ(v) = hv, vi (2.6) 6 QUEER POISSON BRACKETS for v ∈ H, we get ρ′′ v = 2 1, where we have used the identification L2(H; R) ≃ L∞(H) given by (2.5). From definition it follows that δℓ(ρ) = 2. On the other hand, any kinematic tangent vector to H at 0 can be identified with some w ∈ H and w · ρ = hw, 0i + h0, wi = 0. Thus δℓ is in fact a queer tangent vector. One can extend δℓ to a queer constant operational vector field on H, which we will denote by the same symbol. Let us note that [AMR02, Thm. 4.2.16] states that for manifolds M modelled on Banach spaces with norm smooth away from the origin, a certain space of derivations is isomorphic to the vector space of kinematic vector fields on M. In this reference, a derivation D on the Banach manifold M is a collection of linear maps C ∞(M, F ) → C ∞(M, F ) for all Banach spaces F , such that for any f ∈ C ∞(M, F ), g ∈ C ∞(M, G), and any bilinear map B : F ×G → H, the following Leibniz rule holds D (B(f, g)) = B(Df, g) + B(f, Dg), (2.7) where F , G, and H are Banach spaces. An example of such a derivation is the Lie derivative. Let us show that existence of δℓ in Example 2.5 is not a contradiction with this result. Namely the operational vector field δℓ cannot be extended to a derivation in the sense of [AMR02]. Proposition 2.6. The queer operational vector field δℓ constructed in Exam- ple 2.5 cannot be extended to a derivation on all C ∞(H, F ) spaces, where F is any Banach space. Proof. Let us assume that there exists an extension Dℓ of δℓ. Let B be the natural duality pairing between H∗ and H. Consider the maps f : H → H∗, v 7→ hv, ·i and g equal to the identity map on H. Then B(f, g)(v) = hv, vi = ρ(v), and Dℓ(cid:0)B(f, g)(cid:1)(v) = δℓ(ρ)(v) = ℓ(2 1) = 2. On the other hand, B(Dℓf, g)(v) + B(f, Dℓg)(v) = B(Dℓf (v), v) + hv, Dℓg(v)i. This expression vanishes for v = 0, hence (2.7) cannot be satisfied for any extension of δℓ. (cid:3) Proposition 2.7. Let δ be an operational vector field of finite order on a manifold M. Then the set of points at which it is queer is open while the set of points at which it is kinematic is closed in M. QUEER POISSON BRACKETS 7 Proof. Let n be the order of δ. The set of points at which δ is not queer is the intersection ℓ−1 k (0) of level sets of zero sections of coefficients ℓk : n T k=2 (Lk(TmM; R))∗ of δ. Since functionals ℓk are continuous, the M → F above intersection is a closed set. m∈M The set of points at which δ is kinematic is ℓ−1 k (0)∩ℓ−1 1 (T M), where we regard T M as a subbundle of F forward to check that T M is a closed subset of T ∗∗M using local trivializa- tion. (cid:3) (L1(TmM; R))∗ = T ∗∗M. It is straight- m∈M n T k=2 3. Queer Poisson brackets In this section we will construct Poisson brackets which are localizable in the sense of the following definition: Definition 3.1. A Poisson bracket on a manifold M is a bilinear operation { · , · } : C ∞(M) × C ∞(M) → C ∞(M) satisfying (i) skew-symmetry: {f, g} = −{g, f }; (ii) Jacobi identity: (cid:8){f, g}, h(cid:9) + (cid:8){g, h}, f(cid:9) + (cid:8){h, f }, g(cid:9) = 0; (iii) Leibniz rule: {f, gh} = {f, g}h + g{f, h}; for all f, g, h ∈ C ∞(M). A Poisson bracket { · , · } on M is called localizable if it has a localization, that is, a family consisting of a Poisson bracket { · , · }U on every open subset U ⊆ M, which satisfy { · , · }M = { · , · } and are compatible with restric- tions, i.e., if U ⊆ V and f, g ∈ C ∞(V ) then {f, g}V U = {f U , gU}U . If this is the case, then for any function h ∈ C ∞(M), its corresponding Hamil- tonian vector field is the operational vector field given by Xh(f )(m) := {hU , f }U (m) (3.1) for all f ∈ C ∞(U) and m ∈ U, for every open subset U ⊆ M. Remark 3.2. A version of Peetre's theorem on a Banach space E was proved in [WD73] to the effect that if a linear map T : C ∞(E) → C ∞(E) is local in the sense that supp T f ⊂ supp f for all f ∈ C ∞(E), then T is a differen- tial operator of locally finite order provided that E satisfies the condition of B∞ smoothness (existence of bump functions with Lipschitz property for all derivatives). This condition is satisfied e.g. for Hilbert spaces, but not for the Banach space of real sequences that are convergent to zero. From compatibility with restrictions it follows that operational vector fields (including Hamiltonian vector fields) are local in this sense. Thus in the case 8 QUEER POISSON BRACKETS of B∞ smooth Banach spaces they are differential operators of locally finite order. In the following we denote by V2 T ∗∗M the bundle of skew-symmetric bilinear functions on the fibers of cotangent bundle T ∗M of a Banach mani- fold M. Definition 3.3. A localizable Poisson bracket { · , · } on M is of order one at m ∈ M if there exists a skew-symmetric bounded bilinear functional Πm : T ∗ mM → R with mM × T ∗ {f, g}U (m) = Πm(f ′ m, g′ m) (3.2) open neighborhoods U of m and all f, g ∈ C ∞(U). Otherwise we say that { · , · } is queer at m ∈ M. If there exists a smooth section Π of the bundle V2 T ∗∗M satisfying (3.2) at every point m ∈ M, then we say that Π is the Poisson tensor of the Poisson bracket { · , · }. Remark 3.4. In the above definition, if the Poisson bracket is of order one at some point m ∈ M then there exists only one functional Πm satisfying (3.2), as the differentials of locally defined functions at a given point m span the whole T ∗ mM. Theorem 3.5. Let δ1 and δ2 be two commuting operational vector fields on a Banach manifold M, and define {f1, f2}U := (δ1)U (f1) (δ2)U (f2) − (δ2)U (f1) (δ1)U (f2), for all f1, f2 ∈ C ∞(U), for every open subset U ⊆ M. Then { · , · } := { · , · }M is a localizable Poisson bracket with a localization consisting of the brackets { · , · }U . If moreover δ1 and δ2 are linearly independent at some point m ∈ M, then the Poisson bracket { · , · } is queer at the point m if and only if at least one the operational vector field δ1 and δ2 is queer at m. Proof. Bilinearity and skew-symmetry of { · , · } are obvious. Jacobi identity follows from the commutativity of δ1 and δ2 just like in the case of canonical Poisson bracket on R2. This can also be seen e.g. as the special case n = 2 of [Fil85, Prop. 2]. The Leibniz rule for { · , · } follows easily from (2.1). Com- patibility with restrictions follows from the definition of operational vector fields. Now assume that δ1 and δ2 are linearly independent at m ∈ M. If none of δ1 and δ2 is queer at m, then it follows by Remark 2.3 that their values at m QUEER POISSON BRACKETS 9 satisfy (δ1)m, (δ2)m ∈ T ∗∗ T ∗ m M. Then (3.2) is satisfied if we define Πm : T ∗ mM × mM → R by Πm(µ, ν) = (δ1)m(µ) (δ2)m(ν) − (δ2)m(µ) (δ1)m(ν) for all µ, ν ∈ T ∗ mM, hence { · , · } is not queer at m ∈ M. Conversely, assume that { · , · } is not queer at m ∈ M, hence we have (3.2). Since the linear functionals (δ1)m, (δ2)m : C ∞ m (M) → R are linearly independent by hypothesis, there exist an open subset U1 ⊆ M with m ∈ U1 and a function f1 ∈ C ∞(U1) satisfying with (δ1)m(f1) = 0 and (δ2)m(f1) 6= 0. Then for every open subset U ⊆ M with m ∈ U and every f ∈ C ∞(U) we obtain {f1U ∩U1, f U ∩U1}U ∩U1(m) = ((δ2)U ∩U1(f1U ∩U1))(m) · ((δ1)U ∩U1(f U ∩U1))(m) = (δ2)m(f1) · (δ1)m(f ) hence by (3.2) (δ1)m(f ) = ((δ1)U ∩U1(f U ∩U1))(m) = 1 (δ2)m(f1) Πm((f1)′ m, f ′ m) and this shows that the operational tangent vector (δ1)m has order 1 at m. One can similarly prove that the operational tangent vector (δ2)m has order 1 at m and this completes the proof. (cid:3) One can use Theorem 3.5 and Proposition 2.4 to construct queer Poisson brackets on lp spaces for 1 ≤ p ≤ 2. Again we will present the case p = 2 in more detail. Example 3.6 (concrete queer Poisson bracket). Now let us take M = H × R. Denote points of M as (v, x). As the first operational vector field let us take δℓ from Example 2.5 acting in v variable, and for the second -- ∂ ∂x . They commute and thus by Theorem 3.5 define a queer Poisson bracket on H × R: {f, g}(v, x) := δℓ(v)f (·, x) ∂g ∂x (v, x) − ∂f ∂x (v, x)δℓ(v)g(·, x). Note that this Poisson bracket has pathological properties: it does not allow Hamiltonian formalism in the usual sense since its corresponding Hamiltonian vector fields are in general only operational vector fields, e.g. for the function h(v, x) = −x is Xh := {h, ·} = δℓ. Obviously it is not a section of T M. Since in the constructed example δℓ was a differential operator of the second order, it will not lead to an evolution flow on M. Note that the system of Hamilton equations d dt f (v(t), x(t)) = (Xhf )(v(t), x(t)) 10 QUEER POISSON BRACKETS for f ∈ C ∞(M) is not even a well posed problem. Namely for the function ρ given by (2.6) we get d dt ρ(v(t)) = 2. (3.3) Now consider the function f (v, x) = hv, wi for a fixed vector w ∈ H. One sees that f ′′ = 0 and thus Xhf = 0. Since the vector w was arbitrary, it follows that d dt v(t) = 0. As demonstrated a queer Poisson bracket does not lead to the dynamics in the usual way. However it may be possible to consider the dynamics not on the initial manifold but on some jet bundle or higher (co)-tangent bundle, see e.g. [BGG15] and references therein. Taking this into account, from the point of view of applications in physics (including classical mechanics) one should explicitly assume the existence of Poisson tensor in the definition of Poisson Banach manifold. This also ensures the existence of the map ♯ : T ∗M → T ∗∗M defined by ♯(µm) = Πm(µm, ·), µm ∈ T ∗ mM. (3.4) Definition 3.7. A Banach Poisson manifold (M, { · , · }) is a Banach man- ifold M equipped with a localizable Poisson bracket { · , · } for which there exists a Poisson tensor and the corresponding map ♯ satisfies ♯(T ∗M) ⊂ T M. (3.5) This definition is a clarification of the definition of Banach Poisson mani- folds given in [OR03, Def. 2.1], where the localizability property and the ex- istence of Poisson tensor or ♯ map were not explicitly assumed, but were assumed implicitly. In consequence all Banach Poisson manifolds considered there (including Banach Lie -- Poisson spaces) do satisfy the corrected defini- tion. The condition (3.5) on the map ♯ was introduced in [OR03] and guarantees that Hamiltonian vector fields are kinematic and it is equivalent to the bilinear functional Πm : T ∗M × T ∗M → R being separately weak∗-continuous. Acknowledgements We wish to thank Tirthankar Bhattacharyya and Andreas Kriegl for helpful discussions. The work of the first-named author was supported by a grant of the Romanian National Authority for Scientific Research and Innovation, CNCS -- UEFISCDI, project number PN-II-RU-TE-2014-4-0370. The work of the third-named author was supported by the Labex CEMPI (ANR-11-LABX- 0007-01). QUEER POISSON BRACKETS 11 References [AMR02] R. Abraham, J. E. Marsden, T. S. Ratiu: Manifolds, Tensor Analysis, and Ap- [B´00] plications. Springer-Verlag, Berlin-Heidelberg, third edition, 2002. P. Bóna: Extended quantum mechanics. Acta Physica Slovaca, 50:1 -- 198, 2000. [CP12] [CM74] [BGG15] A. J. Bruce, K. Grabowska, J. Grabowski: Higher order mechanics on graded bundles. Journal of Physics A: Mathematical and Theoretical, 48(20):205203, 2015. P. R. Chernoff, J. E. Marsden: Properties of Infinite Dimensional Hamiltonian Systems, Lecture Notes in Mathematics, volume 425. Springer Verlag, 1974. P. Cabau, F. Pelletier: Almost Lie structures on an anchored Banach bundle. J. Geom. Phys., 62(11):2147 -- 2169, 2012. G. Dito: Deformation quantization on a Hilbert space. In Noncommutative Geometry and Physics (edited by Y. Maeda, et al.), pages 139 -- 157. World Scientific, Singapore, 2005. L. D. Fadeev: Hamiltonian interpretation of inverse-scattering method. In Solitons (edited by R. K. Bullough, P. J. Caudrey). Springer-Verlag, 1980. [Fad80] [Dit05] [KM97] [Fil85] [Ida11] [FHH+01] M. Fabian, P. Habala, P. Hájek, V. M. Santalucía, J. Pelant, V. Zizler: Functional analysis and infinite-dimensional geometry. Springer Science & Business Me- dia, 2001. V. Filippov: n-Lie algebras. Sibirsk. Mat. J., 26(6):879 -- 891, 1985. C. Ida: Lichnerowicz -- Poisson cohomology and Banach Lie algebroids. Ann. Funct. Anal, 2(2):130 -- 137, 2011. A. Kriegl, P. W. Michor: The convenient setting of global analysis, volume 53. American Mathematical Society, 1997. A. Odzijewicz, T. S. Ratiu: Banach Lie -- Poisson spaces and reduction. Comm. Math. Phys., 243:1 -- 54, 2003. H. Pitt: A note on bilinear forms. Journal of the London Mathematical So- ciety, 1(3):174 -- 180, 1936. R. A. Ryan: Introduction to tensor products of Banach spaces. Springer Sci- ence & Business Media, 2002. J. C. Wells, C. R. DePrima: Local automorphisms are differential operators on some Banach spaces. Proceedings of the American Mathematical Society, 40(2):453 -- 457, 1973. [WD73] [Rya02] [OR03] [Pit36]
1204.0712
1
1204
2012-04-03T15:33:51
Creator-annihilator domains and the number operator
[ "math.FA", "math-ph", "math-ph" ]
We show that for the bosonic Fock representation in infinite dimensions, the maximal common domain of all creators and annihilators properly contains the domain of the square-root of the number operator.
math.FA
math
Creator-annihilator domains and the number operator P. L. Robinson Abstract We show that for the bosonic Fock representation in infinite dimensions, the maximal common domain of all creators and annihilators properly contains the domain of the square-root of the number operator. Introduction A standard construction of the bosonic Fock representation of a complex Hilbert space V is in terms of creators and annihilators on the symmetric algebra SV in which the number operator N scales elements by homogeneous degree; the symmetric algebra is completed relative to a canonical inner product and these various operators are extended to maximal domains in the resulting Fock space S[V ]. 1 It is well-known that all creators and annihilators are defined on the 2 . A natural question (raised on page 16 of [2] by domain of the square-root N Berezin, among others; compare page 65 of [4]) is whether the domain of N coincides with the maximal common domain of all creators and annihilators. Here, we demonstrate that the answer to this question is affirmative when V is finite-dimensional but negative when V is infinite-dimensional. 1 2 As noted above, the specific question that is answered in this paper ap- pears in [2]. Among many standard references concerning the bosonic Fock representation, we cite [1] and [3]. The particular approach taken here (in- volving the full antidual of the symmetric algebra) was introduced in [6] as a means to establishing a generalized version of the classic Shale theorem 1 on the implementation of symplectic automorphisms, independently of the generalization previously presented in [5]. Fock-lore We begin by recalling certain familiar elements of Fock-lore, pertaining to the construction of bosonic Fock space and the various operators defined therein. For traditional accounts, see a standard text such as [1] or [3]; for an account in line with the present paper, see [6]. To be explicit, let V be a complex Hilbert space. Extend its inner prod- uct h··i to the symmetric algebra SV = Ln≥0 S nV by declaring that the homogeneous summands (S nV : n ≥ 0) be perpendicular and that if n ≥ 0 and x1, . . . , xn, y1, . . . , yn ∈ V then hx1 · · · xny1 · · · yni = X nY hxjyp(j)i p j=1 where p runs over all permutations of {1, . . . , n}; the resulting complex Hilbert space completion is (by definition) the bosonic Fock space S[V ] = Ln≥0 S n[V ]. For our purposes, it is convenient to introduce also the full antidual SV ′ comprising all antilinear functionals SV → C. This full antidual SV ′ is naturally a commutative, associative algebra under the product defined by Φ, Ψ ∈ SV ′ =⇒ [ΦΨ](θ) = (Φ ⊗ Ψ)(∆θ) where the coproduct ∆ : SV → SV ⊗ SV arises when the canonical isomor- phism S(V ⊕ V ) ≡ SV ⊗ SV follows the homomorphism SV → S(V ⊕ V ) induced by the diagonal map V → V ⊕ V . The inner product on SV engen- ders an algebra embedding SV → SV ′ : φ 7→ h·φi and S[V ] is identified with the subspace of SV ′ comprising all bounded anti- linear functionals. Let v ∈ V . The creator c(v) is defined initially on SV as the operator of multiplication by v: φ ∈ SV =⇒ c(v)φ = vφ. 2 The annihilator a(v) is defined initially on SV as the linear derivation that kills the vacuum 1 ∈ C = S0V and sends w ∈ V = S1V to hvwi: thus, if v1, . . . , vn ∈ V then a(v)(v1 · · · vn) = nX hvvjiv1 · · ·cvj · · · vn j=1 where the circumflex b· calculation, c(v) and a(v) are mutually adjoint: signifies omission. As may be verified by direct φ, ψ ∈ SV =⇒ hψc(v)φi = ha(v)ψφi. Accordingly, the creator c(v) and annihilator a(v) extend to SV ′ by antidu- ality: if Φ ∈ SV ′ and ψ ∈ SV then and [c(v)Φ](ψ) = Φ[a(v)ψ] [a(v)Φ](ψ) = Φ[c(v)ψ]. Finally, the corresponding (mutually adjoint) operators in Fock space S[V ] ⊂ SV ′ are defined by restriction to the (coincident) natural domains D[c(v)] = {Φ ∈ S[V ] : c(v)Φ ∈ S[V ]} and D[a(v)] = {Φ ∈ S[V ] : a(v)Φ ∈ S[V ]}. The number operator N is defined initially on SV by the rule n ≥ 0 =⇒ NS nV = nI and extends to SV ′ by antiduality: Φ ∈ SV ′, ψ ∈ SV =⇒ [NΦ](ψ) = Φ[Nψ]. The number operator in S[V ] ⊂ SV ′ is defined by restriction to the natural domain D[N] = {Φ ∈ S[V ] : NΦ ∈ S[V ]} which may be identified in terms of the decomposition S[V ] = Ln≥0 S n[V ] as D[N] = {X Φn ∈ S[V ] : X knΦnk2 < ∞}. n≥0 n≥0 3 We remark that the number operator N in S[V ] is selfadjoint (indeed, pos- itive): Ln≥0 S n[V ] is its spectral decomposition, whence powers of N are readily described in concrete terms; in particular, D[N 1 2 ] = {Φ ∈ S[V ] : X nkΦnk2 < ∞}. n≥0 Theorems Having established sufficient background, we now proceed to our primary 2 ] to the maximal common domain of all creators task: that of relating D[N and annihilators in Fock space. 1 Let u ∈ V be (for convenience) a unit vector: the unitary decomposition V = CuL u⊥ induces (for each n ≥ 0) a unitary decomposition S nV = M {S p(Cu) ⊗ S q(u⊥)}. Decomposing φ ∈ S nV as p+q=n φ = X p+q=n up ⊗ ψq we note (by perpendicularity) that kφk2 = X kupk2kψqk2 = X p! kψqk2 p+q=n p+q=n and that kc(u)φk2 = X kup+1k2kψqk2 = X (p + 1)! kψqk2 p+q=n p+q=n whence it follows that kc(u)φk2 ≤ (n + 1)kφk2. This inequality continues to apply when φ lies in the closure S n[V ]; in par- ticular, S n[V ] ⊂ D[c(u)]. Now, if Φ ∈ D[N 2 ] then 1 kc(u)Φk2 = X kc(u)Φnk2 ≤ X (n + 1)kΦnk2 = kN 1 2 Φk2 + kΦk2 n≥0 n≥0 whence Φ ∈ D[c(u)] = D[a(u)]. Lifting the convenient hypothesis that u ∈ V be a unit vector, we have justified the following result. 4 Theorem 1 If u ∈ V then D[N 1 2 ] is contained in D[c(u)] = D[a(u)]. Thus, D[N 1 2 ] is contained in the maximal common domain of all creators and annihilators in Fock space. We now consider the reverse containment in case V is finite-dimensional, with (u1, . . . , um) a unitary basis. If the integers n1, . . . , nm ≥ 0 have sum n then for each j ∈ {n1, . . . , nm} c(uj)a(uj)(un1 1 · · · unm m ) = nj(un1 1 · · · unm m ) so mX j=1 c(uj)a(uj)(un1 1 · · · unm m ) = n(un1 1 · · · unm m ). By linearity, it follows that mX j=1 c(uj)a(uj) φ = nφ whenever φ ∈ S nV and indeed whenever φ ∈ S n[V ] by the boundedness of creators and annihilators on homogeneous elements of Fock space. Now, if Φ ∈ D[a(u1)] ∩ · · · ∩ D[a(um)] then ∞ > mX j=1 ka(uj)Φk2 = mX X j=1 n≥0 ka(uj)Φnk2 whence (valid) passage of annihilators across the inner product as creators yields ∞ > X hΦn mX c(uj)a(uj)Φni = X nkΦnk2 which places Φ in D[N 1 2 ]. This justifies the following result. n≥0 j=1 n≥0 Theorem 2 If (u1, . . . , um) is a unitary basis for the finite-dimensional V then 1 D[N 2 ] = D[a(u1)] ∩ · · · ∩ D[a(um)] and if Φ lies in this domain then kN 1 2 Φk2 = mX j=1 ka(uj)Φk2. In particular, if V is finite-dimensional then D[N 1 2 ] coincides with the maximal common domain of all creators and annihilators in Fock space. 5 The case in which V is infinite-dimensional is different. To see this, let Φ = Pn>0 Φn ∈ SV ′ be defined by n > 0 =⇒ Φn = λn(un)n/n! where (λn : n > 0) is a complex sequence and where the unit vectors (un : n > 0) in V are perpendicular. From n > 0 =⇒ kΦnk2 = λn2k(un)nk2/(n!)2 = λn2/n! it follows that and that Φ ∈ S[V ] ⇐⇒ X n>0 λn2 n! < ∞ Φ ∈ D[N 1 2 ] ⇐⇒ X n>0 λn2 (n − 1)! < ∞. Further, let v ∈ V : if n > 0 then a(v)Φn = hvuniλn(un)n−1/(n − 1)! whence ka(v)Φk2 = X hvuni2 n>0 Now, if (λn : n > 0) is chosen so that λn2 (n − 1)! . X λn2/n! < ∞ = X λn2/(n − 1)! n>0 n>0 then Φ ∈ S[V ] \ D[N by K > 0 then 1 2 ] while if also (λn2/(n − 1)! : n > 0) is bounded above ka(v)Φk2 ≤ X hvuni2K ≤ Kkvk2 n>0 which places Φ in D[a(v)] = D[c(v)]. Of course, these conditions are easily satisfied: for example, when n > 0 simply take λn2/(n − 1)! = 1/n. As the vector v ∈ V is arbitrary, the following result is justified. Theorem 3 If V is infinite-dimensional then D[N in the maximal common domain of all creators and annihilators. 2 ] is properly contained 1 Without proof, we remark that similar results hold for higher powers of 2 ] is contained in the maximal the number operator: thus, if k > 0 then D[N common domain of all degree k polynomials in creators and annihilators, containment being strict precisely when V is infinite-dimensional. k 6 References [1] J. C. Baez, I. E. Segal and Z. Zhou, Introduction to Algebraic and Con- structive Quantum Field Theory. Princeton University Press (1992). [2] F. A. Berezin, The Method of Second Quantization. Academic Press (1966). [3] O. Bratteli and D. W. Robinson, Operator Algebras and Quantum Statis- tical Mechanics II. Springer-Verlag (1981). [4] J. T. Ottesen, Infinite Dimensional Groups and Algebras in Quantum Physics. Springer-Verlag (1995). [5] S. M. Paneitz, J. Pedersen, I. E. Segal and Z. Zhou, Singular Operators on Boson Fields as Forms on Spaces of Entire Functions on Hilbert Space. J. Functional Analysis 100 (1991) 36-58. [6] P. L. Robinson, The bosonic Fock representation and a generalized Shale theorem. University of Florida preprint (1998); arXiv:1203.5841v1. Department of Mathematics University of Florida Gainesville, FL 32611 e-mail: [email protected] 7
1609.08585
3
1609
2017-11-04T12:47:25
Finite-Dimensional Representations constructed from Random Walks
[ "math.FA", "math.GR", "math.PR" ]
Given a $1$-cocycle $b$ with coefficients in an orthogonal representation, we show that any finite dimensional summand of $b$ is cohomologically trivial if and only if $\| b(X_n) \|^2/n$ tends to a constant in probability, where $X_n$ is the trajectory of the random walk $(G,\mu)$. As a corollary, we obtain sufficient conditions for $G$ to satisfy Shalom's property $H_{\mathrm{FD}}$. Another application is a convergence to a constant in probability of $\mu^{*n}(e) -\mu^{*n}(g)$, $n\gg m$, normalized by its average with respect to $\mu^{*m}$, for any finitely generated amenable group without infinite virtually Abelian quotients. Finally, we show that the harmonic equivariant mapping of $G$ to a Hilbert space obtained as an $U$-ultralimit of normalized $\mu^{*n}- g \mu^{*n}$ can depend on the ultrafilter $U$ for some groups.
math.FA
math
FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS ANNA ERSCHLER AND NARUTAKA OZAWA ABSTRACT. Given a 1-cocycle b with coefficients in an orthogonal representation, we show that every finite dimensional summand of b is cohomologically trivial if and only if kb(Xn)k2/n tends to a constant in probability, where Xn is the trajectory of the random walk (G, µ). As a corollary, we obtain sufficient conditions for G to satisfy Shalom's property HFD. Another application is a convergence to a constant in proba- bility of µ∗n(e) − µ∗n(g), n ≫ m, normalized by its average with respect to µ∗m, for any finitely generated infinite amenable group without infinite virtually abelian quotients. Finally, we show that the harmonic equivariant mapping of G to a Hilbert space obtained as an U -ultralimit of normalized µ∗n − gµ∗n can depend on the ultrafilter U for some groups. 1. INTRODUCTION Convention. Throughout the paper, G is a compactly generated locally compact group with a distinguished relatively compact symmetric subset Q which contains an open gen- erating neighborhood e of G, and µ is a symmetric probability measure on G that satisfies the following conditions: • µ is absolutely continuous with respect to the Haar measure m, • inf{ dµ • R xd dm (x) : x ∈ Q} > 0, G dµ(x) < ∞ for all d. Here xG := min{n : x ∈ Qn} (except that eG := 0). Note that · G is a length function, that is, it satisfies xG = x−1G and xyG ≤ xG + yG. Put BG(r) := {x ∈ G : xG ≤ r}. Formulation of the results. Throughout the paper, we will work with real Hilbert spaces and orthogonal representations. This is purely for our convenience and all results (but not the proofs) hold true for complex Hilbert spaces and unitary representations (except that the statement of Theorem 2.4 has to be slightly modified), because any complex Hilbert space HC is also a real Hilbert space with the real inner product (v, w) 7→ ℜhv, wiHC , and any 1-cocycle (defined below) with coefficients in a unitary representation can be regarded as the one with coefficients in the corresponding orthogonal representation. Let π : G y H be an orthogonal representation on a real Hilbert space H. Recall that a 1-cocycle (or simply a cocycle) is a continuous map b : G → H which satisfies the 1- cocycle identity: b(gx) = b(g) + πgb(x) for all g, x ∈ G. It is a 1-coboundary if there is v ∈ H such that b(x) = v − πxv for all x ∈ G. We note that b is a 1-coboundary if and The work of the fist named author is partially supported by the ERC grant GroIsRan. The work of the second named author is partially supported by JSPS KAKENHI Grant Number 26400114. He also expresses his gratitude to the organizers of the programs "Classification of operator algebras: complexity, rigidity, and dynamics", "Von Neumann Algebras", the Mittag-Leffler Institute, the Hausdorff institute and Ecole Normale, Paris for the hospitality during his visits. The authors would like to thank the referee for helpful remarks. 1 2 ANNA ERSCHLER AND NARUTAKA OZAWA only if it is bounded on G (Proposition 2.2.9 in [2]). Every cocycle b satisfies that b(e) = 0 and kb(x) − b(y)k = kb(x−1y)k ≤ kbkQx−1yG, where kbkQ := supg∈Q kb(g)k < ∞. A cocycle b is said to be µ-harmonic (or simply harmonic) if R b(gx) dµ(x) = b(g) for all g, or equivalentlyR b(x) dµ(x) = 0. Any cocycle b gives rise to an affine isometric action A : G×H → H by A(g, v) = πgv +b(g) (see Chapter 2 in [2]). Conversely, for any (affine) isometric action on a Hilbert space and a point v ∈ H, the map b(g) = A(g, v)− v defines a 1-cocycle, and harmonicity of this cocycle is same as harmonicity of the orbit map g 7→ A(g, v). Under an appropriate assumption on the decay of a non-degenerate measure µ, it is known that a compactly generated locally compact group G admits a non-zero µ- harmonic cocycle with respect to some orthogonal representation if and only if G does not satisfy Kazhdan' property (T). Existence of a non-zero harmonic cocycle on groups which do not satisfy property (T) is proved by Mok ([25, Cor. 0.1]), Korevaar and Schoen [22, Thm 4.1.2] for finitely presented groups (and not discrete definition of harmonicity) and in general case (and discrete definition of harmonicity) by Shalom in [32, Thm 6.1]. We will give somewhat more constructive proof of this fact in Section 4. See also Gromov [14, Section 3.6], [15, Section 7A] Fisher and Margulis [11], Lee and Peres [23, Thm 3.8], Ozawa [29] as well as the book by Bekka, de la Harpe, and Valette [2] for a non-exhaustive list of references about this result. We say that a 1-cocycle b is finite-dimensional if the π(G)-invariant subspace span b(G) is finite-dimensional. If H = Li Hi is some orthogonal decomposition of H into π(G)- invariant subspaces, then b = Li PHi b is a decomposition of b into 1-cocycles PHi b (with respect to πHi ). We call each PHi b a summand of b. We say that such summand is cohomologically trivial if it is a 1-coboundary. Given a probability measure µ on G, let Xn denote the trajectory of the random walk (G, µ), that is, Xn = s1s2 ··· sn where increments si ∈ G are independent and chosen with respect to µ. The corresponding probability measure and its expectation are denoted by P and E. The value of a Hilbert valued µ-harmonic 1-cocycle along a trajectory of the random walk (G, µ) is a martingale, and therefore E(cid:2)kb(Xn)k2(cid:3) = n Xk=1 That is, the expected value 1 n For any (not necessarily harmonic) 1-cocycle b, the expected value 1 n limit (see Lemma 2.2). Theorem A below characterizes the case when the random variable 1 E(cid:2)kb(Xk)k2 − kb(Xk−1)k2(cid:3) = n E(cid:2)kb(X1)k2(cid:3). E(cid:2)kb(Xn)k2(cid:3) is equal to a constant, not depending on n. E(cid:2)kb(Xn)k2(cid:3) has a nkb(Xn)k2 tends to a constant. Theorem A. Let G be a compactly generated locally compact group with a probability measure µ on G as in Convention. Let b : G → H be a 1-cocycle. Then the following conditions are equivalent: (1) Any finite-dimensional summand of b is cohomologically trivial. (2) 1 nb(Xn)2 tends to a constant in probability. Now assume moreover that b is harmonic and put c = RG kb(x)k2 dµ(x). Then the limit β := lim n→∞ 1 2c2 E(cid:2)(cid:12)(cid:12) kb(Xn)k2 n 2(cid:3) − c(cid:12)(cid:12) always exists, and β = 0 if and only if (1) and (2) hold. If β 6= 0, then b has a cohomo- logically non-trivial finite-dimensional summand of dimension ≤ 1/β. FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 3 A more precise version of Theorem A will be given in Theorem 2.4, where we describe the limit distribution of kb(Xn)k/√n. This theorem has the following corollary: Corollary. Let b be a harmonic cocycle. Then, b is a direct sum of (possibly infinitely many) finite-dimensional cocycles if and only if lim supn every c > 0. P(kb(Xn)k < c√n) > 0 for Recall that a group G is said to have Shalom's property HFD if every orthogonal rep- resentation π with non-zero reduced cohomology group H 1(G, π) contains a non-zero finite-dimensional subrepresentation. In Corollary 2.5 we show that G satisfies Shalom's property HFD if at least one of the two following conditions hold: either lim inf n kµ∗n − µ∗(1+δ)nk1 < 2 for some δ > 0 or lim supn µ∗n(BG(c√n)) > 0 for all c > 0. Theorem A and its corollaries develop the argument from [29]. While the main result of [29] is a new proof of Gromov's polynomial growth theorem, the paper also provides a more general criterion for the property HF D for a finitely generated group in terms of convolutions of random walks is given in Section 4 of [29]. It is shown in [10] that wreath products of Z with finite groups satisfy the assumption of that criterion, providing examples of groups of super-polynomial growth where the criterion applies. The assumption of the criterion from Section 4 in [29] uses shifted convolution, and it is not clear whether this assumption is defined by an unmarked Cayley graph of G. Assume that (G, µ) is a simple random walk on G, that is, µ is equidistributed on a finite generating set of G. The conditions of (1) as well as of (2) of Corollary 2.5 are clearly defined by the unmarked Cayley graph of G. We do not know any group which satisfies the assumption of (1) or of (2) of Corollary 2.5 and for which we know that it violates the assumption of Section 4 of [29]. But the conditions of Corollary 2.5 are easier to check than the assumption from [29]. For example, it is easily applicable to solvable Baumslag–Solitar groups, lamplighter groups Z ⋉ LZ F with F finite, or to polycyclic groups obtained as extension of Z2 by M ∈ SL(2, d) with eigenvalues of absolute value 6= 1. See Section 3 for more examples. We do not know any group which satisfies Shalom's property and does not satisfy the assumption of Corollary 2.5. Given a not necessarily harmonic cocycle b on a group without property (T), a harmonic cocycle can be obtained taking averages of b (see Mok, Korevaar Schoen, Shalom [22, 25, 31], and in particular this can be achieved averaging with respect to a probability measure µ (see e.g. Gromov, Lee–Peres [14,23]). In Section 4 we study the cocycles bµ,U , constructed as a ultralimit in ℓ2(G) of normalized µ∗n − gµ∗n on a finitely generated amenable group G. Kesten's criterion [21] (see also [1]) implies that µ∗n is a sequence of almost invariant vectors in ℓ2(G), and one can moreover show (see Theorem 4.3) that the limit is a harmonic 1-cocycle. Applying Theorem A to this 1-cocycle, one obtains Theorem B. Let G be a finitely generated infinite amenable group without virtually abelian infinite quotients. Let µ be a finitely-supported symmetric non-degenerate probability mea- sure. Then (µ∗2n(e) − µ∗2n(X2m))/α(m, n) tends to a constant in probability µ∗2m as m → ∞ and n ≫ m. Here α(m, n) = µ∗2n(e) − µ∗2n+2m(e) is the average of µ∗2n(e) − µ∗2n(g) with respect to µ∗2m. Namely Take n much larger than m. Observe that a group is amenable if and only if µ∗2n(g)/µ∗2n(e) is close to 1 in probability with respect to µ∗2m. Theorem B gives a sufficient condition for the concentration of the second order term of µ∗2n. lim m→∞ lim sup n→∞ = 0. E(cid:12)(cid:12)(cid:12)(cid:12) µ∗2n(e) − µ∗2n(X2m) µ∗2n(e) − µ∗2n+2m(e) − 1(cid:12)(cid:12)(cid:12)(cid:12) 4 ANNA ERSCHLER AND NARUTAKA OZAWA Theorem B applies in particular to any finitely generated amenable torsion group (such as Grigorchuk groups Gw [12]) or to any finitely generated amenable simple group (such as commutator full topological groups of minimal shifts on Z (which are simple by a result of Matui [24] and amenable by a result of Juschenko–Monod [19]), or to simple groups of intermediate growth constructed recently by Nekrashevych [26]. If µ is equidistributed on a finite generating set of G, then the assumption of Theorem B depends only on the unmarked Cayley graph of (G, µ). In particular, the theorem gives a necessary condition for an amenable group to be simple in terms of unmarked Cayley graphs. In general, it is known that the property of being simple can not be defined by the unmarked Cayley graphs, as it is shown by Burger and Mozes [4] (their examples are isometric to product of two trees and they are non-amenable). It is to our knowledge an open problem whether a property of being a torsion group can be verified geometrically. Geometric group theory tries to recover properties of a group from the word metrics of this group. Given a group G, generated by a finite set S, its action on a metric space X and a point x0 ∈ X, the group G is equipped with two metrics: the word metric dG,S(g, h) as well as dX,x0(g, h) = dX (gx0, hx0). It seems interesting to study which properties of the action, or of the group G, can be recovered from these two metrics. Theorem A as well as Corollary 2.5 provide examples of such situation, for X being a Hilbert space and a group G acting by affine transformations of X. Fix a non-principal ultrafilter U on the natural numbers N. Let bp,q µ,U be the mapping to a vector space equipped with a metric, constructed as U ultralimit of normalized (µ∗n)q − g(µ∗n)q, considered as elements of ℓp(G) (see Section 5). This means that we divide g(µ∗n)q − (µ∗n)q by the lp norm of this expression, considered as a function on g, and then we take the ultralimit with respect to U . By the construction, the lp norm of bp,q U,µ,G is one. We recall that any ultralimit of Hilbert spaces is a Hilbert space, so that for p = 2 and any q ≥ 0 we obtain a cocycle with respect to some orthogonal representation of H. In particular, for q = 1 and p = 2 , bp,q µ,U coincides up to a multiplicative constant with the harmonic cocycle bµ,U , studied in the proof of Theorem B in Section 4. In general, for p 6= 2, we obtain a cocycle with respect to some isometric representation on an abstract Lp-space. µ,U , p ≥ 1, q ≥ 0 (in particular, the In Theorem C below we show that the cocycles bp,q harmonic cocycle bµ,U ) can depend on the choice of a non-principal ultrafilter U . Theorem C. Take p = 1 or 2 and q = 0, 1, or 2. For any D ≥ 2 there exist torsion groups G1, G2, . . . , GD such that the following holds. Consider finitely supported symmetric non- degenerate measures µi on Gi and put G = QD j=1 µi. For each j = 1, . . . , D there exists a non-principal ultrafilter U such that the limiting cocycle bp,q µ,U factors through G ։ Gi. j=1 Gi and µ = QD Theorem C shows in particular that there exist at least D mutually distinct limiting cocycles among {bp,q µ,U : U}, and at least D mutually distinct subgroups among possible kernels of such cocycles. Such groups G admit g1, g2 ∈ G such that the ratio (µ∗2n(e) − µ∗2n(g1))/(µ∗2n(e) − µ∗2n(g2)) does not have a limit as n → ∞. The groups Gi are constructed as piecewise automatic groups [9], they can be chosen to be of sub-exponential word growth, but in such a way that for each j the group Gj is in some sense very close to a non-amenable group on some scale while on this particular µ,U is mainly from Gj scale it does not happen to other Gk, j 6= k. The contribution to bp,q on this scale, and the kernel of bp,q µ,U containsQk6=j Gk. FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 5 The kernels of cocycles bp,q µ,U are particular cases of what we call ℓp-thin subgroups: this is a natural family of subgroups, related to the shifts (µ∗n)q (see Definition 5.1), which for p = 2, q = 1 is related to amenability, for p = q = 1 to Poisson-Furstenberg boundary and for q = 0, p ≥ 1 to growth of groups (see Lemma 5.6), these groups in some situation may depend on p (see Example 5.9) and on the measure µ (see Remark 5.10). Since the group G in the statement of the theorem is a torsion group, it does not admit a virtual quotient to an infinite cyclic group. In particular, taking p = 2 we can apply the conclusion of Theorem B to (G, µ) to claim that µ∗n(e) − µ∗n(g), normalised by its average α(m, n) is close to a constant in probability µ∗m, for n ≫ m. In other words, for each n ≫ m µ∗m is concentrated on a set where normalized µ∗n(e) − µ∗n(g) is close to its mean value, but in view of C these sets may depend essentially on n. We are grateful to Pierre de la Harpe for comments on the preliminary version of this paper. 2. HARMONIC COCYCLES AND FINITE-DIMENSIONAL SUMMANDS We now recall from Sections 4 and 5 in [16] that the space Z 1(G, π) of 1-cocycles is a Hilbert space under the norm kbkL2(µ) := (cid:16)ZG kb(x)k2 dµ(x)(cid:17)1/2 , and it decomposes into an orthogonal direct sum of approximate 1-coboundaries and µ- harmonic 1-cocycles. We will say b is normalized when kbkL2(µ) = 1. Lemma 2.1. The space Z 1(G, π) of 1-cocycles is a Hilbert space with respect to the norm k · kL2(µ). Moreover the norms k · kL2(µ) and k · kQ are equivalent. Proof. We observe that Z 1(G, π) is a Banach space w.r.t. the norm k · kQ (see [2, Chap- G dµ(x))1/2kbkQ. The other side inequality follows, via ter 3]), and that kbkL2(µ) ≤ (R x2 Open Mapping Theorem, from the fact that any measurable locally integrable 1-cocycle into a separable Hilbert space is automatically continuous modulo a null set. However, following [16], we give a more direct proof here. Take an open generating neighborhood U of e such that U ⊂ Q and an open neighborhood V of e such that V 2 ⊂ U . We observe dµ∗2 that (R kb(x)k2 dµ∗2(x))1/2 ≤ 2kbkL2(µ) and ε := inf x∈UV dm (x) > 0. Thus, for every g ∈ U one has kb(g)k2 = m(V )−1ZV kb(gx) − πgb(x)k2 dm(x) ≤ 2m(V )−1[ZgV kb(x)k2 dm(x) +ZV kb(x)k2 dm(x)] ≤ 4ε−1m(V )−1ZUV kb(x)k2 dµ∗2(x) ≤ 16ε−1m(V )−1kbk2 L2(µ). Since there is N ∈ N such that Q ⊂ U N , this proves that the norms k · kL2(µ) and k · kQ are equivalent, and that Z 1(G, π) is a Hilbert space w.r.t. the norm k · kL2(µ). (cid:3) The reduced 1-cohomology group H 1(G, π) := Z 1(G, π)/B1(G, π) is defined to be the space Z 1(G, π) of 1-cocycles modulo the closure of the subspace B1(G, π) of 1- coboundaries. We note that B1(G, π) = B1(G, π) if π is finite-dimensional, by Theo- rem 1 in [16]. See Chapter 3 in [2] for an introduction to first reduced cohomology groups. 6 Thus, ANNA ERSCHLER AND NARUTAKA OZAWA Z 1(G, π) = B1(G, π) ⊕ B1(G, π)⊥ and H 1(G, π) ∼= B1(G, π)⊥. We observe that b ∈ Z 1(G, π) belongs to B1(G, π)⊥ if and only if it is µ-harmonic in the sense R b(x) dµ(x) = 0 or equivalently R b(gx) dµ(x) = b(g) for all g ∈ G. Indeed, this follows from the identities b(x−1) + π−1 x b(x) = b(e) = 0 and Z hb(x), v − πxvi dµ(x) = 2hZ b(x) dµ(x), vi. We note that every summand of a µ-harmonic 1-cocycle is µ-harmonic and that every non-zero µ-harmonic 1-cocycle is not a 1-coboundary. We recall the general fact about orthogonal representations. Let (π,H) be an orthogonal representation of G and put T0 := E(cid:2)π(X1)(cid:3) = Z π(g) dµ(g). Then, T0 is a self-adjoint contraction on the Hilbert space H such that T k 0 = E(cid:2)π(Xk)(cid:3) for every k. By strict convexity of a Hilbert space, a vector v ∈ H satisfies T0v = v if and only if πgv = v for µ-a.e. g, which is equivalent to that v is π(G)-invariant. Thus by spectral theory, the operators 1 k=0 π(Xk)(cid:3) converge in strong operator topology to the orthogonal projection P0 onto the subspace of π(G)-invariant vectors. One moreover has convergence in probability 0 = E(cid:2) 1 n Pn−1 n Pn−1 k=0 T k → 0. Indeed, to prove it, one may assume P0 = 0 and in this case π(Xk)v − P0vk k P 1 n n−1 Xn=0 1 n n−1 Xn=0 E(cid:2)k π(Xk)vk2(cid:3) = 1 n2 n−1 Xk,l=0 hT k−l 0 v, vi → 0. Lemma 2.2. For every b ∈ Z 1(G, π) = B1(G, π) ⊕ B1(G, π)⊥, one has lim n 1 n L2(µ), E(cid:2)kb(Xn)k2(cid:3) = kbharmk2 E(cid:2)kb(Xn)k2(cid:3) > 0. where bharm is the B1(G, π)⊥ summand in the above decomposition. In particular, b is nonzero in H 1(G, π) if and only if lim 1 n Proof. Let T0 := R π(g) dµ(g). If c ∈ B1(π,H) is a 1-coboundary, c(x) = v − πxv, then for every n one has 1 n E(cid:2)kc(Xn)k2(cid:3) = 2 nh(1 − T n 0 )v, vi ≤ 2h(1 − T0)v, vi = kck2 L2(µ). Since c 7→ E(cid:2)kc(Xn)k2(cid:3) is norm-continuous by Lemma 2.1, the above inequality holds for all c ∈ B1(G, π). Hence, for any c ∈ B1(G, π), by approximating it by cm ∈ B1(G, π), one has lim sup n 1 n E(cid:2)kc(Xn)k2(cid:3) = lim sup n E(cid:2)k(c − cm)(Xn)k2(cid:3) ≤ kc − cmk2 L2(µ) → 0. 1 n Now let b = c + bharm ∈ B1(G, π) + B1(G, π)⊥ be given. Note that since bharm is µ∗n-harmonic, it is orthogonal to c in L2(µ∗n). Consequently, one has lim n 1 n E(cid:2)kb(Xn)k2(cid:3) = lim n 1 n E(cid:2)kc(Xn)k2 + kbharm(Xn)k2(cid:3) = kbharmk2 L2(µ). (cid:3) FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 7 It is not clear whether 1 is the case for any µ-harmonic 1-cocycle b (cf. Footnote 2 in [23]). n2 E(cid:2)kb(Xn)k4(cid:3) is bounded for every 1-cocycle b. However, it Lemma 2.3. For every d, one has sup sup n b 1 nd E(cid:2)kb(Xn)k2d(cid:3) < ∞, 1 nd where the supremum runs over all normalized µ-harmonic 1-cocycles b. Proof. We fix a universal orthogonal representation (π,H) and consider the operators Un from the space of µ-harmonic cocycles into L2d(µ∗n;H), given by Unb = n−1/2b. Since G (cid:3)1/2d kbkQ kUnbk = ( E(cid:2)kb(Xn)k2d(cid:3))1/2d ≤ n1/2 E(cid:2)X12d i=1 ai)2d ≤ n2d−1Pn i=1 a2d i (by the Holder inequality (Pn for ai ≥ 0), the operators Un are bounded by Lemma 2.1. The lemma claims that Un's are uniformly bounded. For this, by Principle of Uniform Boundedness, it suffices to show supn kUnbk < ∞ for each b. (The use of PUB can be avoided if one does the following proof more meticulously.) We E(cid:2)kb(Xn)k2d(cid:3) ≤ (2d − 1)!! for each normalized harmonic in fact prove that lim supn cocycle b, by induction on d. Here (2d− 1)!! = Qd k=1(2k− 1). The case d = 1 is clear. By induction hypothesis and the Cauchy–Schwarz inequality when k is odd, we may assume that there is C > 0 such that E(cid:2)kb(Xn)kk(cid:3) ≤ Cnk/2 for all k ≤ 2(d − 1). It follows that 1 nd E(cid:2)kb(Xn)k2d(cid:3) = ZZ kb(x) − b(y)k2d dµ∗n−1(x)dµ(y) = ZZ (kb(x)k2 − 2hb(x), b(y)i + kb(y)k2)d dµ∗n−1(x)dµ(y) = ZZ kb(x)k2d +(cid:18)d 1(cid:19)kb(x)k2(d−1)kb(y)k2 + 4(cid:18)d 2(cid:19)kb(x)k2(d−2)hb(x), b(y)i2 dµ∗n−1(x)dµ(y) + C′n(2d−3)/2 n ≤ E(cid:2)kb(Xn−1)k2d(cid:3) + (d + 2d(d − 1)) · (2d − 3)!! · nd−1 + C′nd−3/2 ≤ ··· ≤ = (2d − 1)!! · nd + o(nd), ((2d − 1)!! · dnd−1 + C′kd−3/2) Xk=1 where C′ is some constant depending on d but not on n. This finishes the proof. (cid:3) We start the proof of Theorem A. Recall that the tensor product Hilbert space H ⊗ H is canonically identified with the space of Hilbert–Schmidt operators S2(H) on H via v′ ⊗ v ↔ Sv′⊗v, where Sv′⊗v(u) = hu, viv′. Under this identification, the operators πg⊗πg on H⊗H act on S2(H) by conjugation Ad πg : S 7→ πgSπ∗g . Every Hilbert–Schmidt operator is compact and every compact self-adjoint operator S has a unique spectral decomposition S = Pi λiEi where λi ∈ R are the non-zero eigenvalues of S and Ei are the finite-rank orthogonal projections onto the corresponding eigenspaces. If v ∈ H ⊗ H is (π ⊗ π)(G)- invariant, then Sv is Ad π(G)-invariant and so are the spectral projections Ei's, which means that EiH are finite-dimensional π(G)-invariant subspaces. Now let us consider a 1-cocycle b : G → H and put w := Z (b ⊗ b)(x) dµ(x) ∈ H ⊗ H and T := Z πg ⊗ πg dµ(g). 8 ANNA ERSCHLER AND NARUTAKA OZAWA Then, T is a self-adjoint contraction on H ⊗ H, which is positivity preserving as an op- erator on S2(H). By the previous discussion, 1 k=0 T k converges in strong operator topology to the orthogonal projection P from H ⊗ H onto the subspace of (π ⊗ π)(G)- k=0 T kw converges to P w in norm and SP w is a invariant vectors. positive Hilbert–Schmidt operator which is Ad π(G)-invariant. For any π(G)-invariant closed subspace K ⊂ H, one has n Pn−1 n Pn−1 In particular, 1 PKSP wPK = S(PK⊗PK)P w = SP (PK⊗PK)w = SP wK, where wK = R (bK ⊗ bK)(x) dµ(x) for the cocycle bK = PKb. If b is finite-dimensional, then the trace Tr is norm-continuous and Tr(SP w) = Tr(lim n 1 n n−1 Xk=0 ST kw) = Tr(Sw) = kbk2 L2(µ). In general, one has the spectral decomposition λiEi SP w = Xi (∗) where λ1, λ2, . . . is a finite or infinite sequence of strictly positive numbers and EiH's are finite-dimensional π(G)-invariant subspaces. Thus for bi := Eib and b∞ := b − (Pi bi), one has the direct sum decomposition b = b∞ + Pi bi. We claim that each 1-cocycle bi, i 6= ∞, is nonzero and that b∞ is weakly mixing in the sense that it does not admit a nonzero finite-dimensional summand anymore. First, put E∞ := 1 −Pi Ei and observe that for wi := (Ei ⊗ Ei)w = R (bi ⊗ bi)(x) dµ(x), one has SP wi = EiSP wEi = λiEi, including the case i = ∞ and λ∞ := 0. It follows that b = b∞ ⊕ P⊕i bi and SP w = SP w∞ ⊕ P⊕i SP wi in accordance with H = E∞H ⊕ Li EiH. That SP w∞ = 0 means that b∞ is weakly mixing. Thus kP wk 6= 0 if and only if b has a nonzero finite-dimensional summand. Moreover, one has Tr(SP w) = Xi Tr(SP wi ) = Xi λi Tr(Ei) = Xi kbik2 L2(µ) and kP wk2 = Tr(S2 P w) = Xi λ2 i Tr(Ei). For the proof of Theorem A, in view of Lemma 2.2 and the fact that any nonzero µ- harmonic 1-cocycle is cohomologically non-trivial, we may assume that the 1-cocycle b is µ-harmonic. For such b, we have the following more precise form of Theorem A. For any θ ≥ 0 and any finite or infinite (possibly null) sequence σk of positive numbers, we denote by χ(θ, σk) the distribution of pθ2 +Pk σ2 kg2 k, where gk are independent standard centered Gaussian random variables. Theorem 2.4. Let G be as in Convention. Let b be a normalized µ-harmonic 1-cocycle. Let w, P w, and SP w = Pi λiEi be as defined in (∗) before the formulation of the theorem. Then, E(cid:2)(cid:12)(cid:12) 2(cid:3) = kP wk2 ≤ (min − 1(cid:12)(cid:12) to χ(θ, σk), where θ = kE∞bkL2(µ), and σ2 multiplicities (i.e., σk = λ1/2 θ2 + Pk σ2 Moreover, the random variables 1√nkb(Xn)k converge in distribution and in moments k are positive eigenvalues of SP w counted with l=1 dim ElH), which satisfy L2(µ) = 1. One has θ > 0 if and only if b admits a weakly mixing l=1 dim ElH < k ≤ Pi dim EiH)−1. for Pi−1 k = kbk2 i i 1 2 lim n→∞ kb(Xn)k2 n FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 9 summand; and σk > 0 for some k if and only if b admits a non-zero finite-dimensional summand. Proof of Theorem A and Theorem 2.4. Let b be a normalized µ-harmonic 1-cocycle. In the discussion above, we already saw kP wk 6= 0 if and only if b has a nonzero finite- dimensional summand. Moreover the above formula implies kP wk2 = Xi λ2 i Tr(Ei) ≤ (max λi) Tr(SP w) ≤ (min since Tr(SP w) = Pi λi Tr(Ei) ≤ 1. Note that Tr(Ei) = dim EiH. Next, we prove that E(cid:2) kb(Xn)k2 Z (b ⊗ b)(x) dµ∗n(x) = ZZ (b ⊗ b)(xy) dµ∗n−1(x) dµ(y) − 12(cid:3) → 2hP w, wi = 2kP wk2. Recall that Tr(Ei))−1, n i i = ZZ (b ⊗ b)(x) + (πx ⊗ πx)(b ⊗ b)(y) dµ∗n−1(x) dµ(y) = Z (b ⊗ b)(x) dµ∗n−1(x) + T n−1w = (1 + T + ··· + T n−1)w, andR kb(x)k2 dµ∗n(x) = n (see [23] and [29]). Hence E(cid:2)kb(Xn)k4(cid:3) = Z kb(x)k4 dµ∗n(x) = ZZ (kb(x) − b(y)k2)2 dµ∗n−1(x) dµ(y) = ZZ (cid:2)kb(x)k4 + 4hb(x), b(y)i2 + kb(y)k4 + 2kb(x)k2kb(y)k2(cid:3) dµ∗n−1(x) dµ(y) = E(cid:2)kb(Xn−1)k4(cid:3) + 4h = 4h ≤ 3n2 + O(n). (n − k)T k−1w, wi + n E(cid:2)kb(X1)k4(cid:3) + n(n − 1) T kw, wi + E(cid:2)kb(X1)k4(cid:3) + 2(n − 1) Xk=1 Xk=0 n−2 n−1 By Bounded Convergence Theorem, this implies that kb(Xn)k2 n E(cid:2)(cid:12)(cid:12) − 1(cid:12)(cid:12) 2(cid:3) = E(cid:2) 4 n2h = 1 n2kb(Xn)k4 − n−1 Xk=1 (n − k)T k−1w, wi + 2 nkb(Xn)k2 + 1(cid:3) 1 n (E(cid:2)kb(X1)k4(cid:3) − 1) 1 → 2hP w, wi. nb(Xn)2 − 1(cid:12)(cid:12) E(cid:2)(cid:12)(cid:12) Now since supn 3(cid:3) < ∞ by Lemma 2.3, the sequence 1 tends to a constant (which is necessarily 1) in probability if and only if one has E(cid:2)(cid:12)(cid:12) 2(cid:3) → 0. This completes the proof of Theorem A and the first part of Theorem 2.4. 1(cid:12)(cid:12) For the second half of Theorem 2.4, we first note that convergence in distribution and convergence in moments are equivalent in our setting. Indeed, by the moments condition supn E(cid:2)kb(Xn)k2d(cid:3) < ∞ (Lemma 2.3), convergence in distribution implies that in nb(Xn)2 1 nd 1 nb(Xn)2− 10 ANNA ERSCHLER AND NARUTAKA OZAWA moments (see [3, Corollary 25.12]). And conversely, since the normal distribution and the distributions χ(θ, σk) are uniquely determined by their moments (see [3, Theorem 30.1]), convergence in moments to such a distribution implies that in distribution (see [3, Theorem 30.2]). We use Martingale Central Limit Theorem (Theorem 35.12 in [3]) to prove that for any v ∈ H the random variables Sn := n−1/2hb(Xn), vi converge to a normal dis- tribution N (0, q(v)) where q(v) = hSP wv, vi. Consider the martingale array Sn,k := n−1/2hb(Xk), vi, k = 1, . . . , n, and put Yn,k := Sn,k − Sn,k−1 = n−1/2hπ(Xk−1)b(X−1 k−1Xk), vi. Since X−1 k−1Xk has the same distribution as X1, one has n Xk=1 E(cid:2)Y 2 n,k k X1, . . . , Xk−1(cid:3) = 1 n n Xk=1 h(π ⊗ π)(Xk−1)w, v ⊗ vi P → hP w, v ⊗ vi = q(v), and, for every ε, n Xk=1 E(cid:2)Y 2 n,k1{Yn,k≥ε}(cid:3) ≤ E(cid:2)kb(X1)k2kvk21{kb(X1)k≥εn1/2}(cid:3) → 0. This shows that the array Sn,k satisfies the assumption of the Martingale CLT [3, Thm 35.12], and we can conclude that Sn,n ⇒ N (0, q(v)) in distribution. Now recall that SP w = Pi λiEi and b = b∞ + P bi, and take an orthonormal basis {vi,j : j = 1, . . . , Tr(Ei)} of EiH. Then, by the previous paragraph, n−1/2hb(Xn), vi,ji converges in distribution to a centered Gaussian random variable gi,j with variance q(vi,j) = λi. Moreover, for any βi,j ∈ R, the random variables Pi,j βi,jn−1/2hb(Xn), vi,ji con- verge in moments to N (0, q(Pi,j βi,jvi,j )), where q(Xi,j βi,j vi,j) = Xi,j β2 i,jλi = Xi,j β2 i,j q(vi,j). This means that the family {hn−1/2b(Xn), vi,ji}i,j are asymptotically independent as n → ∞. Thus, for any k ∈ N, one has bi(Xn)2 = n−1/2hb(Xn), vi,ji2 ⇒ 1 n λig2 i,j, k k k Xi=1 Xj Xi=1 Xj Xi=1 where gi,j are independent standard centered Gaussian random variables. Since lim k sup n 1 nXi>k E(cid:2)( bi(Xn)2)d(cid:3) ≤ lim k CdkXi>k bik2d L2(µ) = 0 where Cd is a constant independent of k (by Lemma 2.3), one has lim n E(cid:2)( 1 nkXi bi(Xn)k2)d(cid:3) = lim nkb∞(Xn)k2 → kb∞k2 k lim n nkb(Xn)k2 → kb∞k2 L2(µ) +Pi,j λig2 for every d. Also, since 1 proof, one has 1 1 nk k Xi=1 E(cid:2)( bi(Xn)k2)d(cid:3) L2(µ) in moments by the first half of the (cid:3) i,j ∼ χ(θ, σk)2 in moments. FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 11 Recall that a group G is said to have Shalom's property HFD ([31]) if every orthogonal representation π with H 1(G, π) 6= 0 contains a non-zero finite-dimensional subrepresen- tation. In other words, G has property HFD if and only if every µ-harmonic 1-cocycle b decomposes into a (possibly infinite) direct sum of finite-dimensional summands. By The- orem 2.4, the latter happens for b if and only if limn µ∗n({x ∈ G : kb(x)k ≤ c√n}) > 0 for all c > 0. Corollary 2.5. Assume either (1) lim inf n kµ∗n − µ∗(1+δ)nk1 < 2 for some δ > 0 or (2) lim supn µ∗n(BG(c√n)) > 0 for all c > 0. Then, G has Shalom's property HFD. Proof. We prove a stronger statement that if G does not have property HFD, then for every δ > 0 there are c > 0 and a sequence (En)n of open subsets in G such that µ∗n(En) → 1 and µ∗(1+δ)n(BG(c√n)EnBG(c√n)) → 0. Suppose that there is µ-harmonic 1-cocycle b : G → H without a non-zero finite- dimensional summand. We can assume that this cocycle is normalized. Take any 0 < δ < 1. Put c := (20kbkQ)−1δ and Then, for every x ∈ En and y, z ∈ BG(c√n) one has kb(yxz)k2 ≤ kb(x)k2 + 2kb(x)kkb(y) + πyxb(z)k + kb(y) + πyxb(z)k2 < (1 + δ/2)n En := {x ∈ G : kb(x)k2 < (1 + δ/4)n}. Hence the result follows from Theorem A. (cid:3) Remark 2.6. By Kingman's subadditive ergodic theorem, the linear rate of escape lim n 1 nXn(ω)G = lim n 1 n EXnG =: lµ exists and is constant for a.e. ω ∈ (G, µ)N. Hence either of the conditions (1) and (2) in Corollary 2.5 implies that lµ = 0 and in particular that G is amenable ([17]). Remark 2.7. It is known that Z ≀ Z does not satisfy property HF D ([31, 5.4.1]). Shalom shows that any infinite amenable group with HFD admits a virtual quotient to Z ([31, 4.3.1]). By Corollary 2.5, any non-degenerate random walk on a group without virutal homomorphisms to Z (or Z ≀ Z) does not satisfy either of the conditions (1) or (2). It is apparently on open problem whether the wreath product Z2 ≀ (Z/2Z) has property HFD (see [31, 6.6]); the simple random walk on it does not satisfy either of the conditions (for "switch-walk-switch" random walks it follows from Dvoretzky–Erd os theorem ([7, 18]) that the number of distinct sites of a simple random walk on Z2 visited until the time n is asymptotically equivalent to cn/ log(n), where c > 0 is a constant. 3. MORE ON THE PROPERTY HF D We elaborate on Corollary 2.5. It says G has property HFD provided that (G, µ) satisfies the following property. We say a µ-random walk Xn is cautious if lim sup n P( max k=1,...,nXkG < c√n) > 0 for every c > 0. We look at stability of this property under extension. Let N be a closed normal subgroup of G with a length · N which may not be proper. We say N is strictly exponentially distorted in G if there exists a constant C ≥ 1 such that 1 C log(hN + 1) − C ≤ hG ≤ C log(hN + 1) + C 12 ANNA ERSCHLER AND NARUTAKA OZAWA for all h ∈ N . We will denote by · G/N the length induced by the compact generating neighborhood QN of e in G/N . Proposition 3.1. Let N ⊳ G be a closed normal subgroup which is strictly exponentially distorted, and let ¯µ be the push-out probability measure of µ to G/N . If (G/N, ¯µ) is cautious, then so is (G, µ) and in particular G has Shalom's property HFD. Proof. It suffices to show that there is a constant D ≥ 1 with the following property (cf. [35, Lemma 3.4]). Let si ∈ G be such that siG ≤ 1 and put gk := s1 ··· sk ∈ G and Mn := maxk=1,...,n gkNG/N . Then, one has maxk=1,...,n gkG ≤ D(Mn + log n + 1). To show such D exists, for each k, pick hk ∈ N such that g−1 k NG/N ≤ Mk. Then, h−1 k−1hkN ≤ exp(4CMk). Hence hkN ≤ n exp(4CMn) for all k ≤ n and so k−1hkG ≤ 2Mk + 1 ≤ 3Mk and so h−1 k hkG = g−1 max k=1,...,nhkN ≤ C log(2n exp(4CMn)) ≤ (D − 1)(Mn + log n + 1) for some constant D ≥ 1. Since g−1 k hkG ≤ Mn, we are done. (cid:3) Shalom ([31, Theorem 1.13]) has shown that polycyclic groups have property HFD by invoking Delorme's theorem ([6]) that connected solvable Lie groups have the correspond- ing property, and asked if there is another proof of HFD. It is plausible that all connected solvable groups are cautious. We note that in light of Osin's result ([28]) this problem reduces to the case for connected Lie groups with polynomial volume growth. Corollary 3.2. Let K be a non-archimedean local field and Zd y K n be a semi-simple linear action such that the semi-direct product Zd ⋉ K n is compactly generated. Then, Zd ⋉ K n has Shalom's property HFD. Proof. Let ν0 be the standard nearest neighborhood random walk on Zd and ν1 be a uni- form probability measure on the compact subgroup {x ∈ K : x ≤ 1}. Since (Zd, ν0) is cautious, for µ = 1 ), the random walk (Zd ⋉ K n, µ) is cautious. (cid:3) 2 (ν0 + ν⊗n 1 4. HARMONIC COCYCLE bµ,U CONSTRUCTED FROM DIFFERENCES OF SHIFTS OF µ∗n In this section, we give a rather "explicit" (although we crucially use a non-principal ultrafilter) construction of a non-zero harmonic cocycle on a group that does not satisfy Kazhdan's property (T). In particular, when G is a discrete finitely generated amenable group, a normalized µ-harmonic cocycle bµ will be obtained as an ultralimit of the se- quence µ∗n − gµ∗n ∈ ℓ2(G) after normalization. Throughout this section, we assume (in addition to Convention) that µ is compactly supported and µ = µ′∗2 for some symmetric probability measure µ′ on G. We fix a non-principal ultrafilter U on N and denote by limU the corresponding ultra- limit. Then, the ultrapower Hilbert space HU of a given Hilbert space H is defined to be HU := ℓ∞(N;H)/{(vn)∞n=1 : limU kvnk = 0} with the inner product h[v′n]n, [vn]ni := limU hv′n, vni, where [vn]n is the equivalence class of (vn)n ∈ ℓ∞(N;H). An orthogonal representation π of G on H gives rise to the ultrapower representation πU on HU by πU g [vn]n = [πgvn]n. (NB: In general, the ultra- power representation is no longer continuous.) We apply this construction to an orthogonal representation (π,H) which admits an approximate invariant vectors but no non-zero in- variant vectors. By definition, such an orthogonal representation exists if and only if G does not satisfy Kazhdan's property (T) (see [2]). FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 13 Lemma 4.1. Let (π,H) be an orthogonal representation which admits an approximate in- variant vectors but no non-zero invariant vectors, and consider the positive and contractive operator T := π(µ) on H. Then, there is a unit vector v ∈ H such that the corresponding probability measure ν on [0, 1], defined by the formula 0 tn dν(t) = hT nv, vi, Z 1 satisfies 1 ∈ supp ν and ν({1}) = 0. Proof. Let ET denote the spectral measure corresponding to the self-adjoint operator T . Since (π,H) admits approximate invariant vectors, the spectrum of T contains 1, which means that ET ([1 − 1/n, 1]) 6= 0 for any n. Hence, there is a unit vector v ∈ H such that ET ([1 − 1/n, 1])v 6= 0 for any n. On the other hand, ET ({1}) = 0 since (π,H) has no non-zero invariant vectors. The probability measure ν(· ) := hET (· )v, vi corresponding to v satisfies the desired conditions. (cid:3) Take (π,H, v) as above and put T = π(µ). In case G is a discrete finitely generated infinite amenable group, one can take (π,H, v) to be (λ, ℓ2(G), δe) by Kesten's theorem ([21]). Consider the coboundary cn : G → H given by cn(g) = T n/2v − π(g)T n/2v and its normalization bn := kcnk−1 L2(µ)cn. We note that L2(µ) = 2h(T n − T n+1)v, vi = 2Z 1 kcnk2 0 tn(1 − t) dν(t). We will define the cocycle bµ to be the ultralimit of bn. For continuity of bµ, we need equi-continuity of bn's. Observe that for every g ∈ G, one has cn(g) = −ZG ( dµ dm − g dµ dm )(x)cn−2(x) dm(x). Let K = Q supp µ (recall that Q is a relatively compact generating subset of G and that supp µ is assumed compact) and take a constant C which satisfies kckK ≤ CkckL2(µ) for every cocycle c (see Lemma 2.1). Then by the above equality, for every g ∈ Q, one has kbn(g)k ≤ kcn−2kK dm ∈ L1(G), the function g 7→ k dµ kcnkL2(µ) · k dµ dm − g dµ dmkL1 ≤ C kcn−2kL2(µ) kcnkL2(µ) dm −g dµ Since dµ of bn's follows from the following auxiliary lemma. Lemma 4.2. Let ν be a probability measure on [0, 1] such that 1 ∈ supp ν and ν({1}) = 0. Then, γ(n) := R 1 0 tn(1 − t) dν(t) satisfies γ(n) ց 0 and γ(n + 1)/γ(n) ր 1. dmkL1 is continuous. Thus, equi-continuity dµ dm − g dµ dmkL1. · k Proof. The first assertion is obvious. Since γ(n + 1) = Z tn/2(1 − t)1/2 · n(n+2)/2(1 − t)1/2 dν(t) ≤ γ(n)1/2γ(n + 2)1/2, the sequence γ(n+1)/γ(n) is increasing and has a limit δ ≤ 1. Suppose for a contradiction that δ < 1. Then, one has γ(n) ≤ Cδn and so R 1 0 tn dν(t) = P∞k=n γ(k) ≤ C′δn for every n, where C and C′ are some constant independent of n. This implies supp ν ⊂ [0, δ], a contradiction. Hence δ = 1. (cid:3) Since bn's are equi-continuous and kbn(g)k ≤ gGkbnkQ is bounded for each g, the formula bµ(g) := [bn(g)]n ∈ HU 14 ANNA ERSCHLER AND NARUTAKA OZAWA defines a continuous map such that bµ(gh) = bµ(g) + πU g bµ(h). Since bµ is continuous, the ultrapower orthogonal representation πU is continuous when restricted to span b(G). Hence bµ is a 1-cocycle. It is normalized: kbµk2 L2(µ) = Z limUkbn(x)k2 dµ(x) = limU Z kbn(x)k2 dµ(x) = 1, where, to interchange the ultralimit and integration, we have used the fact that µ is com- pactly supported and bn's are equi-continuous. The constructed 1-cocycle bµ may depend on the choice of a non-principle ultrafilter U (see Theorem C), and we will write bµ,U instead of bµ when we want to emphasize the role of the ultrafilter U . The following re- proves the results of Mok ([25]), Korevaar–Schoen ([22]), and Shalom ([32]) mentioned in Introduction. Theorem 4.3. Let G be a compactly generated locally compact group which does not have Kazhdan's property (T) and µ, (π,H, v), and bµ be as above. Then, bµ is a normalized µ-harmonic cocycle. Proof. It only remains to prove that bµ is harmonic. Put γ(n) = R tn(1 − t) dν(t). Then, one has kZ bn(x) dµ(x)k2 = γ(n) − γ(n + 1) 2γ(n) → 0 by Lemma 4.2. Hence, for every v′ = [v′n]n ∈ HU , one has hZ bµ(x) dµ(x), v′i = Z limUhbn(x), v′ni dµ(x) = limU Z hbn(x), v′ni dµ(x) = limUhZ bn(x) dµ(x), v′ni = 0. This meansR bµ(x) dµ(x) = 0 and bµ is harmonic. In case G is a discrete amenable group and (π,H, v) = (λ, ℓ2(G), δe), a computation (cid:3) yields that and kcnk2 L2(µ) = 2(µ∗n(e) − µ∗n+1(e)) kbµ(g)k2 = limUkbn(g)k2 = limU µ∗n(e) − µ∗n(g) µ∗n(e) − µ∗n+1(e) . Proof of Theorem B. By Theorem A we know that E(cid:2) kc(Xm)k2 malized harmonic cocycle c without non-zero finite-dimensional summands. We will show that in case G does not admit any non-zero harmonic finite-dimensional cocycle (which is the case when G is a finitely generated amenable group without virtually abelian infinite quotients), this convergence is uniform for normalized harmonic cocycles c on G. Indeed, we have seen in the proof of Theorem A that − 12(cid:3) → 0 for any nor- m E(cid:2)kc(Xm)k2 m − 12(cid:3) ≤ 4 m2h m−1 Xk=1 (m − k)T k−1w, wi + 1 mkck4 Q E(cid:2)X14 G(cid:3) → 0 for every normalized µ-harmonic 1-cocycle c, where T = R (π ⊗ ¯π)g dµ(g) and w = R (c ⊗ ¯c)(g) dµ(g). Note that kckQ is uniformly bounded by Lemma 2.1. Therefore, it suffices to prove that limk kT kwk = 0 uniformly for c. Suppose that the latter is not the FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 15 case: there are ε > 0, a subsequence km → ∞, and normalized harmonic cocycles cm with the corresponding Tm and wm such that kT km m wmk ≥ ε for all m. Fix a non-principal ultrafilter U and let cU denote the U -ultralimit cocycle of the sequence (cm)m, with the corresponding objects denoted by TU and wU . Then, cU is a normalized harmonic cocycle. Moreover since t2k is decreasing in k for any t ∈ [−1, 1], one has for each k hT 2k U wU , wUi = limUhT 2k m wm, wmi ≥ limUhT 2km m wm, wmi ≥ ε2. Let Q denote the spectral projection of TU corresponding to eigenvalues {−1, +1}. Then, kQwUk2 = lim k hk−1(1 + T 2 + T 4 + ··· + T 2(k−1))wU , wUi ≥ ε2. U QwU = QwU , the vector QwU is invariant under (π ⊗ ¯π)g for all g ∈ supp µ∗2. Since T 2 However since G0 := hsupp µ∗2i has finite-index in G, it does not admit a non-zero µ∗2-harmonic 1-cocycle, which implies that QwU = 0 (as discussed in the proof of Theo- rem A). We have arrived at a contradiction. It follows that if G satisfies the assumption of Theorem B, then E(cid:2) kc(Xm)k2 uniformly for normalized µ-harmonic 1-cocycles c. In particular, m −12(cid:3) → 0 lim m lim sup n E(cid:2)kbn(Xm)k2 m − 12(cid:3) = lim m sup U E(cid:2)kbµ,U (Xm)k2 m − 12(cid:3) = 0. (Note that lim supn λn = supU limU λn for any bounded sequence λn.) Since by Lemma 4.2, this completes the proof of Theorem B (after exchanging µ with µ∗2). (cid:3) 1 m = lim n µ∗n(e) − µ∗n+1(e) µ∗n(e) − µ∗n+m(e) 5. ℓp-THIN SUBGROUPS 5.1. Definitions. Take a finitely generated group G equipped with a probability measure µ, and ask again what information about its subgroups and quotient groups one can obtain by looking on the behavior the random walk (G, µ). To ensure the existence of non- trivial quotients, we may search normal subgroups of G defined by convolutions of G. A more general question one can ask is what are possible (not necessarily normal subgroups) defined in such terms. Definition 5.1. [ℓp-thin subgroups Hµ,p,q]. Let G be an infinite group generated by a finite set S, and µ be a probability measure on G. Fix some q ≥ 0, p ≥ 1 and a sequence ni tending to ∞. Assume that µ is such that (µ∗n)q is in lp(G) for all n(this holds for example if µ has finite support). Let α(n) denotes the maximum of ℓp norm of (µ∗n)q − g(µ∗n)q, where the maximum is taken over g ∈ S. Consider g ∈ G for which (µ∗ni )q − g(µ∗ni)qp/α(ni) → 0 as i → ∞. If G contains at least two elements, then by the triangular inequality in ℓp, such elements form a subgroup of G, which we we call the main ℓp-thin subgroup and which we denote by Hµ,p,q,ni (and Hµ,p for short, if ni is specified and q = 1). Now we define ℓp-thin subgroups associated an arbitrary function α(n). Consider g such that (µ∗n)q − g(µ∗n)qp/α(n) tends to 0 as n tends to infinity. By triangular in- equality in ℓp such elements form a subgroup of G, which we denote Hµ,p,q,α. We call this subgroup ℓp-thin subgroup associated to α(n). Remark 5.2. For q = 0 in the definition above we use the convention 00 = 0; the ℓ1 norm in this case is therefore the cardinality of the symmetric differences of the supports of µ∗n and gµ∗n, that is the cardinality of the set of points x such that either x is in the support 16 ANNA ERSCHLER AND NARUTAKA OZAWA of µ∗n and gx is not in this support or vice versa. In the definition we have assumed that p ≥ 1. We can extend the definition for the case p = 0, defining α(n) as the maximum of the cardinality of the support of (µ∗n)q − g(µ∗n)q, where the maximum is taken over g ∈ S. In this case we obtain H0,1,µ = H1,0,µ for all µ. Observe that that if the support of µ is a finite symmetric generating set containing the identity, then the support of µ∗n is the ball of radius n in the word metric associated to S. It is clear that the scaling sequence α(n) depends of a finite generating set S up to multiplication by a constant only, and thus the definition of main ℓp-thin subgroups does not depend on the choice of S. In many situation the limit behavior of (µ∗n)q − g(µ∗n)q does not depend on the sub- sequence of possible n's. However, in some situation this quantity, and the corresponding ℓp-thin subgroups may depend on the choice of a subsequence, see Theorem C and Corol- lary 5.11. Remark 5.3. If p ≥ 1, it is known that a normalized sequence vn ∈ ℓ1(G) is almost invariant in ℓ1 with respect to the shift by some element g ∈ G if and only if v1/p (which is clearly a sequence in ℓp(G)) is almost invariant in ℓp with respect to the shift by g (see e.g. the proof of Theorem 8.3.2 in [30]. This implies that the main ℓp-thin subgroups satisfy Hp,1 = H1,p = Hp/q,q for any p, q ≥ 1 whenever (µ∗ni )p does not admit a subsequence of almost invariant vectors in ℓ1. This happens for example for p = 2, if G non-amenable and for p = 1 if the Poisson boundary of (G, µ) is non-trivial, for all ni ([20]). n It is possible that the statement of Remark 5.3 remains valid without the assumption of non-almost-invariance. Instead (µ∗n)q in the Defintion 5.1, one can consider more generally a sequence of functions fi and consider the difference of corresponding shifted functions, as a function of g. We have already remarked that for p = 2, q = 1, µ being equidistributed on a finite symmetric set of G, the values of bµ,U are defined by the unmarked Cayley graph of G. In particular, for p = 2, q = 1 and µ being a measure equidistributed on a finite generating set S , the ℓp-thin subgroups can be described in terms of unmarked Cayley graph of (G, S): Remark 5.4. p = 2, q = 1, µ is symmetric measure on G. Fix a sequence αi, tending to infinity. An element g belongs to the subgroup Hµ,2,1,α if and only if (µ2n(e) − µ∗2n(g))/α2 n) → 0 as n → 0. In particular, if µ is equidistributed on a finite symmetric generating set S, subgroups Hµ,2,1,αi are defined by unmarked Cayley graph of (G, S). 2 = µ∗n2 2 + gµ∗n2 Proof. Observe that gµ∗n − µ∗n2 2 − 2hµ∗n, gµ∗ni = 2µ∗n2 2 − 2hµ∗n, gµ∗ni = 2(Px∈G(µ∗n(x))2 −Px∈G µ∗n(x)µ∗n(gx)), Since µ is symmetric, this is equal to 2(Px∈G µ∗n(x)µ∗n(x−1)−Px∈G µ∗n(gx)µ∗n(x−1) = 2(µ2n(e)−µ∗2n(g)). If µ is equidistributed on a finite symmetric generating set S, observe that µ∗2n(e) and µ∗2n(g) are defined by the unmarked Cayley graph of (G, S) and the vertex in this Cayley graph corresponding to g. Remark 5.5. In a particular case when q = 1, p = 2 and G is non-amenable, the main ℓ2- thin subgroup in 5.1 coincides with the group, studied by Elder and Rogers in [8]. However, if q = 1, p = 2 and G is amenable, the group defined in the above cited paper coincides with G, while the main ℓ2-thin subgroup Hµ,p is never equal to G (for any infinite group G). FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 17 Now assume that µ has finite support, and consider the mappings bp,q µ,U , defined in the in- troduction. Namely, for any non-principal ultrafilter U on N, put αp,q(n) = maxs∈S k(µ∗n)q− s(µ∗n)qkp and define the cocycle bp,q µ,U : G → ℓp(G)U by bp,q µ,U (g) = [αp,q(n)−1((µ∗n)q − g(µ∗n)q)]n ∈ ℓp(G)U . µ,U (G) is contained in a G-invariant separable Lp-subspace of ℓp(G)U . The cocycle bp,q µ,U is independent, modulo scalar multiple, of the choice of the finite gener- ating subset S. We note that ℓp(G)U is an abstract Lp-space on which G acts isometrically. Hence bp,q Lemma 5.6. 1) [Direct products, q = 0, p ≥ 1] Let G be a direct product of A of subexponential growth and B of exponential growth, and let µ = µA× µB where µA(e) > 0. Then there exists a subsequence ni such that subgroup Hµ,0,p(G) = Hµ,p,0(G) contains A. Moreover, for any ni as above, any ultrafilter U such that U (ni) = 1 for q = 0 and p ≥ 1 satisfy bp,q 2) [Direct product, p = q = 1] Let G be a direct product of a group A and B; let µA, µB be non-degenerate measures on A and B such that the Poisson boundary of a random walk (A, µA) is trivial and Poisson boundary of (B, µB) is non-trivial. Put µ = µA × µB. Then for any choice of ni the main ℓ1-thin subgroup Hµ,1,1(G) contains A. Moreover, for any ultrafilter U it holds b1,1 U,µ = bp,q U,µB . U,µ = b1,1 U,µB . 3) [Direct products, q = 1, p = 2] Let G be a direct product of an amenable A and non-amenable group B, µ = µA × µB. Then for any ni, the main ℓ2-thin subgroup Hµ,1,2(G) = Hµ,2,1(G) contains A. Moreover, for any ultrafilter U it holds b2,1 . U,µ = b2,1 U,µB Proof. First we prove the claims of 1), 2), 3) about ℓp-thin subgroups. Observe that since B is of exponential growth, for any finite set S there exists v > 1 such that vG,S(n) ≥ vn for all n. This implies that for each finite generating set SB of B and each C1 < 1 there exists C2 > 0 such that for all n at least C1n among balls of radius i = 1, ..., n have boundary greater than C2vB,SB (i). (Indeed, otherwise vB,SB (n) ≤ Rn(1−C1) (1+C2)C1n, where RB denotes the cardinality of B, SB, and taking C1 close to 1 and C2 close to 0 we would get a contradiction). B among the the balls of radius i = 1, ..., n have boundary at most ǫ2vA(i). Since A is of subexponential growth, for each C and any ǫ1, ǫ2 > 0 at least (1 − ǫ2)n Consider a generating set S = SA × SB, where SA, SB are generating sets of A, B respectively. We have BS(i) = BSA(i)× BSB (i). Here BG,S(i) denotes the ball of radius i in G, S. Observe also that for S ∈ SA it holds sBG,S(i) \ BG,S(i) = s(BA,SA(i) × BB,SB (i)) \ (BA,SA(i) × BB,SB (i)) = (sBA,SA(i) \ BA,SA(i)) × BB,SB (i), and the cardinality of this set is at most 2(vA,SA(i) − vA,SA(i − 1))vB,SB (i), and with the same argument the cardinality of sBG,S(i) \ BG,S(i), for s ∈ SB is at least 2/SB(vB,SB (i) − vB,SB (i−1)vA,SA(i) for some s ∈ SB. This shows that there exists a sequence ni, tending to infinity, such that vA,SA(ni) − vA,SA(ni − 1) vA,SA(ni) vB,SB (ni) vB,SB (ni) − vB,SB (ni − 1) tends to 0 as i tends to infinity. By Remark 5.2 we know that for any group it holds Hµ,0,p(G) = Hµ,p,0(G). Note that for any ni as above the this thin subgroup Hµ,0,p(G) = Hµ,p,0(G) with respect to a subsequence ni contains all s ∈ SA. Therefore, in this case this subgroup contains A. A − aµ∗n A )1. Since the non- degenerate walk (A, µA) has trivial Poisson-Furstenberg boundary, for any a ∈ A it holds 2) We recall that µ∗n = µ∗n A )µ∗n(B). It holds therefore (µ∗n − aµ∗n)1 = (µ∗n B . Take a ∈ A. Observe that aµ∗n − µ∗n = (µ∗n A − aµ∗n A µ∗n 18 ANNA ERSCHLER AND NARUTAKA OZAWA A − aµ∗n B − bµ∗n 3) For g = (g1, g2), g1 ∈ A, g2 ∈ B, µ∗n(g1, g2) = µ∗n A )1 → 0 as n tends to ∞, and therefore (µ∗n − aµ∗n)1 → 0 as n tends (µ∗n to ∞ (see Kaimanovich Vershik [20]). The above mentioned characterization also shows that since the Poisson boundary of (B, µB) is non-trivial, there exists b ∈ B such that (µ∗n B )1 ≥ c > 0, and hence (µ∗n − bµ∗n)1 ≥ c > 0 for some positive constant c and all n. B (g2). For h ∈ A, µ∗n(h(g1, g2))/µ∗n(g1, g2) = µ∗n A (g1) → 1 as n → ∞, by [1] since A is amenable [1]. Analogously, for h ∈ B it holds µ∗n(h(g1, g2))/µ∗n(g1, g2) = B (hg2)/µ∗n µ∗n B (g2) → Ch, where Ch 6= 1 for some h among generators of B, since B is non-amenable [1]. This implies that the scaling sequence α(n) is equivalent up to mul- tiplicative constant to µ∗n(e) = µ∗n B (e). Using Remark 5.4 we conclude that for all s ∈ SA sµ∗n − µ∗n2/α(n) → 0, and hence any s ∈ A, s belongs to the ℓ2 thin subgroup for q = 1, p = 2. By Remark 5.3 we know that under assumption of 3) it holds Hµ,1,2(G) = Hµ,2,1(G). A (hg1)/µ∗n A (g1)µ∗n A (e)µ∗n Now to prove the claims about the cocycles, take g = (a, b) ∈ A × B, put g′ = (e, b) and g′′ = (a, e). It holds g = g′g′′. Under the assumption on p and q in 1), 2), 3) observe that (µ∗ni )q − g′(µ∗ni )qp − (µ∗ni )q − g′′(µ∗ni )qp ≤ (µ∗ni )q − g(µ∗ni)qp ≤ ≤ (µ∗ni )q − g′(µ∗ni )qp + (µ∗ni )q − g′′(µ∗ni )qp and that (µ∗ni )q − g′(µ∗ni )qp = (µ∗ni B )q − g′(µ∗ni 1) of Remark 5.7 and completes the proof of the Lemma. A )qpµni A p. This allows us to use 1) Let αp,q A (n) be the maximal ℓp norm of (µ∗n B (n) is defined analogously. Let θµ(n) be equal to αp,q Remark 5.7. G = A× B, µ = µA × µB, S = SA × SB, SA and SB are finite generating sets of A and B. G (n) be the maximal ℓp norm of (µ∗n)q−s(µ∗n)q, where the maximum is taken A )q, the maximum is G (n) divided by B (ni) tends to zero for some sequence ni and U is a over s ∈ S; and let αp,q over s ∈ SA and αp,q the ℓp norm of (µ∗n)q. If θp,q non-principal ultrafilter such that U ({ni}) = 1, then bp,q 2) Take q = 1, p = 2. Put θ(n) := (µ∗2n − µ∗2n+1)/µ∗2n. Then θ(n) = θµ(n)2. In particular, if θA(ni)/θB(ni) tends to zero and U is a non-principal ultrafilter such that U ({ni}) = 1, then the corresponding harmonic cocycle is defined by that of B, that is bµ,U = bµB ,U . A )q − s(µ∗n A (ni)/θp,q µ,U = bp,q µB ,U . Remark 5.8. The fact that A×B, A is of subexponetial growth, B is of exponential growth, satisfy the claim of 1), Lemma 5.6 not only for some sequence ni but for all sequences can be shown to be equivalent to a positive answer to both following questions A): is it true that no subset of balls is a Foelner sequence in A? B): Is it true that all balls form a Foelner sequence in A? To our knowledge, it is not known whether to answer to A) is positive for all groups of exponential growth (this question is mentioned e.g. in [34]), and whether the answer to B) is positive for all groups of subexponential growth. Example 5.9 (Dependance of ℓp-thin subgroups on p). Let G = Fm × Zd ≀ A, where m ≥ 2, d ≥ 3 and A is a finite group containing at least two elements. Let µ be a non- degenerate symmetric finitely supported measure. Then ℓ2-thin subgroup is not equal to ℓ1-thin subgroup. FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 19 Proof. Observe that the ℓ2-thin subgroup Hµ,1,2 = Hµ,2,1 contains Zd ≀ A by 3) of Lemma 5.6 (in fact, it is equal to Zd ≀ A), while there exists g ∈ Zd ≀ A which does not belong to ℓ1-thin subgroup since the Poisson boundary of Zd ≀ A is non-trivial. Remark 5.10. Let G = C ≀ A, where C is an infinite group of at least cubic growth and A is a finite group containing at least two elements. Let µ be a symmetric finitely supported "switch-walk-switch" measure on G. One can show that Hµ,1,1 is a finite subgroup of G. One can also show that for any integer k ≥ 0 there exists µ as above such that Hµ,1,1 is isomorphic to Am. In particular, this main ℓ1 thin subgroup Hµ,1,1 depends on the choice of a finitely supported symmetric measure µ and this subgroup is not normal. Proof of Theorem C. Assume d = 2 (the general case d ≥ 2 is analogous). We construct G1 and G2 as piecewise-automatic groups with returns of automata τ1, τ2, where τ1, τ2 : A × X → A, the group generated by (A, τ1) is of intermediate growth, τ2 : A × X → A, the group H2 generated by (A, τ2) is non-amenable, and the action of A, considered as generators of H, is contracting for the action of τ1 for each brach of the rooted tree (see [9]). More precisely, we chose automata τ1 and τ2 with the following properties: τ2 is a finite state automaton, containing e, a, b, c, d as its states, such that e acts trivially and a, b, c, d generate the free product Z/2Z ∗ (Z/2Z + Z/2Z) in the group generated by τ2. If the states of τ2 are e, a, b, c, d and the alphabet is 0, 1, we take as τ1 the standard finite state automaton for the first Gigorchuk group (A = {e, a, b, c, d}, X = {0, 1}. In this case we can take as G1 and G2 either piecewise-automatic group or a piece-wise automatic group with returns defined by τ1, τ2 and ti, Ti, i ≥ 1, Ti−1 < ti < Ti. We do not know if τ2 as above exists, and therefore we consider as in [9] an automaton τ2 with the space of states possibly larger than e, a, b, c, d (such automata exist by the result of Olijnyk [27], that shows that any free product of finite groups imbeds in a group generated by a finite state automaton), and we take as τ1 the standard finite state automaton for the first Grigorchuk group, (extended to some larger alphabet than 0 and 1 if the alpaheth of τ2 contains more than two letters) and consider the corresponding piecewise automatic group with returns Gτ1,τ2(ti, Ti). To construct G1 and G2, we fix τ1, τ2 and construct sequences t1 i−1 < i ) by a simultaneous inductive procedure and we put G1 = i and t2 i , T 1 i , T 2 i (T 1 t1 i < T 1 Gτ1,τ2(t1 i , T 2 i , T 1 i < T 2 i−1 < t2 i ) and G2 = Gτ1,τ2(t2 i , T 2 i ) . We need the following properties of piece-wise autmatic group with returns Gτ1,τ2(ti, Ti) (see proof of Proposition 1 in [9]). There exist Ψ : N → N and for each i there exist "comparison groups" A(t1, T1, t2, T2, . . . ti) and B(t1, T1, t2, T2, . . . ti, Ti), such that the following holds for all non-decreasing sequences ti, Ti: (1) all groups A(t1, T1, t2, T2, . . . Ti−1) have a finite index subgroup which imbeds as a subgroup in a finite direct power of the the first Grigorchuk group G1 (generated by (A, τ1), a surjective homomorphism to the group, generated by the automaton (A, τ2), (2) all groups B(t1, T1, t2, T2, . . . Ti−1, ti) have a finite index subgroup which admits (3) the balls of radius Ψ(ti) in G(t1, t2, ..., T1, T2, ...) and A(t1, T1, t2, T2, . . . ti−1, Ti−1) (4) the balls of radius Ψ(Ti) in G(t1, t2, ..., T1, T2, ...) and B(t1, T1, t2, T2, . . . Ti−1, ti, ) coincide, coincide. Let G, SG, H, SH be finitely generated groups such that the balls of radius R + C in the marked Cayley graphs of G, SG, H, SH coincide. Let µH and µG are measures which 20 ANNA ERSCHLER AND NARUTAKA OZAWA are equal after the identifications of these balls and such that lG(s) ≤ C for any s in the support of µG. Observe that for any n ≤ R the scaling functions in the definition of ℓp-thin subgroups are equal : αG,µG,p(n) = αH,µH ,p(n), α′G,µG,p(n) = α′H,µH ,p(n) and for each g in the ball of radius C in the Cayley graph of (G, SG) ℓp norms of g(µ∗n G )q are equal to the ℓp norm of h(µ∗n H )q for h being the corresponding element in the ball of radius C of (H, SH ). H )q − (µ∗n G )q − (µ∗n 1, T 1 1 , t1 1, T 1 1 , t1 1, T 2 1 , t1 2, T 2 2 , . . . t2 1, T 2 1 , t1 2, T 2 2, T 1 2 , . . . T 1 2, T 1 2 , . . . T 1 i−1 and t2 i > M∗i and T 2 i−1), B = B(t2 Suppose that we have chosen already t1 Now suppose that we have chosen already t1 2 , . . . t2 i . For any ǫ > 0 there exist Mi such that for all M′i > Mi there exists M∗i with the following i > M∗i and any n : Mi < n < M′i the ratio of ℓp property. For any t1 norms s1(µ∗n)q − (µ∗n)q and s2(µ∗n)q − (µ∗n)q in G = G1 × G2 is smaller than ǫ for all s ∈ S1 and some s ∈ S2. To prove this , we combine the observation about Cayley graphs above with the claims 1), 2), 3) of Lemma 5.6, for A = A(t1 i ). Tthe group A is of intermediate growth and hence this group is amenable and finitely supported random walks have trivial boundary, B has a finite index subgroup sujecting to a non-amenable group, and hence non-amenable. 1, T 1 2 , . . . t2 For any ǫ > 0 there exist Ni such that for all N′i > Ni there exists N∗i with the following i+1 > M∗i and any n : Ni < n < N′i the ratio of ℓp property. For any T 1 norms of s2(µ∗n)q − (µ∗n)q and s1(µ∗n)q − (µ∗n)q in G = G1 × G2 is smaller than ǫ for all s2 ∈ S2 and some s1 ∈ S1. there exist sequences ni, mi tending to infinity, such that the following holds. The ratio of ℓp norms of s1(µ∗ni )q − (µ∗ni)q and the scaling sequence α(ni) tend to 0 for all s1 ∈ S1. This implies that all s1 ∈ S1, as well as all g ∈ G1 belong to the main ℓp thin subgroup Hµ,p,q , corresponding to ni. The ratio of ℓp norms of s2(µ∗mi )q − (µ∗mi )q and the scaling sequence α(mi) tend to 0 for all s2 ∈ S2. This implies that all s2 ∈ Se, as well as all g ∈ G2 belong to the main ℓp thin subgroup Hµ,p,q , corresponding to mi. Consider an ultrafilter Um such that U (mi) = 1 and an ultrafilter Un such that U (ni) = 1. Using 1), 2), 3) of Lemma 5.6 we also observe that bp,q This implies that for some choice of t1 i > N∗i and t2 is equal to bp,q is equal to bp,q i , T 1 i and t2 and that bp,q 1, T 2 1 , t1 2, T 2 1 , t1 2, T 1 2 , . . . t1 i , T 2 i i and t2 i , T 2 i . µ,Un µ2,Un µ,Um µ1,Um . Corollary 5.11. Let Gi, µi be as in the formulation of Theorem C. Take q = 0, 1 or 2 and p = 1 or 2. For each j : 1 ≤ j ≤ D there exists ni,j such that for all the main ℓp-thin subgroup Hµ,p,qof G with respect to ni = ni,j contains Qk:k6=j Gk. In particular, there exist at least D not equal ℓp-thin subgroups. REFERENCES [1] Andr´e Avez, Limite de quotients pour des marches al´eatoires sur des groupes, C. R. Acad. Sci. Paris S´er. A-B 276 (1973), A317–A320 (French). MR0315750 [2] Bachir Bekka, Pierre de la Harpe, and Alain Valette, Kazhdan's property (T), New Mathematical Mono- graphs, vol. 11, Cambridge University Press, Cambridge, 2008. MR2415834 [3] Patrick Billingsley, Probability and measure, 3rd ed., Wiley Series in Probability and Mathematical Statistics, John Wiley & Sons, Inc., New York, 1995. A Wiley-Interscience Publication. MR1324786 [4] Marc Burger and Shahar Mozes, Lattices in product of trees, Inst. Hautes ´Etudes Sci. Publ. Math. 92 (2000), 151–194 (2001). MR1839489 de Cornulier, Fixed [5] Yves point and almost fixed point, after Gromov and V.Lafforque (http://www.normalesup.org/ cornulier/grolaf.pdf). [6] Patrick Delorme, 1-cohomologie des repr´esentations unitaires des groupes de Lie semi-simples et r´esolubles. Produits tensoriels continus de repr´esentations, Bull. Soc. Math. France 105 (1977), no. 3, 281–336 (French). MR0578893 FINITE-DIMENSIONAL REPRESENTATIONS CONSTRUCTED FROM RANDOM WALKS 21 [7] A. Dvoretzky and P. Erdos, Some problems on random walk in space, Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability, 1950., University of California Press, Berkeley and Los Angeles, 1951, pp. 353–367. MR0047272 [8] Murray Elder and Cameron Rogers, Random walks and amenability of groups (https://arxiv.org/abs/1605.04065). [9] Anna Erschler, Piecewise automatic groups, Duke Math. J. 134 (2006), no. 3, 591–613, DOI 10.1215/S0012- 7094-06-13435-X. MR2254627 [10] A. Erschler, Almost invariance of distributions for random walks on groups, preprint, arXiv:1603.01458. [11] David Fisher and Gregory Margulis, Almost isometric actions, property (T), and local rigidity, Invent. Math. 162 (2005), no. 1, 19–80, DOI 10.1007/s00222-004-0437-5. MR2198325 [12] R. I. Grigorchuk, Degrees of growth of finitely generated groups and the theory of invariant means, Izv. Akad. Nauk SSSR Ser. Mat. 48 (1984), no. 5, 939–985 (Russian). MR764305 [13] Mikhael Gromov, Groups of polynomial growth and expanding maps, Inst. Hautes ´Etudes Sci. Publ. Math. 53 (1981), 53–73. MR623534 (83b:53041) [14] M. Gromov, Random walk in random groups, Geom. Funct. Anal. 13 (2003), no. 1, 73–146, DOI 10.1007/s000390300002. MR1978492 [15] , Asymptotic invariants of infinite groups, Geometric group theory, Vol. 2 (Sussex, 1991), London Math. Soc. Lecture Note Ser., vol. 182, Cambridge Univ. Press, Cambridge, 1993, pp. 1–295. MR1253544 [16] Alain Guichardet, Sur la cohomologie des groupes topologiques. II, Bull. Sci. Math. (2) 96 (1972), 305–332 (French). MR0340464 [17] Y. Guivarc'h, Sur la loi des grands nombres et le rayon spectral d'une marche al´eatoire, Conference on Random Walks (Kleebach, 1979), Ast´erisque, vol. 74, Soc. Math. France, Paris, 1980, pp. 47–98, 3 (French, with English summary). MR588157 [18] Naresh C. Jain and William E. Pruitt, The range of random walk, Proceedings of the Sixth Berkeley Sym- posium on Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Univ. California Press, Berkeley, Calif., 1972, pp. 31–50. MR0410936 [19] Kate Juschenko and Nicolas Monod, Cantor systems, piecewise translations and simple amenable groups, Ann. of Math. (2) 178 (2013), no. 2, 775–787, DOI 10.4007/annals.2013.178.2.7. MR3071509 [20] V. A. Kaımanovich and A. M. Vershik, Random walks on discrete groups: boundary and entropy, Ann. Probab. 11 (1983), no. 3, 457–490. MR704539 [21] Harry Kesten, Symmetric random walks on groups, Trans. Amer. Math. Soc. 92 (1959), 336–354. MR0109367 [22] Nicholas J. Korevaar and Richard M. Schoen, Global existence theorems for harmonic maps to non- locally compact spaces, Comm. Anal. Geom. 5 (1997), no. 2, 333–387, DOI 10.4310/CAG.1997.v5.n2.a4. MR1483983 [23] James R. Lee and Yuval Peres, Harmonic maps on amenable groups and a diffusive lower bound for random walks, Ann. Probab. 41 (2013), no. 5, 3392–3419, DOI 10.1214/12-AOP779. MR3127886 [24] Hiroki Matui, Some remarks on topological full groups of Cantor minimal systems, Internat. J. Math. 17 (2006), no. 2, 231–251, DOI 10.1142/S0129167X06003448. MR2205435 [25] Ngaiming Mok, Harmonic forms with values in locally constant Hilbert bundles, Proceedings of the Con- ference in Honor of Jean-Pierre Kahane (Orsay, 1993), 1995, pp. 433–453. MR1364901 [26] V. Nekrashevych, Palyndromic subshifts and simple periodic groups of intermediate growth, preprint, arXiv:1601.01033. [27] A. S. Oliınyk, Free products of finite groups and groups of finitely automatic permutations, Tr. Mat. Inst. Steklova 231 (2000), no. Din. Sist., Avtom. i Beskon. Gruppy, 323–331 (Russian, with Russian summary); English transl., Proc. Steklov Inst. Math. 4 (231) (2000), 308–315. MR1841761 [28] D. V. Osin, Exponential radicals of solvable Lie groups, J. Algebra 248 (2002), no. 2, 790–805, DOI 10.1006/jabr.2001.9036. MR1882124 [29] Narutaka Ozawa, A functional analysis proof of Gromov's polynomial growth theorem, preprint, arXiv:1510.04223. [30] Hans Reiter and Jan D. Stegeman, Classical harmonic analysis and locally compact groups, 2nd ed., London Mathematical Society Monographs. New Series, vol. 22, The Clarendon Press, Oxford University Press, New York, 2000. MR1802924 [31] Yehuda Shalom, Harmonic analysis, cohomology, and the large-scale geometry of amenable groups, Acta Math. 192 (2004), no. 2, 119–185, DOI 10.1007/BF02392739. MR2096453 (2005m:20095) [32] , Rigidity of commensurators and irreducible lattices, Invent. Math. 141 (2000), no. 1, 1–54, DOI 10.1007/s002220000064. MR1767270 22 ANNA ERSCHLER AND NARUTAKA OZAWA [33] Yves Stalder, Fixed point properties in the space of marked groups, Limits of graphs in group theory and computer science, EPFL Press, Lausanne, 2009, pp. 171–182. MR2562144 [34] Romain Tessera, Volume of spheres in doubling metric measured spaces and in groups of polynomial growth, Bull. Soc. Math. France 135 (2007), no. 1, 47–64 (English, with English and French summaries). MR2430198 [35] Russ Thompson, The rate of escape for random walks on polycyclic and metabelian groups, Ann. Inst. Henri Poincar´e Probab. Stat. 49 (2013), no. 1, 270–287, DOI 10.1214/11-AIHP455 (English, with English and French summaries). MR3060157 A.E.: C.N.R.S., D ´EPARTEMENT DE MATH ´EMATIQUES ET APPLICATIONS, ´ECOLE NORMALE SUP ´ERIEURE, PSL RESEARCH INSTITUTE, 45, RUE D'ULM, 75005, PARIS, FRANCE E-mail address: [email protected] N.O.: RESEARCH INSTITUTE OF MATHEMATICAL SCIENCES, KYOTO UNIVERSITY, KITASHIRAKAWA- OIWAKE, KYOTO 606-8502, JAPAN E-mail address: [email protected]
1603.07494
1
1603
2016-03-24T09:30:37
Asymmetric Fuglede Putnam's Theorem for operators reduced by their eigenspaces
[ "math.FA" ]
Fuglede-Putnam Theorem have been proved for a considerably large number of class of operators. In this paper by using the spectral theory, we obtain a theoretical and general framework from which Fuglede-Putnam theorem may be promptly established for many classes of operators.
math.FA
math
ASYMMETRIC FUGLEDE PUTNAM'S THEOREM FOR OPERATORS REDUCED BY THEIR EIGENSPACES F. LOMBARKIA AND M. AMOUCH Abstract. Fuglede-Putnam Theorem have been proved for a considerably large number of class of operators. In this paper by using the spectral theory, we obtain a theoretical and general framework from which Fuglede-Putnam theorem may be promptly established for many classes of operators. 1. Introduction and basic definitions Fuglede-Putnam Theorem have been studied in the last two decades by several authors and most of them have essentially proved such theorem for special classes of operators. Many times the arguments used, to prove Fuglede-Putnam Theorem are similar, but in this paper we show that it is possible to bring back up this theorem from some general common properties. We use the spectral theory to obtain a theoretical and general framework from which Fuglede-Putnam theorem is established, and we can deduce that Fuglede-Putnam Theorem hold for many classes of operators. Let H be an infinite complex Hilbert space and consider two bounded linear operators A, B ∈ L(H). Let LA ∈ L(L(H)) and RB ∈ L(L(H)) be the left and the right multiplication operators, respectively, and denote by dA,B ∈ L(L(H)) either the elementary operator ∆A,B(X) = AXB − X or the generalized derivation δA,B(X) = AX − XB. Given T ∈ L(H), ker(T ), R(T ), σ(T ) and σp(T ) will stand for the null space, the range of T , the spectrum of T and the point spectrum of T . Recall that if M, N are linear subspaces of a normed linear space V, then M is orthogonal to N in the sense of Birkhoff, M ⊥N for short, if kmk ≤ km + nk for all m ∈ M and n ∈ N. It is known that if A, B ∗ ∈ L(H) are hyponormal operators, then dA,B satisfies the asymmetric Putnam Fuglede commutativity property ker dA,B ⊆ ker dA∗,B∗, hence ker dA,B⊥R(dA,B). From the fact that hyponormal operators are closed under translation and multiplication by scalars, B. P. Duggal in [12] deduced that if A and B ∗ are hyponormal, then ker(dA,B − λI) ⊆ ker(dA∗ ,B∗ − λI) holds for every complex number λ, where λ is the conjugate of λ. An operator T ∈ L(H) is said to be p-hyponormal, 0 < p ≤ 1, if T ∗2p ≤ T 2p, where T = (T ∗T ) 2 . An invertible operator T ∈ L(H) is log-hyponormal if log T ∗2 ≤ log T 2. In [12] B. P. Duggal proved that if A and B ∗ are p-hyponormal or log-hyponormal, then ker(dA,B −λI) ⊆ ker(dA∗,B∗ − λI) and ker(dA,B − λI)⊥R(dA,B − λI), for every complex number λ. An operator T ∈ L(H) is said to be w-hyponormal if (T ∗ 2 ≥ T ∗, It is shown in [2, 3] that the class of w-hyponormal properly contains see [19]. 2 T T ∗ 2 ) 1 1 1 1 2000 Mathematics Subject Classification. 47B47, 47B20, 47B10. Key words and phrases. Hilbert space; elementary operator; operators reduced by their eigenspaces, polaroid operators, property (β). 1 2 F. LOMBARKIA AND M. AMOUCH t 2 2 2 2 , 1 the class of p-hyponormal (0 < p ≤ 1,) and log-hyponormal. T. Furuta, M. Ito and T. Yamazaki [18] introduced a very interesting class A operators defined by − T 2 ≥ 0, and they showed that class A is a subclass of paranormal operators T (i.e., kT xk2 ≤ kT 2xkkxk, for all x ∈ H ) and contains w-hyponormal operators. An operator T ∈ L(H) is said to be class A(s, t), 0 < s, t if T ∗2t ≤ (T ∗tT 2sT ∗t) t+s . Then T ∈ A( 1 2 ) if an only if T is w-hyponormal and T ∈ A(1, 1) if an only if T is class A. I. H. Jeon and I. H. Kim [20] introduced quasi-class A operators defined by T ∗(T − T 2)T ≥ 0, as an extension of the notion of class A operators. K. Tanahash, I. H. Jeon, I, H. Kim and A. Uchiyama [25] introduced k-quasi-class A operators defined − T 2)T k ≥ 0, for a positive integer k as an extension of the notion by T ∗k(T of quasi-class A operators, for interesting properties of k-quasi-class A operators, called also quasi-class (A, k), see [16, 25]. In [9, Lemma 2.4], [5, Theorem 3.6] and [13, Lemma 2.4] the authors proved that if A, B ∗ ∈ L(H) are w-hyponormal operators with ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then dA,B satisfies the asymmetric Putnam Fuglede commutativity property ker dA,B ⊆ ker dA∗,B∗ . Recently B.P. Duggal, C. S. Kubruslly and I. H. Kim in [13, Theorem 2.5] have proved that if A ∈ A(s1, t1) and B ∗ ∈ A(s2, t2), 0 < s1, s2, t1, t2 ≤ 1 are such that ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then δA,B satisfies the asymmetric Putnam Fuglede commutativity property ker δA,B ⊆ ker δA∗,B∗ . In this paper we prove that all the precedent results are a consequence of our main results. Since the class of k-quasi-class A operators contains properly the class A(s, t), 0 < s, t operators, it therefore provides a unified approach in studying the asymmetric Putnam Fuglede commutativity property for k-quasi-class A operators. Now we recall some definitions Definition 1.1. An operator T ∈ L(H) has Bishop's property (β) if for every open set U ⊂ C and every sequence of analytic functions fn : U → X, with the property that (T − λI)fn(λ) → 0 uniformly on every compact subset of U, it follows that fn → 0, again locally uniformly on U. Bishop's property (β) implies Dunford property (C), also T satisfies property (β) if and only if T ∗ satisfies property (δ) [22, Theorem 2.5.5]. For more information on property (β), property (δ) and Dunford's condition (C) we refer the interested reader to [22]. Recall that the ascent p(T ) of an operator T , is defined by p(T ) = inf{n ∈ N : ker T n = ker T n+1} and the descent q(T ) = inf{n ∈ N : R(T n) = R(T n+1)}, with inf ∅ = ∞. It is well known that if p(T ) and q(T ) are both finite then p(T ) = q(T ). We denote by Π(T ) = {λ ∈ C : p(T − λI) = q(T − λI) < ∞} the set of poles of the resolvent. An operator T ∈ L(H) is called Drazin invertible if and only if it has finite ascent and descent. The Drazin spectrum of an operator T is defined by σD(T ) = {λ ∈ C : T − λI is not Drazin invertible}. In the sequel we shall denote by accD and isoD, the set of accumulation points and the set of isolated points of D ⊂ C, respectively Definition 1.2. An operator T ∈ L(H) is said to be polaroid if isoσ(T ) ⊆ Π(T ). It is easily seen that, if T ∈ L(H) is polaroid, then Π(T ) = E(T ), where E(T ) is the set of eigenvalues of T which are isolated in the spectrum of T 3 An important subspace in local spectral theory is the the quasinilpotent part of T is defined by H0(T ) = {x ∈ H : lim n→∞ kT n(X)k 1 n = 0}. It is easily seen that ker T n ⊂ H0(T ) for every n ∈ N, see [1] for information on H0(T ). The range-kernel orthogonality of dA,B in the sense of G. Birkhoff was studied by numerous mathematicians, see [4, 5, 10, 21, 26] and the references therein. A sufficient condition guaranteeing the range-kernel orthogonality of dA,B is that ker dA,B ⊆ ker dA∗,B∗ [10]. The main objective of this paper is to give sufficient conditions to have ker(dA,B − λI) ⊆ ker(dA∗,B∗ − λI), for every complex number λ. After section one where several basic definitions and facts will be recalled, in section two, we prove that if A and B ∗ are reduced by each of its eigenspaces, polaroid and have property (β), then ker(dA,B −λI) ⊆ ker(dA∗,B∗ −λI) for every complex number λ and that the elementary operator dA,B satisfies the range-kernel orthogonality kXk ≤ kX − (dA,B − λI)Y k, for all X ∈ ker(dA,B − λI) and Y ∈ L(H). We apply the results obtained to k-quasi-class A operators. We prove that dA,B satisfies the asymmetric Putnam Fuglede commutativity property ker dA,B ⊆ ker dA∗,B∗, if A and B ∗ are k-quasi-class A operator with ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B. This generalizes results given in [12, Theorem 2.3], [5, Theorem 3.6], [13, Lemma 2.4] and [13, Theorem 2.5]. 2. Main results Let T ∈ L(H) be reduced by each of its eigenspaces. If we let M = W{ker(T − µI), µ ∈ σp(T )} (where W(.) denotes the closed linear span, it follows that M reduces T . Let T1 = T M and T2 = T M ⊥. By [7, Proposition 4.1] we have • T1 is normal with pure point spectrum, • σp(T1) = σp(T ), • σ(T1) = clσp(T1)(here cl denotes closure), • σp(T2) = ∅. The classical and most known form of the Fuglede Putnam theorem is the fol- lowing Theorem 2.1. [15, 24] If X, A and B are bounded operators acting on complex Hilbert space H such that A and B are normal, then AX = XB =⇒ A∗X = XB ∗. Theorem 2.2. Suppose that A, B ∗ ∈ L(H) are reduced by each of its eigenspaces, polaroid and have property (β), then ker(δA,B − λI) ⊆ ker(δA∗,B∗ − λI), ∀λ ∈ C. Proof. Since A and B ∗ are reduced by each of its eigenspaces, then then there exists M1 = _{ker(A − βI), β ∈ σp(A)} and M2 = H ⊖ M1 on the one hand and N1 = _{ker(B ∗ − αI), α ∈ σp(B ∗)} and N2 = H ⊖ N1 4 F. LOMBARKIA AND M. AMOUCH on the other hand such that A and B have the representations A = (cid:18) A1 0 A2 (cid:19) on H = M1 ⊕ M2 0 and B = (cid:18) B1 0 B2 (cid:19) on H = N1 ⊕ N2 0 Recall from [14] that σ(δA,B) = σ(A) − σ(B), we consider the following cases: Case 1: If λ ∈ C\σ(δA,B), then ker(δA,B − λI) = {0} and hence ker(δA,B − λI) ⊆ ker(δA∗,B∗ − λI). Case 2: If λ ∈ isoσ(δA,B), then there exists finite sequences {µi}n where µi ∈ isoσ(A) and νi ∈ isoσ(B) such that λ = µi − νi, for all 1 ≤ i ≤ n. i=1 and {νi}n i=1, λ 6∈ σ(δAi,Bj ) for all 1 ≤ i, j ≤ 2 other than i = j = 1. Consider X ∈ ker(δA,B − λI) such that X : N1 ⊕ N2 −→ M1 ⊕ M2 have the representation X = [Xkl]2 k,l=1. Hence (δA,B − λI)(X) = (cid:18) (δA1,B1 − λI)(X11) (δA2,B1 − λI)(X21) (δA1,B2 − λI)(X12) (δA2,B2 − λI)(X22) (cid:19) = 0. Observe that δAi,Bj − λI is invertible for all 1 ≤ i, j ≤ 2 other than i = j = 1. Hence X22 = X21 = X12 = 0. Since A1 − µi and B1 − νi are normal, it follows from Fuglede-Putnam theorem that (A∗ 1 − µi)X11 − X11(B ∗ 1 − νi) = 0, consequently X = X11 ⊕ 0 ∈ ker(δA∗,B∗ − λI). Case 3: If λ ∈ accσ(δA,B), then there exists µ ∈ σ(A) and ν ∈ σ(B) such that λ = µ − ν ∈ (σ(A) − accσ(B)) or λ = µ − ν ∈ (accσ(A) − σ(B)). Since A and B are polaroid, then µ ∈ accσ(A) = σD(A) and ν ∈ accσ(B) = σD(B), it follows from [8, Theorem 2.4] that µ ∈ σD(A1) ∪ σD(A2) and ν ∈ σD(B1) ∪ σD(B2), since σp(A2) = σp(B ∗ 2 ) = ∅, then σD(A2) = σ(A2) and σD(B2) = σ(B2). Hence µ ∈ σD(A1) ∪ σ(A2) and ν = σD(B1) ∪ σ(B2). Let X ∈ ker(δA,B − λI) such that X : N1 ⊕ N2 −→ M1 ⊕ M2 have the representation X = [Xkl]2 k,l=1. Hence (δA,B − λI)(X) = (cid:18) (δA1,B1 − λI)(X11) (δA2,B1 − λI)(X21) (δA1,B2 − λI)(X12) (δA2,B2 − λI)(X22) (cid:19) = 0. We consider the following cases • µ ∈ σ(A1) and ν ∈ σD(B1), or • µ ∈ σ(A1) and ν ∈ σ(B2), or • µ ∈ σ(A2) and ν ∈ σD(B1), or • µ ∈ σ(A2) and ν ∈ σ(B2). or • µ ∈ σD(A1) and ν ∈ σ(B1), or 5 • µ ∈ σD(A1) and ν ∈ σ(B2), or • µ ∈ σ(A2) and ν ∈ σ(B1). We start by studying these cases • If µ ∈ σ(A1) and ν ∈ σD(B1). Observe that δAi,Bj − λI is invertible for all 1 ≤ i, j ≤ 2 other than i = j = 1, we have X ∈ ker(δA,B − λI). Hence X12 = X21 = X22 = 0. Since A1 − µ and B1 − ν are normal, it follows from Fuglede-Putnam theorem that (A∗ 1 − µi)X11 − X11(B ∗ 1 − νi) = 0, consequently X = X11 ⊕ 0 ∈ ker(δA∗,B∗ − λI). • If µ ∈ σ(A1) and ν ∈ σ(B2), we have X ∈ ker(δA,B − λI), then X22 = X21 = X11 = 0. Since µ ∈ σ(A1) = clσp(T1), then there exist a sequence (µn) ∈ σp(A1) such that µn −→ µ, so for all n ∈ N there exist a non 1 − µn). We have (A1 − µ)X12 = null vector xn ∈ ker(A1 − µn) = ker(A∗ X12(B2 − ν) this implies that (B ∗ 1 − µ), consequently 1 − µ)xn. Since µn is an eigenvalue, so there exists 12(A∗ (B ∗ 2 −ν) 1 −µn), such that lim kxn −zk = 0. Since (B ∗ a no null vector z ∈ ker(A∗ is injective, we obtain X12 = 0, hence 12xn = X ∗ 2 − ν)X ∗ 2 − ν)X ∗ 12 = X ∗ 12(A∗ X = 0 ∈ ker(δA∗,B∗ − λI). 1 ) = σ(B ∗ • If µ ∈ σ(A2) and ν ∈ σD(B1), we have X ∈ ker(δA,B − λI), then X11 = X22 = X12 = 0 and (A2 − µ)X21 = X21(B1 − ν). Since ν ∈ σD(B1), then ν ∈ σD(B ∗ 1 ), so there exist a sequence (νn) ∈ σp(B ∗ 1 ) such that νn −→ ν, so for all n ∈ N there exist a non null vector xn ∈ 1 − νn) = ker(B1 − νn). We have (A2 − µ)X21 = X21(B1 − ν), ker(B ∗ consequently (A2 − µ)X21xn = X21(B1 − ν)xn. Since νn is an eigenvalue, so there exists a no null vector z ∈ ker(B1 − νn), such that lim kxn − zk = 0. Since (A2 − µ) is injective, we obtain X21 = 0, hence 1 )\E(B ∗ X = 0 ∈ ker(δA∗,B∗ − λI). • If µ ∈ σ(A2) and ν ∈ σ(B2). Since A has property (β) it follows from [6, Remarks 3.2] that A2 has property (β), applying [1, Theorem 2.20] we get H0(A2−µ) is closed and from [22, Proposition 1.2.20] That σ(A2H0(A2−µ)) ⊆ {µ}. If σ(A2H0(A2−µ)) = ∅, then H0(A2−µ) = {0}, the case σ(A2H0(A2−µ)) = {µ} is not possible, since the operator A2 does not contains isolated points. Hence H0(A2 − µ) = {0}, we have X ∈ ker(δA,B − λI), then X21 = X12 = X11 = 0 and (A2−µ)X22 = X22(B2−ν), this implies that, if t ∈ H0(B2−ν), then X22t ∈ H0(A2 − µ) = {0}. Hence X22t = 0. Since t ∈ H0(B2 − ν), using properties of quasinilpotent part, we get (B2 − ν)(t) ∈ H0(B2 − ν), consequently N2 = H0(B2 − ν). So X22 = 0, hence X = 0 ∈ ker(δA∗,B∗ − λI). The other cases will be proved similarly. (cid:3) Theorem 2.3. Let A, B ∈ L(H). If all the eigenvalues of A, B ∗ are reduced by each of its eigenspaces, polaroid and have property (β), then ker(∆A,B − λI) ⊆ ker(∆A∗,B∗ − λI), ∀λ ∈ C. 6 F. LOMBARKIA AND M. AMOUCH Proof. Since A and B ∗ are reduced by each of its eigenspaces, then then there exists M1 = _{ker(A − βI), β ∈ σp(A)} and M2 = H ⊖ M1 on the one hand and N1 = _{ker(B ∗ − αI), α ∈ σp(B ∗)} and N2 = H ⊖ N1 on the other hand such that A and B have the representations A = (cid:18) A1 0 A2 (cid:19) on H = M1 ⊕ M2 0 and B = (cid:18) B1 0 B2 (cid:19) on H = N1 ⊕ N2. 0 Recall from [14] that σ(∆A,B) = σ(A)σ(B) − {1}. We consider the following cases. Case 1: If λ ∈ C\σ(∆A,B), the result is immediate. Case 2: If λ ∈ isoσ(∆A,B) and λ 6= −1, then there exists finite sequences {µi}n i=1 and {νi}n i=1, where µi ∈ isoσ(A) and νi ∈ isoσ(B) such that λ = µiνi − 1, for all 1 ≤ i ≤ n. λ 6∈ σ(∆Ai,Bj ) for all 1 ≤ i, j ≤ 2 other than i = j = 1. Let X ∈ ker(∆A,B − λI) such that X : N1 ⊕ N2 −→ M1 ⊕ M2 have the representation X = [Xkl]2 k,l=1. Hence (∆A,B − λI)(X) = (cid:18) (∆A1,B1 − λI)(X11) (∆A2,B1 − λI)(X21) (∆A1,B2 − λI)(X12) (∆A2,B2 − λI)(X22) (cid:19) = 0. Observe that ∆Ai,Bj − λI is invertible for all 1 ≤ i, j ≤ 2 other than i = j = 1. Hence X22 = X21 = X12 = 0. Since A1 and B1 are normal, it follows from Fuglede- Putnam theorem and [11, Theorem 2] that 1 1 + λ A∗ 1X11B ∗ 1 − X11 = 0, consequently X = X11 ⊕ 0 ∈ ker(∆A∗,B∗ − λI). If λ = −1, we consider the case −1 ∈ isoσ(∆A,B), that is 0 ∈ isoσ(LARB), hence either 0 ∈ isoσ(A) and 0 ∈ isoσ(B) or 0 ∈ isoσ(A) and 0 6∈ σ(B) or 0 ∈ isoσ(B) and 0 6∈ σ(A). If 0 ∈ isoσ(A) and 0 ∈ isoσ(B). Let X : N1 ⊕ N2 −→ M1 ⊕ M2, have the matrix representation X = [Xkl]2 If X ∈ ker(LARB), then (cid:18) A1X11B1 A1X12B2 A2X21B1 A2X22B2 (cid:19) = 0, it follows that X22 = X21 = X12 = 0, and 1 . Thus X ∈ ker(LA∗RB∗ ). The proofs of k,l=1. RB∗ A1X11B1 = 0, hence X11 ∈ ker LA∗ the other remaining cases are similar. 1 Case 3: If λ ∈ accσ(∆A,B) = (accσ(A)σ(B) − 1) ∪ (σ(A)accσ(B) − 1), and λ = −1 then 0 ∈ σ(A)accσ(B) or 0 ∈ accσ(A)σ(B). Since A and B are polaroid, then 0 ∈ accσ(A) = σD(A) and 0 ∈ accσ(B) = σD(B). • If 0 ∈ σ(A1) and 0 ∈ σD(B1). Observe that LAi RBj is invertible for all 1 ≤ i, j ≤ 2 other than i = j = 1, we have X ∈ ker(LARB). Hence X12 = X21 = X22 = 0. Since A1 and B1 are normal, it follows from Fuglede-Putnam theorem that 7 A∗ 1X11B ∗ 1 = 0, consequently X = X11 ⊕ 0 ∈ ker(LA∗ RB∗ ). • If 0 ∈ σ(A1) and 0 ∈ σ(B2). Let X ∈ ker(LARB), then X22 = X21 = X11 = 0. We have A1X12B2 = 0 this implies that A1X12 = 0. Since A1 is normal, then A∗ 1X12 = 0, consequently A∗ 2 = 0. Hence 1X12B ∗ X = (cid:18) 0 X12 0 (cid:19) ∈ ker(LA∗RB∗ ). 0 • If 0 ∈ σ(A2) and 0 ∈ σD(B1), this case will be proved similarly as the precedent one. • If 0 ∈ σ(A2) and 0 ∈ σ(B2). Let X ∈ ker(LARB), then X12 = X21 = X11 = 0. Since A2 and B ∗ 2 are injective, then X22 = 0. Hence The other cases will be proved similarly. X = 0 ∈ ker(LA∗ RB∗ ). Case 4: If λ ∈ accσ(∆A,B) = (accσ(A)σ(B) − 1) ∪ (σ(A)accσ(B) − 1) and λ 6= −1, then there exists µ ∈ σ(A) and ν ∈ σ(B) such that λ = µν ∈ (σ(A)accσ(B)−1) or λ = µν ∈ (accσ(A)σ(B) − 1). • If µ ∈ σ(A1) and ν ∈ σD(B1). Observe that ∆Ai,Bj − λI is invertible for all 1 ≤ i, j ≤ 2 other than i = j = 1, we have X ∈ ker(∆A,B − λI). Hence X12 = X21 = X22 = 0. Since A1 and B1 are normal, it follows from Fuglede-Putnam theorem and [11, Theorem 2] that 1 1 + λ consequently A∗ 1X11B ∗ 1 − X11 = 0, X = X11 ⊕ 0 ∈ ker(∆A∗,B∗ − λI). • If µ ∈ σ(A1) and ν ∈ σ(B2). Let X ∈ ker(∆A,B − λI), then X22 = X21 = X11 = 0. Since µ ∈ σ(A1) = clσp(T1), then there exist a sequence (µn) ∈ σp(A1) such that µn −→ µ, so for all n ∈ N there exist a non null vector xn ∈ ker(A1 − µn) = ker(A∗ 1 − µn). We have B ∗ 2 X ∗ 12(A∗ 1 − µ) − (1 + λ − µB ∗ 2 )X ∗ 12 = 0, consequently B ∗ 2 X ∗ 12(A∗ 1 − µ)xn − (1 + λ − µB ∗ 2 )X ∗ 12xn = 0. Since µn is an eigenvalue, so there exists a no null vector z ∈ ker(A∗ such that lim kxn − zk = 0. Since B ∗ 1 − µn), 2 is injective, we obtain X12 = 0, hence X = 0 ∈ ker(∆A∗,B∗ − λI). • If µ ∈ σ(A2) and ν ∈ σD(B1), this case will be proved similarly as the precedent one. 8 F. LOMBARKIA AND M. AMOUCH • If µ ∈ σ(A2) and ν ∈ σ(B2). Since A has property (β) it follows from [6, Remarks 3.2] that A2 has property (β), applying [1, Theorem 2.20] we get H0(A2−µ) is closed and from [22, Proposition 1.2.20] That σ(A2H0(A2−µ)) ⊆ {µ}. If σ(A2H0(A2−µ)) = ∅, then H0(A2−µ) = {0}, the case σ(A2H0(A2−µ)) = {µ} is not possible, since the operator A2 does not contains isolated points. Hence H0(A2 − µ) = {0}, we have X ∈ ker(∆A,B − λI), then X21 = X12 = X11 = 0 and A2X22B2 − (1 + λ)X22 = 0, this implies that (A2 − µ)X22(B2 − ν) + ν(A2 − µ)X22 + µX22(B2 − ν) = 0. if t ∈ H0(B2 − ν), then X22t ∈ H0(A2 − µ) = {0}. Hence X22t = 0. Since t ∈ H0(B2 −ν), using properties of quasinilpotent part, we get (B2 −ν)(t) ∈ H0(B2 − ν), consequently N2 = H0(B2 − ν). So X22 = 0, hence X = 0 ∈ ker(∆A∗,B∗ − λI). The other cases will be proved similarly. (cid:3) Theorem 2.4. Suppose that A, B ∗ ∈ L(H) are reduced by each of its eigenspaces, polaroid and have property (β), then R(dA,B − λI) is orthogonal to ker(dA,B − λI), for all λ ∈ C. Proof. Follows from [10, Lemma 4] (cid:3) Corollary 2.5. [12, Lemma 2.1] Suppose that A, B ∗ ∈ L(H) are p-hyponormal or log-hyponormal, then ker(dA,B − λI) ⊆ ker(dA∗,B∗ − λI), ∀λ ∈ C Corollary 2.6. [9, Lemma 2.4] Let A, B ∗ ∈ L(H) be w-hyponormal operators such that ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then ker dA,B ⊆ ker dA∗,B∗. Corollary 2.7. [5, Theorem 3.6],[13, Lemma 2.4] Let A, B ∗ ∈ L(H) be w-hyponormal operators such that ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then ker δA,B ⊆ ker δA∗,B∗ . Corollary 2.8. [13, Theorem 2.5] Let A, B ∗ ∈ L(H). If A ∈ A(s1, t1) and B ∗ ∈ A(s2, t2), 0 < s1, s2, t1, t2 ≤ 1 are such that ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then ker δA,B ⊆ ker δA∗,B∗ . As a nice application of our main results the Fuglede Putnam theorem for k- quasi-class A operators which contains all the precedent classes of operators. Theorem 2.9. Let A, B ∗ ∈ LH) be k-quasi-class A operators, then ker(dA,B − λI) ⊆ ker(dA∗,B∗ − λI), for all non null complex number λ. Proof. We know from [17, Theorem 2.4] that k-quasi-class A operators are polaroid and from [25, Lemma 11] that k-quasi-class A operators have property (β). Since by [25, Lemma 13], A and B ∗ are reduced by each of its eigenspaces, then the conclusion follows from Theorem 2.2 and Theorem 2.3. (cid:3) 9 Theorem 2.10. Let A, B ∗ ∈ L(H) be k-quasi-class A operators such that ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B, then ker dA,B ⊆ ker dA∗,B∗. Proof. The conditions ker A ⊆ ker A∗ and ker B ∗ ⊆ ker B implies that 0 is normal eigenvalue of both A and B ∗. it follows from Theorem 2.2 and Theorem 2.3 that ker dA,B ⊆ ker dA∗,B∗. (cid:3) References [1] P. Aiena Fredholm and local spectral theory with applications to multipliers, Kluewer Acad. publishers, 2004. [2] A. Aluthge and D. Wang, On w-hyponormal operators II , Intergr. Equ. Oper. Theory, 37(2000), 324 -- 331. [3] A. Aluthge and D. Wang, An operator inequality which implies paranormality, Math. In- equality Appl. 2(1999), 113 -- 119. [4] M. Amouch, A note on the range of generalized derivation, Extracta Math. 21(2006), 149 -- 157. [5] A. Bachir and F. Lombarkia, Fuglede Putnam's theorem for w-hyponormal operators, Math. Inequal. App. 12(2012), 777 -- 786. [6] C. Benhida, E. H. Zerouali and H. Zguitti, Spectra of upper triangular operator matrices, Proc. Amer. Math. Soc. 133 (2005), 3013 -- 3020. [7] S. K. Berberian, The Weyl spectrum of an operator, Indiana Univ. Math. J 20(1970), 529 -- 544. [8] M. Berkani and H. Zariouh, Weyl type-theorems for direct sums, Bull. Korean Math. Soc. 49, (2012), No. 5, pp. 10271040. [9] M. cho, S. V. Djordjevi´c, B. P. Duggal and T. Yamazaki, On an elementary operator with w-hyponormal operator entries, Linear Algebra Appl. 433 (2010), 2070 -- 2079. [10] B. P. Duggal, Range kernel orthogonality of derivations, Linear Algebra Appl. 304(2000), 103-108. [11] B. P. Duggal, A remark on generalized Putnam-Fuglede theorems, Proc. Amer. Math. Soc. 129 (2000), 83 -- 87. [12] B. P. Duggal, An elementary operator with log-hyponormal, p-hyponormal entries, Linear Algebra Appl. 428 (2008), 1109 -- 1116. [13] B. P. Duggal, C. S. Kubrursly and I. H. Kim Bishop's property β, a commutativity theorem and the dynamics of class A(s, t) operators, J. Math. Anal. Appl. 427 (2015), 107 -- 113. [14] M. R. Embry and M. Rosenblum, Spectra, tensor products and linear operator equations, Pacific J. Math. 53 (1974), 95-107. [15] B. Fuglede, A commutativity theorem for normal operators, Proc. Nat. Acad. Sci. U. S. A, 36(1950), 35 -- 40. [16] F. G. Gao and X. C. Fang, On k-quasiclass A operators, Math. Inequal. Appl, 2009(2009), Article ID 921634, 1 -- 10. [17] F. G. Gao and X. C. Fang, Weyl's theorem for algebraically k-quasiclass A operators, Opscula Mathematica, 32 (2012), 125 -- 135. [18] T. Furuta, M. Ito and T. Yamazaki, A subclass of operators including class of log-hyponormal and several classes, Sci. Math. 1, 3(1998), 389 -- 403. [19] M. Ito, T. Yamazaki, Relation between two inequalities (cid:16)B r 2 ApB r 2 (cid:17) r p+r ≥ Br and Ap ≥ p+r and their applications, Integral Equations Operator Theory, 44 (2002), p p (cid:16)A 2 Br A 442 -- 450. p 2 (cid:17) [20] I. H. Jeon and I. H. Kim, On operators satisfying T ∗T 2T ≥ T ∗T 2T , Linear Algebra Appl. 418 (2006), 854 -- 862. [21] D. Keckic, Orthogonality of the range and the kernel of some elementary operators, Proc. Amer. Math. Soc. 128 (2000), 3369 -- 3377. [22] K. B. Laursen and M. M. Neumann, An introduction to local spectral theory, Lon. Math. Soc. Monographs, Oxford Univ. Press, 2000. 10 F. LOMBARKIA AND M. AMOUCH [23] F. Lombarkia, Generalized Weyl's theorem for an elementary operator, Bull. Math. Anal. Appl. 3 (4) (2011), 123 -- 131. [24] C. R. Putnam, On normal operators in Hilbert space, Amer. J. Math 73 (1951), 357 -- 362. [25] K. Tanahashi, I. H. Jeon, I. H. Kim and A. Uchiyama, Quasinilotent part of class A or (p,k)-quasihyponormal operators, Oper. Theory Adv. Appl, 187 (2008), 199 -- 210. [26] A. Turnsek, Generalized Anderson's inequality, J. Math. An. Appl, 263 (2001), 121 -- 134. Farida Lombarkia Department of Mathematics, Faculty of Science, University of Batna, 05000, Batna, Algeria. E-mail address: [email protected] Mohamed AMOUCH Department of Mathematics and informatic University Chouaib Doukkali, Faculty of Sciences, Eljadida. 24000, Eljadida, Morocco. E-mail address: [email protected]
1201.5592
1
1201
2012-01-26T18:02:16
Agler-Commutant Lifting on an Annulus
[ "math.FA", "math.OA" ]
The main result is a test function style commutant lifting theorem for an annulus A. The test functions are the minimal inner functions for A. The model space is the Sarason Hardy Hilbert space for A uniquely determined by the fact that its reproducing kernel has no zeros.
math.FA
math
AGLER-COMMUTANT LIFTING ON AN ANNULUS SCOTT MCCULLOUGH∗ AND SAIDA SULTANIC Abstract. This note presents a commutant lifting theorem (CLT) of Agler type for the annulus A. Here the relevant set of test functions are the minimal inner functions on A - those analytic functions on A which are unimodular on the boundary and have exactly two zeros in A - and the model space is determined by a distinguished member of the Sarason family of kernels over A. The ideas and constructions borrow freely from the CLT of Ball, Li, Timotin, and Trent [14] and Archer [11] for the polydisc, and Ambrozie and Eschmeier for the ball in Cn [3], as well as generalizations of the de Branges-Rovnyak construction like found in Agler [5] and Ambrozie, Englis, and Muller [4]. It offers a template for extending the result result in [29] to infinitely many test functions. Among the needed new ingredients is the formulation of the factorization implicit in the statement of the results in [14], [11] and [29] in terms of certain functional Hilbert spaces of Hilbert space valued functions. 1. Introduction Results going back to [5] and including [7], [15], [16], [10] [4], [3] [20], [19] among others view the starting point for Agler-Pick interpolation as a collection of func- tions Ψ, called test functions. Roughly speaking one constructs an operator algebra whose norm is as large as possible subject to the condition that each ψ ∈ Ψ is con- tractive. The corresponding Agler-Schur class, or Ψ-Agler-Schur class, is then the unit ball of this operator algebra and interpolation is within this class. The by now classical example is that of Agler-Pick interpolation in the d-fold polydisc Dd ⊂ Cd with Ψ = {z1, . . . , zd}, where the zj are the coordinate functions [5][7]. In this case the unit ball of the resultant operator algebra of functions on Dd is known as the Agler-Schur class, often denoted Sd. For d = 1, 2 this operator algebra is the same as H ∞(Dd), but generally Sd and H ∞(Dd) are different. The literature contains many articles on the Agler-Schur class and its operator-valued generalizations. A sample of references include [12] [16][25][8]. Of special relevance for this paper is the work of Ambrozie [10] and the subsequent articles [20] and [19], where the set of test functions Ψ is allowed to be infinite with a compact Hausdorff topology. It has long been known that Pick interpolation is a special case of commutant lifting [31] [21] [23] [30]. In this spirit Ball, Li, Timotin, and Trent [14] formulate and prove an Agler-Pick type commutant lifting theorem for the polydisc. Significant refinements of both the statements and proofs of this result appear in the work of Archer [11]. Ambrozie and Eschmeier [3] establish a related CLT for the unit ball in Cn. In [29] we establish a generalization of these results to the case of a 1991 Mathematics Subject Classification. 47A20 (Primary), 47A48, 47A57, 47B32 (Secondary). Key words and phrases. commutant lifting, Agler-Schur class, annulus . ∗ Research supported by NSF grants 0457504 and 0758306. 2 S. MCCULLOUGH AND S. SULTANIC finite collection Ψ together with a distinguished reproducing kernel Hilbert space H 2(k), unlocking the prior tight connection between the coordinate (test) functions {z1, . . . , zd} and the kernel k for the Hardy space H 2(Dd) in the case of the polydisc. In this more general context, the lack of an orthonormal basis explicitly expressible in terms of the test functions necessitated a number of innovations. In this article we pursue an Agler-Pick type commutant lifting theorem with Ψ the infinite collection of minimal inner functions on an annulus A - those with uni- modular boundary values and exactly two zeros inside - and H 2(k) a distinguished choice of Hardy Hilbert space on A - distinguished by the fact that k(z, w) is the only Sarason kernel for A which does not vanish for (z, w) ∈ A × A. In addition to certain measure theoretic considerations necessitated by the infinite collection of test functions, it also turns out that some structures not apparent or exploited in the case of finite test functions become important. We have borrowed freely from [14], [11], [3], [5] [4] and of course [29]. We thank the referee for many substantive suggestions which markedly improved the exposition. 2. Preliminaries and Main Result Fix 0 < q < 1 and let A denote the annulus {z ∈ C : q < z < 1}. The boundary of the annulus comes in two parts, the outer boundary B0 = {z = 1} and the inner boundary B1 = {z = q}. As is customary, D denotes the unit disc. 2.1. The test functions. The minimal inner functions on A are those (non- constant) analytic functions φ : A → D whose boundary values are unimodular and have the minimum number of zeros - two - in A. Up to canonical normaliza- tions, they can be parametrized by the unit circle. If ψ : A → D is a minimal inner function normalized by ψ(√q) = 0 and ψ(1) = 1, z = √q} (see then the second zero w of ψ must lie on the circle T = {z : Section 11). Conversely, if w is a point on this circle T, then there is a (uniquely determined) minimal inner function ψw with ψw(√q) = 0 = ψw(w) normalized by In the case w = √q, this zero has multiplicity two. Hence, letting ψw(1) = 1. Ψ = {ψw : w ∈ T} ⊂ H ∞(A), there is a canonical bijection T → Ψ given by w 7→ ψw which turns out to be a homeomorphism. For z ∈ A, let E(z) denote the corresponding point evaluation on Ψ. Thus E(z) : Ψ → D is the continuous function defined by E(z)(ψ) = ψ(z). 2.2. Transfer functions and the Schur class. In the test function approach to interpolation and commutant lifting, those functions built from the test functions as a transfer function of a unitary colligation play a key role and are known as Agler-Schur class functions. Definition 2.1. A Ψ-unitary colligation is a tuple Σ = (ρ, A, B, C, D,E,H) where (i) E and H are Hilbert spaces; (ii) ρ : C(T) → B(E) is a unital representation; and (iii) the block operator U = (cid:18)A B C D(cid:19) : E ⊕ H → E ⊕ H is unitary. AGLER-COMMUTANT LIFTING ON AN ANNULUS 3 The corresponding transfer function is the function on A with values in B(H) given by where Z : A → B(E) is the function ρ(E(z)). WΣ = D + C(I − ZA)−1ZB, The collection S(A,H) of functions F : A → B(H) with a transfer function representation is called the Schur-Agler class. It coincides with the usual unit ball of H ∞(A) for scalar-valued functions [19] H = C). We believe that, using Agler's rational dilation theorem [5] and arguments like those in [19] or those of [18], the same is true for operator-valued H ∞(A), but postpone further consideration of this issue. 2.3. A Hardy space of the annulus. Results of Sarason [31], Abrahamse and Douglas [2], and Abrahamse [1] among others identify a certain one parameter family of Hardy Hilbert spaces over the annulus which, collectively, play the same role for A as the classical Hardy space plays for D. For t > 0, let µt denote the measure on the boundary of A which is the usual normalized arclength measure on the outer boundary B0 (so that µt(B0) = 1), but is t times normalized arclength measure on the inner boundary B1 (so that µt(B1) = t). Let H 2 t (A) denote the Hardy Hilbert space obtained by closing up functions analytic in a neighborhood of the closure of A in L2(µt). t = H 2 It is straightforward to check that the set (1) ζn = is an orthonormal basis for H 2 zn , n ∈ Z, p1 + tq2n t . In particular, (2) k(z, w; t) = Xn∈Z (zw∗)n 1 + tq2n is the reproducing kernel for H 2 t . Each ϕ ∈ H ∞(A) determines an operator Mt(ϕ) of multiplication by ϕ on H 2 t whose adjoint satisfies Mt(ϕ)∗k(·, w; t) = ϕ(w)∗k(·, w; t). q2t 7→ H 2 It also intertwines Mtq2 and Mt; From equation (2), it is evident that U : H 2 t given by U f = zf is unitary. i.e., U Mtq2(ϕ) = Mt(ϕ)U . Modulo this equivalence, the collection (H 2 t , Mt) is a family of representations of H ∞(A) parametrized by the unit circle. Up to unitary equivalence, these are Sarason's Hardy spaces of the annulus [31] that appear in [1]. They are also, over A, the rank one bundle shifts of Abrahamse and Douglas [2]. The kernel functions k(z, w; t) have theta function representations from which the proposition below follows. From here on, let k(z, w) = k(z, w; 1) and H 2(A) = H 2 1 (A). This is our distinguished Hardy space and its kernel. Set kw(z) = k(z, w). Proposition 2.2. The kernel k(·,·) doesn't vanish in the annulus; i.e., for z, w ∈ A, k(z, w) 6= 0, but it does vanish on the boundary as k(1,−1) = 0. Further, there is a constant C′ > 0 independent of z and w in A so that 1 k(z, w) = C′k(z,−w). If t 6= q2m (for any m), then there exists z, w ∈ A such that k(z, w; t) = 0. 4 S. MCCULLOUGH AND S. SULTANIC A proof of the proposition appears in Section 10. In the sequel, frequent use will be made of the Hilbert space tensor product H 2(k) ⊗ H, where H is itself a Hilbert space. A convenient way to define this Hilbert space is as those (Laurent) series h = Xj∈Z ζj ⊗ hj, for which P khjk2 converges. The inner product is defined by hh, gi = Xhhj, gji. For z ∈ A, the sum h(z) = X ζj (z) ⊗ hj converges absolutely. It follows that, for a fixed g ∈ H, hh(z), giH = hh, kz ⊗ gi. A function W : A → B(H) defines a contraction operator MW on H 2(k) ⊗ H by (3) M ∗ W [kz ⊗ g] = kz ⊗ W (z)∗g if and only if the (operator-valued) kernel A × A ∋ (z, w) 7→ (I − W (z)W (w)∗)k(z, w) is positive semi-definite [9][13]. Because, for h ∈ H 2(k) ⊗ H, hMW h, kz ⊗ gi = hW (z)h(z), gi, it is natural to write (MW h)(z) = W (z)h(z) = (W h)(z) to denote the operator MW and identify it with the function W (z). The following standard lemma will be used often and without comment in the sequel. Lemma 2.3. If Wn : A → B(H) is a sequence of functions which converge pointwise (in the norm topology) to W and if {Wn} is uniformly bounded, then MW is bounded and the sequence (MWn ) converges WOT to W . For expository purposes, we record the following nice relation between the kernel k and the test functions. Proposition 2.4 ([26, 27]). For the test function ψ with zeros √q and w with w = √q the kernel A × A ∋ (z, w) 7→ k(z, w)(1 − ψ(z)ψ(w)∗) = h(I − MψM ∗ ψ)k(·, w), k(·, z)i has rank two and is positive semi-definite. Further, Mψ is a shift of multiplicity two and the kernel of I − MψM ∗ ψ is the span of k(·,√q) and k(·, w) (except of course when w = √q when we must resort to using a derivative). AGLER-COMMUTANT LIFTING ON AN ANNULUS 5 1 2.4. Some representations and the functional calculus. Let T denote an operator on a Hilbert space M with σ(T ) ⊂ A. This spectral condition (as opposed to the more liberal σ(T ) ⊂ A) is imposed because we wish to consider 1 k (T, T ∗) and k does not extend to be analytic in z and w∗ beyond A × A. Let T also denote the corresponding representation T : H ∞(A) → B(M), given by T (f ) = f (T ). We also use the notation Tf = f (T ). Note that T is weakly continuous in the sense that if f, fn ∈ H ∞(A) and fn converges to f uniformly on compact sets, then Tfn converges in operator norm to Tf . 2.4.1. The hereditary functional calculus. Given an operator T and a polynomial p(z, w) = P pj,ℓzj(w∗)ℓ, the hereditary calculus of Agler [5] evaluates p(T, T ∗) = P pj,ℓT j(T ∗)ℓ. The calculus extend to functions f (z, w) which are analytic in z and coanalytic in w on a neighborhood of σ(T ) × σ(T )∗. Here we will not need the full power of the calculus, but we do need a generalization like that found in [4]. For integers j, let Tj denote Tζj , where ζj is defined in equation (1) (with t = 1). For an operator T ∈ B(M) with σ(T ) ⊂ A, and G ∈ B(M), the sum ∞ k(T, T ∗)(G) := TjGT ∗ j X−∞ converges absolutely. The same is also true of 1 k (T, T ∗)(G) := C′ ∞ X−∞ (−1)jTjGT ∗ j . The following Lemma follows from the functional calculus considerations in [4] together with the fact that, by hypothesis, σ(T ) × σ(T ∗) ⊂ A × A (see [22]). Lemma 2.5. Let T, G ∈ B(M) be given. If σ(T ) ⊂ A, then k(T, T ∗)( and likewise, 1 k (T, T ∗)(G)) = G, 1 k (T, T ∗)(k(T, T ∗)(G)) = G. If Gα ∈ B(M) is a (norm bounded) net which converges WOT to G ∈ B(M), k (T, T ∗)(Gα) con- then k(T, T ∗)(Gα) converges WOT to k(T, T ∗)(G); and likewise 1 verges WOT to 1 k (T, T ∗)(G) 2.5. The model operator. The operator of multiplication by z on H 2(k) gives rise to the representation M : H ∞(A) → B(H 2(k)) defined by M (f )g = Mf g = f g. (Note σ(Mζ) = A.) To simplify notation, if H is a Hilbert space, we also use M to denote the representation M ⊗ IH on H 2(k) ⊗ H. We say that M on H 2(k) ⊗ H lifts the representation T : H ∞(A) → B(M) if there is an isometry V : M → H 2(k) ⊗ H so that V T ∗ = M ∗V ; i.e., for each f ∈ H ∞, V T ∗ f V . An application of Runge's Theorem, or simply arguing with Laurent series, together with the considerations in Subsection 2.4 shows that it suffices to assume that V T ∗ f = M ∗ ζ = M ∗ ζ V . If M ⊂ H 2(k) ⊗ H is invariant for M ∗ (that is M ∗ f M ⊂ M for all f ∈ H ∞(A)), then T = V ∗M V given by Tf = P Mf P , where V is the inclusion of M into H 2(k) ⊗ H, is also a representation. Indeed, in this case M lifts T . 6 S. MCCULLOUGH AND S. SULTANIC 2.6. Agler decompositions. Suppose T ∈ B(M) is an operator with σ(T ) ⊂ A and such that T is lifted by M. Further suppose X ∈ B(M) commutes with T ; i.e., Tf X = XTf for all f ∈ H ∞(A). As in Subsection 2.5, note that it suffices to assume that Tζ X = XTζ. An Agler decomposition, for the pair (T, X) is a B(M)-valued measure µ on B(T), the Borel subsets of T (identifying Ψ with T), µ : B(T) → B(M) such that (i) for each ϕ in the scalar Schur class and each Borel set ω, (4) (5) (ii) k(T, T ∗)(µ(ω)) − Tϕk(T, T ∗)(µ(ω))T ∗ ϕ (cid:23) 0 and; 1 k (T, T ∗)(I − XX ∗) = µ(T) −Z Tψdµ(ψ)T ∗ ψ. Here, for self-adjoint operators A and B, the notation A (cid:23) B means A − B is Several remarks are in order. positive semi-definite and similarly A ≻ B means A − B is positive definite. Remark 2.6. The integral on the right hand side of item (ii) is interpreted weakly as follows. Given a measurable partition P = (ωj)n j=1 of T and points S = (sj ∈ ωj), let ∆(P, S, µ) = P Tsj µ(ωj)T ∗ sj . The tagged partitions (P, S) form an directed set ordered by refinement of partitions, and it turns out, because of (4), that the net {∆(P, S, µ) : (P, S)} converges in the WOT and its limit is the integral. Thus the integral here, and the corresponding L2 spaces that appear later, shares much with the integration theory based of Riemann sums and is not so different than others found in the literature. For a recent example, see [24]. Detail of the construction are given in Section 4. Narrowly tailoring the development to the present needs has the virtue of keeping the presentation self contained and ultimately the paper shorter. Remark 2.7. The definition of operator-valued measure requires µ to be WOT countably additive. Thus, the second part of Lemma 2.5 implies that Λ(ω) = k(T, T ∗)(µ(ω)) is also an operator-valued measure. It is not assumed that µ(T) = I. 2.7. The main result. Definition 2.8. Given T ∈ B(M) with σ(T ) ⊂ A, a lifting V T ∗ = M ∗V of T by M on H 2(k) ⊗ H is minimal if Q∗V M is dense in H. Here Q∗P fjζj = f0. In the next section it is shown that a minimal lifting is essentially unique. The following theorem is the main result of this paper. Theorem 2.9. Let M be a separable Hilbert space. Suppose X, T ∈ B(M) and (i) σ(T ) ⊂ A; (ii) M on H 2(k) ⊗ H with V T ∗ = M ∗V is a minimal lifting; and (iii) XTϕ = TϕX for each ϕ ∈ H ∞(A). The following are equivalent. (sc) There is an F ∈ S(A,H) so that XV ∗ = V ∗MF . (ad) There is an Agler decomposition µ : B(T) → B(M) for the pair (T, X). AGLER-COMMUTANT LIFTING ON AN ANNULUS 7 Remark 2.10. It is illuminating to consider the special case of Agler-Pick inter- polation on A. Let z1, . . . , zn ∈ A and w1, . . . , wn ∈ D be given. Let M ⊂ H 2(k) denote the span of {kzj} and let V denote the inclusion of M into H 2(k). Then T defined by T = V ∗M V is lifted by M and its spectrum is the set of {zj}. Define X ∗ on M by X ∗kzj = w∗ j kzj . Then X commutes with T . In this case 1 k h (T, T ∗)(I − XX ∗)kzℓ , kzji = 1 − wjw∗ ℓ and Z Tψhd µ(ψ)T ∗ ψkzℓ , kzji = Z ψ(zj)ψ(zℓ)∗hd µ(ψ)kzℓ , kzwi. Thus part (ii) in an Agler decomposition takes the form, 1 − wj w∗ ℓ = Z [1 − ψ(zj)ψ(zℓ)∗]hd µ(ψ)kzℓ , kzwi. 3. More on Liftings Recall the orthonormal basis {ζn}n∈Z (with t = 1) of equation (1) and let Tj and Mj denote Tζj and Mζj respectively, where M ∗ acting on H 2(k) ⊗ H lifts T ∗ acting on M. The following is a version of a theorem of Ambrozie, Englis, and Muller [4], a result very much in the spirit of the de Branges-Rovnyak construction [17] and related to the results of [6]. Proposition 3.1. Suppose T ∈ B(M) and σ(T ) ⊂ A. If M = M ⊗ IH lifts T with V T ∗ = M ∗V , then V h = X ζj ⊗ RT ∗ j h (6) where R = Q∗V : M → H, the operator Q : H → H 2(k) ⊗ H is defined by and the sum converges in norm. In particular, the (non-decreasing) sum Q∗X fj ⊗ ζj = f0, (7) n Xj=−n TjR∗RT ∗ j converges WOT to the identity. Conversely, if there is an R : M → H so that the sum in equation (7) converges WOT to the identity, then M lifts T via V T ∗ = M ∗V where V is given by equation (6). Moreover, for f ∈ H ∞ and h ∈ H, V ∗(f ⊗ h) = Tf R∗h. (8) Proof. Suppose V : M → H 2(k) ⊗ H is an isometry and V T ∗ f V for all f ∈ H ∞(A). Since V : M → H 2(k) ⊗ H, there exists operators Rj : M → H so that, for h ∈ M, f = M ∗ V h = X ζj ⊗ Rjh, 8 S. MCCULLOUGH AND S. SULTANIC with the sum converging SOT. Now, X ζj ⊗ RjT ∗ mh mh =V T ∗ =M ∗ mV h =X M ∗ mζj ⊗ Rjh. Taking the inner product of both sides of the above equation with 1 ⊗ e (e ∈ H) gives, hR0T ∗ mh, ei = hRmh, ei. With R = R0, this shows Rm = RT ∗ m and thus proves that V takes the form promised in equation (6). That this sum converges in norm follows from the spectral condition on T . To prove the conversely, the hypothesis that the sum converges WOT to the identity implies that V defined as in equation (6) (which converges in norm) is an isometry. We next prove equation (8), from which the conclusion that M lifts T via V T ∗ = M ∗V will follow. To start, note that, for each m ∈ Z, hV ∗ζm ⊗ e, hi =hζm ⊗ e, V hi =he, RT ∗ mhi =hTmR∗e, hi. Hence V ∗ζm ⊗ e = TmR∗e. Next note that, from the computation above, equation (8) holds for Laurent polynomials (finite linear combinations of {ζj : j ∈ Z}). Next, if f ∈ H 2(k), then there is a sequence of Laurent polynomials pn which converge to f in H 2(k) and also uniformly on compact subsets of A. Hence, pn ⊗ h converges in H 2(k) ⊗ H to f ⊗ h and also Tpn converges to Tf in norm, and equation (8) is proved. Next, if both f, g ∈ H ∞, then V ∗Mf g ⊗ e =V ∗f g1 ⊗ e =Tf gR∗e =Tf TgR∗e =Tf V ∗g ⊗ e. Thus, V ∗M = T V ∗ so that M lifts T . Proposition 3.2. Suppose T ∈ B(M) has spectrum in A. If 1 if R ∈ B(M,H) satisfies R∗R = 1 converges WOT to the identity. In particular, M lifts T . k (T, T ∗) (cid:23) 0 and k (T, T ∗), then then the sum in equation (7) (cid:3) Conversely, if G is a positive operator and the sum X TnGT ∗ n converges WOT to the identity, then G = 1 Remark 3.3. It is always possible to choose H = M or H ⊂ M , though the former choice could lead to a representation which is not minimal. k (T, T ∗). Proof. The first part of the proposition follows from I = k(T, T ∗)( 1 k (T, T ∗)(I)) = k(T, T ∗)(R∗R). AGLER-COMMUTANT LIFTING ON AN ANNULUS The hypothesis for the second part of the lemma is k(T, T ∗)(G) = I. Hence, G = 1 k (T, T ∗)(k(T, T ∗)(G)) = 1 k (T, T ∗)(I). 9 (cid:3) Recall the notion of a minimal lifting given in Definition 2.8. Proposition 3.4. The lifting V T ∗ = M ∗V of T on H 2(k) ⊗ H is minimal if and only if there does not exist a proper subspace F ⊂ H such that the range of V lies in H 2(k) ⊗ F . Proof. From the form of V , the smallest subspace F of H such that the range of V lies in H 2(k) ⊗ F is the closure of the range of R = Q∗V . Proposition 3.5. Suppose T ∈ B(M). If σ(T ) ⊂ A and 1 lifts T . k (T, T ∗) (cid:23) 0, then M If both VjT ∗ = M ∗Vj where M is acting on H 2(k) ⊗ Hj, j = 1, 2 are minimal liftings of T , then there is a unitary operator U : H1 → H2 so that (I ⊗ U )V1 = V2; i.e., a minimal lifting is unique up to unitary equivalence. (cid:3) Proof. The first part follows from Proposition 3.2. From Proposition 3.1, where Rℓ = Q∗ ℓ Vℓ : M → Hℓ and Q∗ I = k(T, T ∗)(R∗ Therefore, by Proposition 3.2 R∗ ℓ Rℓ = 1 Vℓh = X ζj ⊗ RℓT ∗ j h, ℓ P ζj ⊗ fj = f0 on H 2(k) ⊗ Hℓ. Moreover, ℓ Rℓ) = X TjR∗ k (T, T ∗) for ℓ = 1, 2. ℓ RℓT ∗ j . From minimality, Rℓ has dense range and therefore there is a unitary operator (cid:3) U : H1 → H2 so that R2 = U R1. It follows that (I ⊗ U )V1 = V2. 4. Some Functional Hilbert Spaces Theorem 2.9 involves operator-valued measures and implicitly certain related functional Hilbert spaces. In this section we sketch out the relevant constructions. Most of what is needed is summarized later as Lemma 6.1 in Section 6. 4.1. General constructions. Let B(T) denote the Borel subsets of the unit circle T. By an operator-valued measure on T we mean a Hilbert space M and a function such that ν : B(T) → B(M) (p) ν(ω) (cid:23) 0 for ω ∈ B(T); and (ca) for each e, f ∈ M, the function is a (complex) measure on B(T). ω 7→ hν(ω)e, fi A (measurable) partition P of T is a finite disjoint collection ω1, . . . , ωn ∈ B(T) whose union is T. A measurable simple function H is a function of the form H = n Xj=1 Kωj cj 10 S. MCCULLOUGH AND S. SULTANIC for some vectors cj ∈ M and partition P . Here, Kω denotes the characteristic function of a set ω. Let S denote the collection of measurable simple functions. If H ′ = The measure ν gives rise to a semi-inner product on S as follows. Pm ℓ=1 Kω′ c′ ℓ is also in S, define ℓ hH ′, Hiν = Xj,ℓ j ν(ωj ∩ ωℓ)c′ c∗ ℓ. In the usual way, this inner product gives rise to a semi-norm, kHk2 ν = hH, Hiν . A tagging S of the partition P consists of a choice of points S = (sj ∈ ωj). The pair (P, S) is a tagged partition. The collection of tagged partitions is a directed set under the relation (P, S) (cid:22) (Q, T ) if Q is a refinement of P . Given F : T → M, let F (P, S) denote the resulting measurable simple function F (P, S) = X Kωj F (sj). Thus, each such F generates the net {F (P, S) : (P, S)} of simple functions. Let R2(ν) denote those F for which the net {F (P, S)} is bounded and Cauchy in S; i.e., those F for which there is a C such that kF (P, S)kν ≤ C for all (P, S), and such that for each ǫ > 0 there is a partition Q such that for any pair (P, S), (P ′, S′) such that P and P ′ both refine Q, (9) ǫ2 > kF (P, S) − F (P ′, S′)k2 ℓ)(F (sj) − F (s′ ℓ)). (F (sj) − F (s′ ℓ))∗ν(ωj ∩ ω′ ν = Xj,k The following are some simple initial observation. Lemma 4.1. Measurable simple functions are in R2(ν). If F ∈ R2(ν) and H ∈ S, then the net hH, F (P, S)iν is Cauchy. If F, G ∈ R2(ν), then the net hF (P, S), G(P, S)iν converges. Proof. The first statement is evident. Given tagged partitions (P, S) and (Q, T ), hF (P, S), G(P, S)iν−hF (Q, T ), G(Q, T )iν ≤hF (P, S) − F (Q, T ), G(P, S)iν + hF (Q, T ), G(P, S) − G(Q, T )iν. This estimate, Cauchy-Schwarz, plus the boundedness hypothesis on the nets proves the third statement. The second statement is a special case of the third. (cid:3) Lemma 4.2. If F, G ∈ R2(ν), then so is F + G. Proof. The boundedness of the net {(F + G)(P, S)} is evident. Given tagged par- titions (P, S) and (P ′, S′), note that k(F + G)(P, S)−(F + G)(P ′, S′)kν ≤kF (P, S) − F (P ′, S′)kν + kG(P, S) − G(P ′, S′)kν . Applying this estimate to appropriate partitions and common refinement proves the result. (cid:3) AGLER-COMMUTANT LIFTING ON AN ANNULUS 11 The assignment, hF, Giν = limhF (P, S), G(P, S)iν defines a semi-inner product on R2(ν) which is also natural to write as (10) hF, Giν = Z hd ν(s)F (s), G(s). We define L2(ν) as the completion, after moding out null vectors, of R2(ν) in the (semi-)norm induced by this (semi-)inner product. Proposition 4.3. Simple functions are dense in L2(ν). In particular, the inclusion M → L2(ν) which sends m ∈ M to the equivalence class of the constant function m is bounded. Moreover, if H = Pn j, then m′ 1 Kωj mj and H ′ = Pn′ hH, H ′iν = Xj,ℓ hν(ωj ∩ ω′ 1 Kω′ j ℓ)mj, m′ ℓi. Proof. Let F ∈ R2(ν) and ǫ > 0 be given. Choose a partition Q such that for all for all tagged partitions (P, S), (P ′, S′), such that P and P ′ refine Q, the inequality (9) holds. Let H = F (Q, T ). Then, ǫ2 > k(H−F )(P, S)k2 In view of Lemma 4.1, the right hand side converges to kH − Fk2 ν and so (measur- able) simple functions are dense in R2(ν). Since R2(ν) is dense in L2(ν) the first statement follows. The second statement is a restatement of the definition of the inner product (cid:3) ν = hH, Hiν−hH, F (P, S)iν−hF (P, S), Hiν +hF (P, S), F (P, S)iν . induced by ν on measurable simple functions. While there is no reason to believe a given continuous M valued function on T should be in L2(ν), there is an important class which is. Proposition 4.4. Suppose f : T → B(M) is continuous and C is a non-negative real number. If, for each s and t and Borel set ω, both f (s)ν(ω)f (s)∗ ≤ Cν(ω) (11) and (f (s) − f (t))ν(ω)(f (s) − f (t))∗ (cid:22) kf (s) − f (t)k2ν(ω), (12) then for each m ∈ M, the function f (s)m is in L2(ν). Proof. Fix a vector m and let F (s) = f (s)∗m. The inequality of equation (11) im- plies the net {F (P, S)} is bounded. A straightforward argument using the uniform continuity of f and the inequality (12) shows that the net {F (P, S)} is Cauchy. Hence F ∈ R2(ν). (cid:3) The algebra C(T) of continuous (scalar-valued) functions on T has a natural representation on L2(ν). Lemma 4.5. If a ∈ C(T) and F ∈ R2(ν), then aF ∈ R2(ν) and moreover, kaFkν ≤ kak∞kFkν. Hence a determines a bounded linear operator τ (a) on L2(ν). The mapping sending a ∈ C(T) to τ (a) ∈ B(L2(ν)) is a unital ∗-representation. 12 S. MCCULLOUGH AND S. SULTANIC Finally, given a, a′ ∈ C(T) and simple measurable functions F = P Kωj mj and F ′ = P Kω′ m′ ℓ, ℓ ha′F ′, aFiν = Xj,ℓ Zωj ∩ω′ ℓ a′(s)a∗(s)hdν(s)m′ ℓ, mji. (13) Proof. Fix F ∈ R2(ν). For any partition P = (ωj) of T and pointing S = (sj ∈ ωj), k(aF )(P, S)k2 ν =Xhν(ωj)a(sj )F (sj), a(sj)F (sj )i =Xa(sj)2hν(ωj ∩ ω)F (sj), F (sj)i ≤kak2 ∞kF (P, S)k2 ν . Thus, since the net {F (P, S)} is bounded, so is the net {aF (P, S)}. (aF )(P, S) − (aF )(R, T ) = G + H, where If (R, T ) is another tagged partition, where R = (θℓ) and T = (tℓ ∈ θℓ), then G =Xj,ℓ H =Xj,ℓ (a(sj) − a(tℓ))Kωj ∩θℓ F (sj), a(tℓ)Kωj ∩θℓ(F (sj ) − F (tℓ)). If ǫ bounds both a(sj ) − a(tℓ) and kF (P, S) − F (R, T )kν and if C is a bound for the net {F (P, S)}, then kGkν ≤ ǫC, kHkν ≤ kak∞ǫ. Thus, using the uniform continuity of a and the fact that the net {F (P, S)} is Cauchy, it is possible to choose a partition Q of sufficiently small width so that if P and R are refinements of Q with taggings S and T respectively, then k(aF )(P, S) − (aF )(R, T )kν =kG + Hkν ≤kGkν + kHkν ≤ (C + kak∞)ǫ. Thus the net {(aF )(P, S)} is Cauchy. Hence aF ∈ R2(ν). Given a partition (P, S), It suffices to prove equation (13) in the case that F = Kωm and F ′ = Kω′m′. h(a′F ′)(P, S), (aF )(P, S)iν =X a′(sj)∗a(sj)hν(ωj ∩ ω ∩ ω′)m′mi =Zω∩ω′ X a′(sj)∗a(sj)Kωjhd ν(s)m′, mi. Given ǫ > 0, if the partition P is chosen, using the uniform continuity of a′a∗, so that k[a′a∗ −Xj a′(sj)a∗(sj )Kωj ]Kω∩ω′k∞ < ǫ, then Zω∩ω′ [a′a∗ −X(a′(sj))∗a(sj)Kωj ]hd ν(s)m′, mi ≤ ǫkν(ω ∩ ω′)k km′k kmk. It follows that the net h(a′F ′)(P, S), (aF )(P, S)iν converges to the integral Zω∩ω′ a′(s)a∗(s)hd ν(s)m′, mi, completing the proof of equation (13). AGLER-COMMUTANT LIFTING ON AN ANNULUS 13 Each a determines a bounded operator on R2(ν) (with norm at most kak∞) and hence extends to a bounded operator τ (a) on all of L2(ν). It remains to prove that τ determines a unital ∗-representation on L2(ν). Evidently τ (1) = I. Using equation (13) twice (first with a = 1 and the second with a = (a′)∗ and a′ = 1), hτ (a′)∗F, F ′iν =hF, τ (a′)Fiν =hF, a′F ′iν =Z (a′)∗(s)hdµ(s)m, m′i =hτ ((a′)∗)F, F ′iν . Hence τ (a)∗ = τ (a∗). Finally, again using equation (13) twice, this time first with a = aa′ a′ = 1, and second with a = a and a′ = (a′)∗, hτ (aa′)F, F ′iν =Z a′a(s)hdµ(s)m, m′iν =Z ((a′)∗)∗a(s)hdµ(s)m, m′i =hτ (a′)∗τ (a)F, F ′iν =hτ (a′)τ (a)F, F ′iν . Thus τ (a′a) = τ (a′)τ (a). (cid:3) 4.2. Agler decompositions again. Suppose µ is an Agler decomposition as de- fined in subsection 2.6. Then both µ, and Λ defined by Λ(ω) = k(T, T ∗)(µ(ω)), are positive B(M)-valued measures on B(T) and the constructions of the previous section apply to both L2(µ) and L2(Λ). Lemma 4.6. If F ∈ R2(Λ), then F ∈ R2(µ) and hF, FiΛ ≥ hF, Fiµ. Thus, the mapping Φ∗ : R2(Λ) → R2(µ) given by F 7→ F induces a contractive linear mapping Φ∗ : L2(Λ) → L2(µ). Proof. This follows immediately from Λ(ω) = k(T, T ∗)(µ(ω)) (cid:23) µ(ω). (cid:3) Given m ∈ M, let Y denote the mapping Y : M → L2(Λ) defined by Y m(ψ) = T ∗ ψm. Here the identification of Ψ, the collection of test functions, with T is in force. Of course, it needs to be verified that Y m(ψ) is indeed in L2(Λ). Let ι denote the inclusion, as constant functions, of M into R2(Λ). Thus, if m ∈ M, then ιm denotes the constant function ιm(ψ) = m. Lemma 4.7. For m ∈ M, the function Y m is in R2(Λ). Moreover, hΛ(T)m, miM = hιm, ιmiL2(Λ) ≥ hY m, Y miL2(Λ). 14 S. MCCULLOUGH AND S. SULTANIC Thus, Y determines a bounded linear operator Y : M → L2(Λ) given by (Y m)(ψ) = T ∗ ψm. In the notation of equation (10), (14) hY ∗Y m, miΛ = hZ d Λ(ψ)T ∗ ψm, T ∗ ψmi. Further, Φ∗Y : M → L2(µ) is bounded and (15) hY ∗ΦΦ∗Y m, m′iν = Z hd µ(ψ)T ∗ ψm, T ∗ ψm′i. Remark 4.8. We interpret equations (14) and (15) as (16) and (17) Y ∗Y = Z Tψd Λ(ψ)T ∗ ψ Y ∗ΦΦ∗Y = Z Tψd µ(ψ)T ∗ ψ respectively. Given a tagged partition (P, S), let ∆(P, S, Λ) = X Tsj Λ(ωj)T ∗ sj and define ∆(P, S, µ) similarly. Thus, ∆(P, S, Λ) is an operator on M and because TsΛ(ω)T ∗ s ≤ Λ(ω), it is positive semidefinite and bounded above by Λ(T). For vectors m, m′ ∈ M, h∆(P, S, Λ)m, m′i = hY m(P, S), Y m′(P, S)iΛ. Thus, the net {∆(P, S, Λ)} converges WOT to the operator of equation (16). It follows that the net { 1 1 k k (T, T ∗)(∆(P, S, Λ))} also converges. On the other hand, (T, T ∗)(∆(P, S, Λ)) = ∆(P, S, µ). Hence the net {∆(P, S, µ)} converges WOT to the operator of equation (17). Proof. By hypothesis, for ϕ in the scalar Schur class and measurable sets ω, (18) Λ(ω) (cid:23) TϕΛ(ω)T ∗ ϕ. Thus, the functions Y m satisfies the hypotheses, with respect to Λ, of Proposition 4.4. It follows that Y m is in L2(Λ) for each m. The moreover follows immediately from equation (18). The rest of the Lemma follows from the definitions. (cid:3) Lemma 4.9. Let µ be an Agler decomposition of the pair (T, X) and let, as in Proposition 3.1, R∗R = 1 k (T, T ∗). Then, R∗R + Y ∗ΦΦ∗Y = XR∗RX ∗ + ι∗ΦΦ∗ι. Proof. Part (ii) of the definition of an Agler decomposition can be written as 1 k (T, T ∗)(I) +ZΨ Tψdµ(ψ)T ∗ ψ = 1 k (T, T ∗)(XX ∗) + µ(T). Because X commutes with T ∗, 1 k (T, T ∗)(XX ∗) = X 1 k (T, T ∗)(I)X ∗ AGLER-COMMUTANT LIFTING ON AN ANNULUS 15 and hence 1 4.7 gives k (T, T ∗)(XX ∗) = XR∗RX ∗. An application of the last part of Lemma R∗R + Y ∗ΦΦ∗Y = XR∗RX ∗ + µ(T). Noting that hι∗ΦΦ∗ιm, m′i = hm, m′iL2(µ) = hµ(T)m, m′i completes the proof. (cid:3) 5. Uniformity of the Test Functions Using the orthonormal basis {ζj} for H 2(k) defined in equation (1), each test function ψ has a Laurent expansion, In this section we show that ψ = Xhψ, ζjiζj . with convergence in the strong operator topology. T ∗ ψ = Xhζj, ψiT ∗ j The section begins with establishing a uniform, independent of ψ, estimate on the rate of convergence of the Laurent series for ψ on compact subsets of A. Lemma 5.1. There is a 0 < ρ < 1 and a constant C so that for all ψ ∈ Ψ and j ∈ Z, hψ, ζji < Cρj. Sktech of proof. There is a function ϕ analytic in a neighborhood of our annulus A such that (a) for z = 1, ϕ(z) = 1; (b) for z = q, ϕ(z) = √q; and (c) ϕ(√q) = 0. 2} (see Section 11). It extends by reflection across both boundaries to be analytic in the annulus {q z < q− 1 It follows that, up to a unimodular constant, if ψ is unimodular on the boundary of A and has exactly two zeros, these being at √q and √qγ (for a necessarily unimodular γ), then 2 < 3 (19) ψ(z) = δ ϕ(z)ϕ(γ∗z) z , for some unimodular δ. In particular equation (19) gives an explicit parametrization of Ψ by T. It now follows that ψ ∈ Ψ is bounded uniformly (independent of ψ) on a larger annulus than A and the result follows. (cid:3) In the following Lemma µ is an Agler decomposition for (T, X). Thus, Λ(ω) = k(T, T ∗)(µ(ω)) and for ϕ in the scalar Schur class, TϕΛ(ω)T ∗ Lemma 5.2. If m ∈ M, then, for each j, the function hζj , ψiT ∗ moreover, independent of j, there is a C > 0 and 0 < ρ < 1 such that ϕ (cid:22) Λ(ω). j m ∈ L2(Λ) and khζj, ψiT ∗ j mkL2(Λ) ≤ Cρj. 16 S. MCCULLOUGH AND S. SULTANIC If G ∈ L2(Λ) and m ∈ M, then (20) hG, Y miL2(Λ) = Xj hG,hζj , ψiT ∗ j miL2(Λ). If F is a measurable simple function, then hF, Φ∗Y miL2(µ) = Xj hF,hζj , ψiT ∗ j miL2(µ). Proof. Given a positive integer N , define σN : T → H ∞(A) by σN (ψ) = Xj≤N hζj, ψiζj . In view of Lemma 5.1, the sequence σN converges to the identity function ψ uni- formly on compact subsets of A. Hence, by Proposition 4.4, for each m ∈ M kT ∗ ψ−σN mkL2(Λ) = kT ∗ ψm − Xj≤N hζj , ψiT ∗ j mkL2(Λ) converges to 0 and equation (20) follows. To finish the proof, choose G = ΦF in equation (20) to obtain hΦF, Y miL2(Λ) =Xj =Xj hΦF,hζj , ψiT ∗ j miL2(Λ) hF,hζj , ψiT ∗ j miL2(µ). using that Φ∗ : L2(Λ) → L2(µ) is the inclusion mapping (and is bounded). The final conclusion of the lemma follows. (cid:3) 6. The Factorization and Lurking Isometry The next several sections, Sections 6, 7, and 8, are devoted to the proof of (ad) implies (sc) in Theorem 2.9 and throughout these sections the relevant hypotheses are in force. Namely, M is a separable Hilbert space, (a) X, T ∈ B(M) commute; (b) σ(T ) ⊂ A; (c) T lifts to M on H 2(k) ⊗ M via V T ∗ = M ∗V and V h = X ζj ⊗ RT ∗ j h, where R∗R = 1 k (T, T ∗); and (d) there exists a measure µ : B(T) → B(M) such that, with Λ(ω) = k(T, T ∗)(µ(ω)), for all Borel subset ω and Schur class functions ϕ and Λ(ω)) − TϕΛ(ω)T ∗ ϕ (cid:23) 0 1 k (T, T ∗)(I − XX ∗) = µ(T) −Z Tψdµ(ψ)T ∗ ψ. Once properly formulated to account for infinitely many test functions, the over- arching strategy for proving results like Theorem 2.9 is now well established, but the presence of infinitely many, and not necessarily orthogonal, test functions requires some reinterpretation of earlier results, revealing new structures. The positivity condition in (ad) (item (d) above) is factored and this factorization produces a AGLER-COMMUTANT LIFTING ON AN ANNULUS 17 lurking isometry and of course an auxiliary Hilbert space. The lurking isometry in turn generates the Ψ-unitary colligation. A good deal of effort is required to show that the resulting transfer function solves the problem and the argument given here is patterned after that in [29], which in turn borrowed from [14] [11] and closely related to those in [3]. The factorization we will need comes from factoring the measure Λ of Remark 2.7. This factorization amounts to the construction of the Hilbert spaces L2(Λ) and L2(µ) in Section 4. The following Lemma summarizes many of the needed results and constructions from Section 4 Lemma 6.1. With the hypotheses above, (i) there exist Hilbert spaces L2(Λ) and L2(µ) which contain densely all simple measurable M-valued functions so that, in particular, the inclusion mapping ι : M → L2(Λ) is bounded (not necessarily isometric); (ii) the space L2(Λ) includes in L2(µ) contractively so that there exists an op- erator Φ whose adjoint Φ∗ : L2(Λ) → L2(µ) is the inclusion mapping; and ψm (that is, the function Y m(ψ) = T ∗ (iii) an operator Y : M → L2(Λ) defined by Y m = T ∗ ψm determines an element of L2(Λ)), which together satisfy the lurking isometry equality, (21) R∗R + Y ∗ΦΦ∗Y = XR∗RX ∗ + ι∗ΦΦ∗ι. Moreover, if (a) a, a′ : T → C are continuous; (b) ω, ω′ are Borel subsets of T; (c) m, m′ ∈ M; and (d) F = Kωm and F ′ = Kω′m′, then aF and a′F ′ are in L2(µ) and haF, a′F ′i = Zω∩ω′ a(ψ)a′(ψ)∗hdµ(ψ)m, m′i. In particular, if F ∈ L2(µ) is simple, then aF determines an element of L2(µ) and kaFk ≤ kak∞kFk. Thus, there is a unital ∗-representation τ : C(T) → B(L2(µ)) such that τ (E(z))F (ψ) = ψ(z)F (ψ). (Recall the identification of Ψ, the collection of test functions, with T.) Condition (i) in the definition of Agler decomposition implies Y is a bounded (in fact contractive) operator into L2(Λ). (Details in Section 4). Further, Φ∗Y m = T ∗ ψm determines an element of L2(µ) and in condition (ii) in the definition of an Agler decomposition equation (5) becomes, Thus hΦ∗Y m, Φ∗Y mi = Z hTψdµ(ψ)T ∗ ψm, mi. 1 k (T, T ∗) − X 1 k (T, T ∗)X ∗ = ι∗ΦΦ∗ι − Y ∗ΦΦ∗Y. Rearranging and using the relation 1 the lurking isometry equality of equation (21). k (T, T ∗) = R∗R of Proposition 3.2 produces 18 S. MCCULLOUGH AND S. SULTANIC 7. The colligation and its Transfer Function Recall there are two parts to the colligation. The unitary matrix and the repre- sentation. 7.1. The unitary matrix. The lurking isometry, equation (21), produces, non- uniquely, the unitary matrix of item (iii) of Definition 2.1. The construction requires an initial enlargement of the space L2(µ). Let ℓ2 denote the usual separable Hilbert space with orthonormal basis {ej : j ∈ N} and define W : L2(µ) → L2(µ) ⊗ ℓ2 by WF = F ⊗ e0. In particular, W is an isometry. Let K and K∗ denote the subspaces of [L2(µ) ⊗ ℓ2] ⊕ H given by the closures of the spans of {(cid:0)WΦ∗Y m ⊕ Rm(cid:1) : m ∈ M}, {(cid:0)WΦ∗ιm ⊕ RX ∗m(cid:1) : m ∈ M} respectively, where ι is the inclusion of M into L2(Λ). The lurking isometry of equation (21) says that the mapping from K to K∗ defined by (cid:0)WΦ∗Y m ⊕ Rm(cid:1) → (cid:0)WΦ∗ιm ⊕ RX ∗m(cid:1) is an isometry. Because K and K∗ have the same codimension (i.e., their orthogonal complements have the same dimension), this isometry can be extended to a unitary U = (cid:18)A∗ C∗ B∗ D∗(cid:19) : ⊕ H giving rise to the usual system of equations, L2(µ) ⊗ ℓ2 → L2(µ) ⊗ ℓ2 ⊕ H , (22) A∗WΦ∗Y + C∗R =WΦ∗ι B∗WΦ∗Y + D∗R =RX ∗. Note that the domain of D and B and the codomain of C is M. 7.2. The representation. Of course we also need the representation ρ : C(T) → B(L2(µ)⊗ ℓ2) of item (ii) in Definition 2.1. We begin with the unital representation τ : C(T) → B(L2(µ)) from Lemma 6.1 (see also Lemma 4.5) and define ρ = τ ⊗ I, where I is the identity on ℓ2. 7.3. The transfer function and its properties. Let E(z) : Ψ → C denote evaluation at z; i.e., E(z)(ψ) = ψ(z). For F ∈ L2(µ), τ (E(z))F (ψ) = ψ(z)F (ψ). The corresponding transfer function is then given by W (z) = D + C(I − ρ(E(z))A)−1ρ(E(z))B. (23) The function W gives rise to the multiplication operator MW on H 2(k) ⊗ M. In the following subsection we make some observations related to MW and the corresponding Ψ-unitary colligation needed in the sequel. There is a canonical auxiliary multiplication operator associated to z 7→ ρ(E(z)) which, as in equation (3), is most conveniently defined in terms of its adjoint. Define Z ∗ : H 2(k) ⊗ [L2(µ) ⊗ ℓ2] → H 2(k) ⊗ [L2(µ) ⊗ ℓ2] by Z ∗(kz ⊗ F ) = kz ⊗ ρ(E(z))∗F. AGLER-COMMUTANT LIFTING ON AN ANNULUS 19 Of course it needs to be checked that, after extending by linearity, this prescription produces a bounded operator, a fact that follows readily from hkz ⊗X Fj ⊗ ej,kw ⊗X Gℓ ⊗ eℓi −Xj hZ ∗kz ⊗ Fj, Z ∗kw ⊗ Gji =Xj =Xj k(w, z)[hF, Gi − hρ(E(z))∗Fj , ρ(E(w))∗Gji Z k(w, z)(1 − ψ(z)∗ψ(w))G(ψ)∗dµ(ψ)F (ψ) and the fact that each k(z, w)(1− ψ(z)ψ(w)∗) is a positive kernel and µ is a positive measure. Here we have used Proposition 2.4 and have actually proved that Z has norm at most one. Thus ρ(E(z)) determines a (multiplication) operator on H 2(k) ⊗ [L2(µ) ⊗ ℓ2] denoted by Z: hZ f, kz ⊗ FiH 2(k)⊗[L2(µ)⊗ℓ2] = hρ(E(z))f(z), FiL2(µ)⊗ℓ2 . Lemma 7.1. Given a simple measurable function F = P Kωℓmℓ ∈ L2(µ) and f ∈ H ∞, (24) Z(f ⊗ (F ⊗ ep)) = X f ζj ⊗ hψ, ζjiF ⊗ ep. Here Kωℓ is the characteristic function of the Borel set ω ⊂ T; ep is the element of ℓ2 with a 1 in the p-th entry and 0 elsewhere; and the symbol ψ denotes the variable in Ψ. In particular, the sum on the right hand side converges. Since hψ, ζji is con- tinuous, it follows, from the moreover part of Lemma 6.1 that hψ, ζjiKωℓ mℓ is in L2(µ). In Section 11 we show that there is a 0 < ρ < 1 and a C such that for all j, hψ, ζji < Cρj (see also Lemma 5.1). Note also, ψ(z) = Xj hψ, ζjiζj (z). Proof. Choose C and ρ as above. It follows from Lemma 4.5 that khψ, ζjiFkL2(µ) ≤ CρjkFk and thus the sum on the right hand side of equation (24) converges. Because simple functions are dense in L2(µ) by item (i) of Lemma 6.1, it suffices to prove the result assuming F = Kωm ⊗ ep, for a Borel set ω. Given z ∈ A and a 20 S. MCCULLOUGH AND S. SULTANIC (very) simple function F ′ = Kω′m′ ⊗ ep hZ(f ⊗ F ⊗ ep), kz ⊗ F ′ ⊗ epi =h(f ⊗ F ⊗ ep), kz ⊗ ρ(E(z))∗(F ′ ⊗ ep)iH 2(k)⊗[L2(µ)⊗ℓ2] f (z)ψ(z)hdµ(ψ)m, m′i f (z)ζj(z)Zω∩ω′hψ, ζjihdµ(ψ)m, m′i f (z)ζj(z)hhψ, ζjiF, F ′i =h(f ⊗ F ), kz ⊗ τ (E(z))∗F ′iH 2(k)⊗L2(µ) =Zω∩ω′ =Xj =Xj =Xj hf ζj ⊗ hhψ, ζjiF ⊗ ep, kz ⊗ F ′ ⊗ epi. Returning to the transfer function W of equation (23), let W = W (z)−D. Before concluding this subsection, we present two key relations amongst V, W, R, Φ, ι and Z. Define J : M → H 2(k) ⊗ M by (cid:3) J m = X ζj ⊗ T ∗ j m. The spectral condition σ(T ) ⊂ A implies this sum converges and J is a bounded operator. Note that (I ⊗ R)J = V . Lemma 7.2. For f ∈ H ∞(A) and F ∈ L2(µ) ⊗ ℓ2, (25) J ∗(I ⊗ ι∗ΦW ∗)Zf ⊗ F = Tf Y ∗ΦW ∗F, and J ∗(I ⊗ ι∗ΦW ∗)f ⊗ F = Tf ι∗ΦW ∗F. Proof. First, suppose f = ζ p for some integer p. Straightforward computation and the fact that M lifts T gives, hζ pζℓ, ζp+ℓiTp+ℓ = TℓT p. Given m ∈ M, ω ∈ B(T) and h ∈ L2(µ) ⊗ ℓ2, let F = Kω ⊗ h, and compute, hJ ∗(I ⊗ ι∗ΦW ∗)Z(zp ⊗ F ), miM zpζℓ ⊗ hψ, ζℓiW ∗F ], ζj ⊗ Φ∗ιT ∗ j miH 2(k)⊗L2(µ) j miH 2(k)⊗M j miH 2(k)⊗L2(µ) p+ℓmiL2(µ) =h(I ⊗ ι∗ΦW ∗)Z(zp ⊗ F ),X ζj ⊗ T ∗ =Xj hI ⊗ W ∗Z(zp ⊗ F ), ζj ⊗ Φ∗ιT ∗ =Xℓ h[Xℓ =Xℓ hζ pζℓ, ζp+ℓihψ, ζℓihW ∗F, Φ∗ιT ∗ =Xℓ ℓ (T ∗)pmiL2(Λ) hΦW ∗F,hζℓ, ψiT ∗ =hΦW ∗F, Y (T ∗)pmiL2(Λ) =hT pY ∗ΦW ∗F, miM. AGLER-COMMUTANT LIFTING ON AN ANNULUS 21 Here we have used the form of V from Proposition 3.1 in the second equality; the description of Z provided by Lemma 7.1 in the fourth; and Lemma 5.2, equation (20) in the seventh. Now use linearity and the fact that the linear span of elements like F is dense in L2(µ) ⊗ ℓ2 to finish the proof of the first part of Lemma 7.2. second. An argument very much like the one that proved the first identity proves the (cid:3) We now use Lemma 7.2 to establish the following Lemma. Lemma 7.3. With notations as above (and A, B and C appearing in the repre- sentation of the transfer function W ), J ∗(I ⊗ ι∗ΦW ∗)[I − Z(I ⊗ A)] = V ∗(I ⊗ C). Proof. For f ∈ H ∞(A) and F ∈ L2(µ) ⊗ ep, J ∗(I ⊗ ι∗ΦW ∗)[I − Z(I ⊗ A)](f ⊗ F ) =J ∗(I ⊗ ι∗ΦW ∗)[f ⊗ F − Z(f ⊗ AF )] =Tf [ι∗ΦW ∗ − Y ∗ΦW ∗A]F =Tf [R∗C]F =V ∗[f ⊗ CF ]. Here both parts of Lemma 7.2 were used in the second equality, equation (22) (i) was used in the third, and Proposition 3.1 in the last. Since the linear span of elements of the form f⊗F is dense in H 2(k)⊗[L2(µ)⊗ℓ2], (cid:3) the result follows. The following Lemma does the heavy lifting in the proof of (ad) implies (sc) in Theorem 2.9. Recall W = W − D. Lemma 7.4. For m ∈ M, J ∗(I ⊗ ι∗ΦW ∗)Z(I ⊗ B)(1 ⊗ m) = V ∗MW(1 ⊗ m). Proof. Choose a sequence 0 < tn < 1 converging to 1 and let Zn = (1 − tn)[I − tnZ(I ⊗ A)]−1. We claim that Zn converges to 0 in the WOT. The first step in proving this claim is to show that Zn is contractive which follows from the following computation in which we have written S in place of Z(I ⊗ A): I − ZnZ ∗ n =(I − tnS)−1[(I − tnS)(I − tnS)∗ − (1 − tn)2](I − tnS)−∗ =tn(I − tnS)−1[−S − S∗ − tn(1 − SS∗)](I − tnS)−∗ =tn(I − tnS)−1[(I − S)(I − S∗) + (1 − tn)(1 − SS∗)](I − tnS)−∗. Noting that S is a contraction - since both Z and A are contractions - it follows that Zn is a contraction. 22 S. MCCULLOUGH AND S. SULTANIC Next observe that, for given f ∈ H 2(k) ⊗ L2(µ) ⊗ ℓ2, z ∈ A and F ∈ L2(µ) ⊗ ℓ2, hf, (tn(I ⊗ A)∗Z ∗)jkz ⊗ Fi hZnf, kz ⊗ Fi =(1 − tn)Xj =(1 − tn)Xj hf, kz ⊗ (tnA∗ρ(E(z))∗)jFi =(1 − tn)hf, kz ⊗ (I − tnA∗ρ(E(z))∗)−1Fi which evidently tends to 0 as tn tends to 1, since kρ(E(z))Ak < 1. The statement about WOT convergence now follows. Let Wn = C(I − tnρ(E(z))A)−1ρ(E(z))B. Because Wn converges pointwise Next, for m, h ∈ M, boundedly to W, MWn converges WOT boundedly to MW. h(I ⊗ C)(I − tnZ(I ⊗ A))−1Z(I ⊗ B)(1 ⊗ m), kz ⊗ hi =h(I − tnZ(I ⊗ A))−1Z1 ⊗ Bm, kz ⊗ C∗hi =h(I − tnρ(E(z))A)−1ρ(E(z))Bm, C∗hi =hWn(z)m, hi =h(MWn 1 ⊗ m)(z), hi =hMWn 1 ⊗ m, kz ⊗ hi. Hence (I ⊗ C)(I − tnZ(I ⊗ A))−1Z(I ⊗ B)(1 ⊗ m) = MWn m. We are now in a position to complete the proof. Using Lemma 7.3, V ∗MWn (1 ⊗ m) =V ∗(I ⊗ C)(I − tnZ(I ⊗ A))−1Z(I ⊗ B)(1 ⊗ m) =J ∗(I ⊗ ι∗ΦW ∗)[I − Z(I ⊗ A)]× (I − tnZ(I ⊗ A))−1Z(I ⊗ B)(1 ⊗ m) =J ∗(I ⊗ ι∗ΦW ∗)Z(1 ⊗ Bm) + (tn − 1)V ∗(I − tnZ(I ⊗ A))−1Z(I ⊗ B)(1 ⊗ m). As n tends to infinity, the left hand side tends to V ∗MW (WOT) and the second term on the right hand side tends to 0 (WOT) completing the proof. (cid:3) AGLER-COMMUTANT LIFTING ON AN ANNULUS 23 8. Proof of (ad) Implies (sc) Using the ingredients assembled in the previous section, the proof that (ad) since Mf , MW commute since V ∗ intertwines Mf and Tf implies (sc) follows readily. For f ∈ H ∞(A) and m ∈ M, V ∗MW (f ⊗ m) =V ∗Mf MW (1 ⊗ m) =Tf V ∗MW (1 ⊗ m) =Tf V ∗[(1 ⊗ Dm) + MW(1 ⊗ m)] =Tf V ∗(1 ⊗ Dm) + Tf J ∗(I ⊗ ι∗ΦW ∗)Z1 ⊗ Bm from Lemma 7.4 =Tf R∗Dm + Tf J ∗(I ⊗ ι∗ΦW ∗)Z1 ⊗ Bm from equation (8) =Tf R∗Dm + Tf ι∗Y ∗ΦW ∗Bm using equation (25) =Tf [R∗D + Y ∗ΦW ∗B]m =Tf XR∗m using the second equation in (22) =XTf R∗m =XV ∗(f ⊗ m) from Proposition 3.1, equation (8). 9. The Converse This section is devoted to the proof of the implication (sc) implies (ad) of Theo- rem 2.9. Accordingly assume hypotheses (i), (ii), and (iii) and also the representa- tion (sc) for X in Theorem 2.9 throughout this section. Thus there is an W with a Ψ-unitary colligation transfer function representation W (z) = D + C(I − ρ(E(z))A)−1ρ(E(z))B W V . For definiteness, write such that V X ∗ = M ∗ U = (cid:18)A B For technical reasons, let, for 0 ≤ r < 1, C D(cid:19) : E ⊕ H → . E ⊕ H Wr(z) = D + C(I − rρ(E(z))A)−1rρ(E(z))B. Like before, let The usual computation reveals, Hr(z) = C(I − rρ(E(z))A)−1. (26) I − Wr(z)Wr(w)∗ = Hr(z)(I − r2ρ(E(z))ρ(E(w))∗)Hr(w)∗. There is a spectral measure E associated with the representation ρ. Thus E : B(T) → B(E) and, in particular, ρ(E(z))ρ(E(w))∗ = ZΨ E(z)E(w)∗dE(ψ) = ZΨ where E(z)(ψ) = ψ(z) has been used. Lemma 9.1. There exists a constant κ > 0 so that H(z)H(w)∗ (cid:22) κk(z, w). (Here the inequality is in the sense of kernels). ψ(z)ψ(w)∗dE(ψ), 24 S. MCCULLOUGH AND S. SULTANIC Proof. Multiplying equation (26) by k(z, w) gives, (27) (I − Wr(z)Wr(w)∗)k(z, w) = Hr(z)[Z k(z, w)(1 − r2ψ(z)ψ(w)∗)dE(ψ) ]Hr(w)∗ On the other hand, with b = √q, since ψ(b) = 0 we have k(z, b)(1 − r2ψ(z)ψ(b)∗) = k(z, b). Thus, (28) k(z, w)(1 − r2ψ(z)ψ(w)∗) (cid:23) k(z, b)k(b, w) k(b, b) . Letting G(z) = k(z,b)√k(b,b) , combining equations (28) and (27), using and E(T) = I, gives, k(z, w) (cid:23) k(z, w)(I − Wr(z)Wr(w)∗) (29) The function g(z) = 1 k(z, w) (cid:23) Hr(z)G(z)G(w)∗Hr(w)∗. G is analytic in a neighborhood of the annulus and is thus a multiplier of H 2(k). In particular, there is an η so that k(z, w)(η2− g(z)g(w)∗) (cid:23) 0. This last inequality is more conveniently written as (30) η2k(z, w) (cid:23) k(z, w)g(z)g(w)∗. Putting equations (29) and (30) together yields, η2k(z, w) (cid:23) k(z, w)g(z)g(w)∗ (cid:23) Hr(z)Hr(w)∗ (cid:3) From Lemma 9.1 it follows that Hr induces a bounded linear operator H∗ r : H 2(k) ⊗ H → E of norm at most √κ, determined by rkz ⊗ e = Hr(z)∗e. H∗ Hence for 0 < r ≤ 1, the formula Qr(ω) = Hr E(ω)H∗ r, for a Borel subset ω of T, defines a B(H2(k) ⊗ H)-valued measure. k(T, T ∗)(µr(ω)) and Λ = Λ1. Let Q = Q1. Define µr(ω) = V ∗Qr(ω)V and let µ = µ1. Finally, let Λr(ω) = Lemma 9.2. For fixed ω, the operators Qr(ω) are uniformly bounded by κ, and, for each ω, the net Qr(ω) converges WOT to Q(ω). Similarly, µ(ω) = V ∗Q(ω)V defines a B(M)-valued measure on T and µr(ω) = V ∗Qr(ω)V converges WOT boundedly to µ(ω). Finally, Λr(T) is uniformly bounded and Λr(ω) converges WOT to Λ(ω) for each Borel set ω. AGLER-COMMUTANT LIFTING ON AN ANNULUS 25 Proof. The uniform bound on the Qr follows immediately from the fact that Hr is uniformly bounded. Next, as r tends to 1, hQr(ω)kz ⊗ e, kw ⊗ fi = hE(ω)Hr(z)∗e, Hr(w)∗fi → hE(ω)H1(z)∗e, H1(w)∗fi. Here we have used, for a fixed z ∈ A, Hr(z) converges in norm to H1(z). Since Qr(ω) is uniformly bounded and converges WOT to Q(ω) against a dense set of vectors, it converges WOT to Q(ω). The spectral condition on T and the fact that µr(T) is uniformly bounded implies Λr(T) is also uniformly bounded. Since µr(ω) converges WOT to µ(ω) it follows that k(T, T ∗)(µr(ω)) converges WOT to k(T, T ∗)(µ(ω)). (cid:3) To complete the proof of Theorem 2.9 it remains to show that µ produces an Agler decomposition for the pair (T, X). Lemma 9.3. For 0 < r < 1 and each Borel set ω ⊂ T, Λr(ω) = V ∗MHr (I ⊗ E(ω))M ∗ Hr V. For any ϕ ∈ H ∞(A) of norm at most one and Borel set ω, Λ(ω) − TϕΛ(ω)T ∗ ϕ (cid:23) 0. Using the definition of Λ, the conclusion of the second part of the lemma is k(T, T ∗)(µ(ω) − Tϕµ(ω)T ∗ From the definition of µ and the lifting property V T ∗ ϕ) (cid:23) 0. µ(ω) − Tϕµ(ω)T ∗ ψ = M ∗ ψV , ϕ = V ∗(Q(ω) − MϕQ(ω)M ∗ ϕ)V. Proof. Let so that for an operator G, kn(z, w) = n X−n ζj (z)ζj(w)∗ kn(M, M ∗)(G) = n X−n MjGM ∗ j , where Mj = Mζj . To prove that kn(M, M ∗)(Qr(ω)) converges WOT to MHr (I ⊗ E(ω))M ∗ Hr , observe, for a Borel set ω, that (31) hkn(M, M ∗)(Qr(ω))kw ⊗ e, kz ⊗ fi =kn(z, w)hE(ω)Hr(w)∗e, H ∗ (cid:22)k(z, w)hE(ω)Hr(w)∗e, H ∗ =hMHr (I ⊗ E(ω))M ∗ (cid:22)hMHr M ∗ kw ⊗ e, kz ⊗ fi, Hr r (z)fi r (z)fi Hr kw ⊗ e, kz ⊗ fi where the inequalities are in the sense of (positive semidefinite) kernels. Since also kn(M, M ∗)(Qr(ω)) is a bounded increasing sequence of positive op- erators equation (31) implies that kn(M, M ∗)(Qr(ω)) converges WOT to MHr (I ⊗ E(ω))M ∗ Hr E(ω))M ∗ Hr . Hence, V ∗kn(M, M ∗)(Qr(ω))V = kn(T, T ∗)(µr(ω)) converges to V ∗MHr (I⊗ V, proving the first part of the Lemma. 26 S. MCCULLOUGH AND S. SULTANIC Similarly, kn(M, M ∗)(MϕQr(ω)M ∗ to MϕMHr (I ⊗ E(ω))M ∗ nition of Qr, Hr ϕ) = Mϕkn(M, M ∗)(Qr(ω))M ∗ ϕ, converges WOT ϕ. Thus, letting n tend to infinity and using the defi- M ∗ hkn(M, M ∗)(Qr(ω)−MϕQr(ω)M ∗ ϕ)kw ⊗ e, kz ⊗ fi →(1 − ϕ(z)ϕ(w)∗)k(z, w)hHr(z)E(ω)Hr(w)∗e, fi The kernel on the right hand side is positive semi-definite because it is the pointwise product of positive semi-definite kernels, and thus lim WOT kn(M, M ∗)[ Qr(ω) − MϕQr(ω)M ∗ ϕ ] (cid:23) 0. Thus, WOT 0 (cid:22)V ∗ lim kn(M, M ∗)(Qr(ω) − MϕQr(ω)M ∗ =k(T, T ∗)[ V ∗Qr(ω)V − TϕV ∗Qr(ω)V T ∗ ϕ ]. ϕ)V Finally, letting r tend to 1 on the right hand side above and applying Lemmas 9.2 and 2.5 gives 0 (cid:22) k(T, T ∗)[µ(ω) − Tϕµ(ω)T ∗ ϕ]. (cid:3) It remains to verify the condition of equation (5). The argument is an elaboration on the proof of the preceding lemma, making use of the approximations Hr and the related operators Hr and MHr . We break the proof into several steps as outlined in the Lemma below. Lemma 9.4. With notations as above: (i) For 0 < r < 1 and each Borel set ω ⊂ T, Λr(ω) = V ∗MHr (I ⊗ E(ω))M ∗ Hr V and converges WOT to Λ(ω). particular, kM ∗ (ii) There is a constant C∗ such that kΛr(T)k ≤ C2 (iii) There is a bounded operator Γ on H 2(k) ⊗ E determined by hΓkw ⊗ f, kz ⊗ gi = k(z, w)Z ψ(z)ψ(w)∗dhE(ψ)f, gi. V k ≤ C∗ independent of r. Hr ∗ for all 0 < r < 1. In (iv) If (ωj) is a Borel partition of T of diameter at most ǫ > 0, then for any choice of points sj ∈ ωj, ǫ > kΓ −X Msj M ∗ sj ⊗ E(ωj)k. (v) The identity k(T, T ∗)(Z Tψd µ(ψ)T ∗ ψ) = Z Tψd Λ(ψ)T ∗ ψ holds. Note, in item (iv) the identification of T with Ψ is in force so that s − t < ǫ means kMs − Mtk = ks − tk∞ < ǫ. AGLER-COMMUTANT LIFTING ON AN ANNULUS 27 Proof of Lemma 9.4. The first part of item (i) is part of Lemma 9.2. The descrip- tion of Λr in terms of MHr and E is the first part of Lemma 9.3. To prove item (ii), note that Λr(T) is uniformly bounded by Lemma 9.2, so Hr follows from this bound on Λr(T) and the there is a C∗. The bound on M ∗ representation of Λr in item (i). k(z, w). Item (iii) is a consequence of the fact that, as kernels, k(z, w)ψ(z)ψ(w)∗ (cid:22) To prove item (iv), first note if (ωj)j is a partition of T, then X Γ(ωj) = Γ. s − MtM ∗ Next observe that if s, t ∈ Ψ and s − t < ǫ, then, since also kMsk,kMtk = 1, we have kMsM ∗ To finish the proof of item (iv), choose any partition (ωj) of Ψ = T of width at most ǫ > 0. Thus, if s, t ∈ ωj, then kMs − Mtk < ǫ; i.e., the sup norm of the difference of the functions s, t : A → D is less than ǫ. Thus, if sj, tj ∈ ωj, then t k < ǫ. kX Msj M ∗ sj ⊗ E(ωj ) −X Mtj M ∗ tj ⊗ E(ωj )k ≤ 2ǫ. Consequently, choosing a sequence of partitions such that the width of the parti- tions tends to zero, the corresponding Riemann sums form a norm Cauchy sequence and thus converge to some operator. At the same time, this sequence converges WOT to Γ, since hX Msj M ∗ sj ⊗ E(ωj )kw ⊗ f, kz ⊗ gi = Xj sj(z)sj(w)∗k(z, w)hd E(ωj )f, gi. Thus the sequence of Riemann sums converges in norm to Γ. Comparing any Riemann sum whose partition has width at most ǫ > 0 with an appropriate term of the sequence just constructed completes the proof of (iv). converge WOT to R Tψd µ(ψ)T ∗ From Lemma 4.7 and Remark 4.8, the Riemann sums ∆(P, S, µ) and ∆(P, S, Λ) ψ respectively. Hence the net k(T, T ∗)(∆(P, S, µ)) converges to the RHS of item (v). On the other hand, we have k(T, T ∗)(∆(P, S, µ)) = ∆(P, S, Λ). Hence k(T, T ∗)(∆(P, S, µ)) converges WOT to both the right and left hand side of (v) and the result follows. (cid:3) ψ and R Tψd Λ(ψ)T ∗ Using Lemma 9.4, the proof that (sc) implies equation (5), and hence the converse of Theorem 2.9, proceeds as follows. From Lemma 9.4 and the representation (I − Wr(z)Wr(w)∗)k(z, w) = Hr(z)[ZΨ (1 − r2ψ(z)ψ(w)∗)k(z, w)dE(ψ)]Hr(w)∗ it follows that Wr )V = V ∗MHr [I − r2Γ]M ∗ The left hand side converges WOT to I − XX ∗ (because M ∗ so does V ∗(I − MWr M ∗ V. Hr W V = V M ∗ W ) and thus Since, by item (i) of Lemma 9.4 with ω = T, V ∗MHr M ∗ Hr Λ(T), it follows that V ∗MHr (I − Γ)M ∗ Hr V. (32) WOT. V ∗MHr ΓM ∗ Hr V → Λ(T) − I + XX ∗ V converges WOT to 28 S. MCCULLOUGH AND S. SULTANIC Fix a vector m ∈ M. Given ǫ > 0, choose, using (iv) of Lemma 9.4 and Lemma 4.7 respectively, a tagged partition (P, S) such that both, 1 kmk + 1 (33) sj ⊗ E(ωj ) − Γk ǫ >kX Msj M ∗ ǫ >hZ Tψd Λ(ψ)T ∗ ψm, mi −XhTsj Λ(ωj)T ∗ sj m, mi. Using item (i) of Lemma 9.4 and equation (32) respectively, choose 0 < r0 < 1 (depending upon (P, S)) such that for r0 ≤ r < 1, (34) ǫ >XhTsj Λ(ωj)T ∗ ǫ >hV ∗MHr ΓM ∗ sj m, mi −XhTsj Λr(ωj)T ∗ sj m, mi Hr V m, mi − h(Λ(T) − I + XX ∗)m, mi. Note that combining the first inequality in equation (33) with item (ii) of Lemma 9.4 gives, (35) hV ∗MHr X Msj M ∗ sj ⊗ E(ωj)M ∗ =hV ∗MHr [X Msj Ms∗ Hr V m, mi − hV ∗MHr ΓM ∗ j ⊗ E(ωj) − Γ]M ∗ Hr V m, mi Hr V m, mi < C2 ∗ ǫ. Similarly, observe that XhTsj Λr(ωj)T ∗ sj (36) m, mi =XhTsj V ∗MHr (I ⊗ E(ωj))M ∗ =XhV ∗MHr Msj (I ⊗ E(ωj))M ∗ =hV ∗MHr [X Msj (I ⊗ E(ωj))M ∗ Hr sj V T ∗ sj M ∗ Hr sj ]M ∗ Hr m, mi V m, mi V m, mi. Putting it all together, it follows from (34), (35), and (36) that hZ Tψd Λ(ψ)T ∗ ψm, mi − h(Λ(T) − I + XX ∗)m, mi ψm, mi −XhTsj Λ(ωj)T ∗ ≤hZ Tψd Λ(ψ)T ∗ + XhTsj Λ(ωj)T ∗ + hV ∗MHr [X Msj (I ⊗ E(ωj ))M ∗ + hV ∗MHr ΓM ∗ <ǫ + ǫ + C∗ǫ + ǫ. sj m, mi sj m, mi −XhTsj Λr(ωj)T ∗ sj − Γ]M ∗ sj m, mi Hr V m, mi Hr V m, mi − h(Λ(T) − I + XX ∗)m, mi Thus, I − XX ∗ = Λ(T) −Z Tψd Λ(ψ)T ∗ ψ. An application of item (v) of Lemma 9.4 completes the proof. 10. Details on the kernel This section gives the details on the basic facts about our kernel k. It requires a digression into theta functions much of which is borrowed from [28]. Begin by recalling the theta function ϑ1(x) = ϑ1(x, q) = 2q 1 4 sin(x)Π∞ n=1(1 − q2n)(1 − q2ne2ix)(1 − q2ne−2ix), AGLER-COMMUTANT LIFTING ON AN ANNULUS 29 and the Jordan-Kronecker function ∞ f (α, p) = Xn=−∞ 1 − pq2n It is well known that these functions are related by ϑ1(x + y) ϑ1(x)ϑ1(y) f (α, p) = C , αn . where x and y are chosen so that α = e2ix and p = e2iy and C is a constant (independent of x, y). Replacing p with −t and thus y with y + π 2 and letting α = zw∗ gives, k(z, w; t) = C ϑ1(x + y + π 2 ) ϑ1(x)ϑ1(y + π 2 ) From its product expansion, it is evident that the zeros of ϑ1 are q2m = e2ix for integers m and thus k(z, w; t) = 0 if and only if tzw∗ = −q2m for some integer m. Thus, unless t = q2ℓ for some ℓ, there exists points z, w ∈ A such that k(z, w; t) = 0. We are interested in the case t = 1 (p = −1 and y = 0 above) which gives, k(z, w; 1) = k(z, w) = C ϑ1(x + π 2 ) ϑ1(x)ϑ1( π 2 ) . In particular, k(z, w) vanishes if and only if zw∗ = −q2m. In particular, k(z, w) does not vanish for both z and w in the annulus, and further for each fixed w ∈ A, as a function of z, the kernel k(z, w) extends beyond the annulus to a meromorphic function. If zw∗ = e2ix, then −zw∗ = e2i(x+ π k(z,−w) = C 2 ) and therefore, ϑ1(x + π) ϑ1(x + π 2 )ϑ1( π 2 ) ϑ1(x) = − C = C′ ϑ1(x + π 1 , 2 )ϑ1( π 2 ) where C′ = θ1( π k(z, w) 2 )−2. It is evident that C′ > 0. 11. Details on the Test Functions Generally the minimal inner functions on a multiply connected domain can be constructed using the Green's functions or as a product of quotients of theta func- tions. In the case of the annulus the first construction is relatively simple to de- scribe, given unique solutions to the Dirichlet problem. The first step is to construct, given a point a ∈ A, an analytic function with mod- ulus one on the outer boundary B0 and constant modulus on the inner boundary B1 with just one zero, at a, in A. There is a harmonic function w whose boundary values (on the boundary of A) agree with the boundary values of log z − a. There is a constant β and an analytic function f on A so that w = ℜ(f + β log(z)). Here ℜ denotes the real part. Note that β can be computed because a harmonic function u = w − β log(z) is the real part of an analytic function on A if and only 30 S. MCCULLOUGH AND S. SULTANIC if the integral of u around the outer boundary agrees with the integral of u around the inner boundary of A; i.e., +2πβ log(q) = Z log exp(it) − adt −Z log q exp(it) − adt. Indeed, a simple computation shows β = log(a) log(q) . In particular, given two points a, b ∈ A, there is a function unimodular on the boundary of A with zeros precisely a and b (with multiplicity if needed) if and only if log(ab) = q. References [1] Abrahamse, M.B. The Pick interpolation theorem for finitely connected domains. Michigan Math. J. 26 (1979), no. 2, 195 -- 203. [2] Abrahamse, M.B.; Douglas, R.G. A class of subnormal operators related to multiply- connected domains Advances in Math. 19 (1976), no. 1, 106 -- 148. [3] Ambrozie, Calin; Eschmeier, Jorg. A commutant lifting theorem on analytic polyhedra. Topo- logical algebras, their applications, and related topics, 83108, Banach Center Publ., 67, Polish Acad. Sci., Warsaw, 2005. [4] Ambrozie, C.-G.; Englis, M.; Muller, V. Operator tuples and analytic models over general domains in Cn, J. Operator Theory 47 (2002), no. 2, 287 -- 302. [5] Agler, Jim. On the representation of certain holomorphic functions defined on a polydisc. Topics in operator theory: Ernst D. Hellinger memorial volume, 47 -- 66, Oper. Theory Adv. Appl., 48, Birkhauser, Basel, 1990. [6] Agler, Jim. The Arveson extension theorem and coanalytic models. Integral Equations Op- erator Theory 5 (1982), no. 1, 608 -- 631. [7] Agler, Jim; McCarthy, John E. Nevanlinna-Pick interpolation on the bidisk. J. Reine Angew. Math. 506 (1999), 191 -- 204. [8] Agler, Jim; McCarthy, John E. Pick interpolation and Hilbert function spaces. Graduate Studies in Mathematics, 44. American Mathematical Society, Providence, RI, 2002. xx+308 pp. ISBN: 0-8218-2898-3. [9] Agler, Jim;McCarthy, John E. Complete Nevanlinna-Pick kernels. J. Funct. Anal. 175 (2000), no. 1, 111 -- 124. [10] Ambrozie, C.-G. Remarks on the operator-valued interpolation for multivariable bounded analytic functions. Indiana Univ. Math. J. 53 (2004), no. 6, 1551 -- 1576. [11] Archer, Robert. Unitary dilations of commuting contractions. PhD thesis, University of New- castle, 2004. [12] Ball, Joseph A.; Bolotnikov, V. Vladimir Realization and interpolation for Schur-Agler-class functions on domains with matrix polynomial defining function in Cn. J. Funct. Anal. 213 (2004), no. 1, 45 -- 87. [13] Beatrous, Frank; Burbea, Jacob. Reproducing kernels and interpolation of holomorphic func- tions. Complex analysis, functional analysis and approximation theory (Campinas, 1984), 25-46, North-Holland Math. Stud., 125, North-Holland, Amsterdam, 1986. [14] Ball, J. A.; Li, W. S.; Timotin, D.; Trent, T. T. A commutant lifting theorem on the polydisc. Indiana Univ. Math. J. 48 (1999), no. 2, 653 -- 675. [15] Ball, Joseph A.; Trent, Tavan T. Unitary colligations, reproducing kernel Hilbert spaces, and Nevanlinna-Pick interpolation in several variables. J. Funct. Anal. 157 (1998), no. 1, 1 -- 61. [16] Ball, Joseph A.; Trent, Tavan T.; Vinnikov, Victor. Interpolation and commutant lifting for multipliers on reproducing kernel Hilbert spaces. Operator theory and analysis (Amsterdam, 1997), 89 -- 138, Oper. Theory Adv. Appl., 122, Birkhauser, Basel, 2001. [17] de Branges, Louis; Rovnyak, James. Canonical models in quantum scattering theory. 1966 Perturbation Theory and its Applications in Quantum Mechanics (Proc. Adv. Sem. Math. Res. Center, U.S. Army, Theoret. Chem. Inst., Univ. of Wisconsin, Madison, Wis., 1965) pp. 295 -- 392 Wiley, New York [18] Dritschel, Michael A.; McCullough, Scott. The failure of rational dilation on a triply con- nected domain. J. Amer. Math. Soc. 18 (2005), no. 4, 873918 (electronic). AGLER-COMMUTANT LIFTING ON AN ANNULUS 31 [19] Dritschel, Michael A.; McCullough, Scott A., Test functions, kernels, realizations and in- terpolation. Operator theory, structured matrices, and dilations, 153 -- 179, Theta Ser. Adv. Math., 7, Theta, Bucharest, 2007. [20] Dritschel, Michael A.; Marcantognini, Stefania; McCullough, Scott. Interpolation in Semi- groupoid Algebras. Journal fur die Reine und Angewandte Mathematik 2006. [21] Douglas, R. G.; Muhly, P. S.; Pearcy, Carl. Lifting commuting operators. Michigan Math. J. 15 1968 385 -- 395. [22] Eschmeier, Jorg. Tensor products and elementary operators. J. Reine Angew. Math. 390 (1988), 4766. [23] Foias, Ciprian; Frazho, Arthur E. The commutant lifting approach to interpolation problems. Operator Theory: Advances and Applications, 44. Birkhauser Verlag, Basel, 1990. [24] Fritz Gesztesy, Rudi Weikard, Maxim Zinchenko. On a class of Model Hilbert Spaces. arXiv:1111.0645v1. [25] Jury, M.T. An improved Julia-Caratheodory theorem for Schur-Agler mappings of the unit ball. arXiv:0707.3423v1. [26] McCullough, Scott, Isometric representations of some quotients of H∞ of an annulus. Inte- gral Equations Operator Theory 39 (2001), no. 3, 335-362. [27] McCullough, Scott The trisecant identity and operator theory. Integral Equations Operator Theory 25 (1996), no. 1, 104-127. [28] McCullough, Scott; and Shen, Li Chien, On the Szego's kernel of an annulus. Proc. Amer. Math. Soc. 121 (1994), no. 4, 1111 -- 1121. [29] McCullough, Scott; and Sultanic, Saida Ersatz commutant lifting with test functions. Com- plex Anal. Oper. Theory 1 (2007), no. 4, 581 -- 620. [30] Sz.-Nagy, Bela. Sur les contractions de l'espace de Hilbert. Acta Sci. Math. Szeged 15, (1953). 87 -- 92 [31] Sarason, Donald. Generalized interpolation in H∞. Trans. Amer. Math. Soc. 127 1967 179 -- 203. Department of Mathematics, University of Florida, Box 118105, Gainesville, FL 32611-8105, USA Sarajevo School Of Science and Technology, Bistrik 7, Sarajevo 71000, Bosnia- Herzegovina E-mail address: [email protected] E-mail address: [email protected]
1804.01370
1
1804
2018-04-04T12:40:25
The Perron solution for vector-valued equations
[ "math.FA" ]
Given a continuous function on the boundary of a bounded open set in $\mathbb{R}^d$ there exists a unique bounded harmonic function, called the Perron solution, taking the prescribed boundary values at least at all regular points (in the sense of Wiener) of the boundary. We extend this result to vector-valued functions and consider several methods of constructing the Perron solution which are classical in the real-valued case. We also apply our results to solve elliptic and parabolic boundary value problems of vector-valued functions.
math.FA
math
THE PERRON SOLUTION FOR VECTOR-VALUED EQUATIONS M. KREUTER Abstract. Given a continuous function on the boundary of a bounded open set in Rd there exists a unique bounded harmonic function, called the Perron solution, taking the prescribed bound- ary values at least at all regular points (in the sense of Wiener) of the boundary. We extend this result to vector-valued functions and consider several methods of constructing the Perron solution which are classical in the real-valued case. We also apply our re- sults to solve elliptic and parabolic boundary value problems of vector-valued functions. 1. Problem Setting, Existence and Uniqueness Let Ω ⊂ Rd be an open and bounded set and let X be a (real) Banach space. The set of all harmonic functions will be denoted by H(Ω, X) := {u ∈ C 2(Ω, X), ∆u = 0}. Given a function f ∈ C(∂Ω, X) we consider the classical Dirichlet problem (u ∈ H(Ω, X) ∩ C(Ω, X) u∂Ω = f. A function u satisfying the above is called a classical solution. Note that the range of f is separable. Hence we will without loss of gener- ality assume that X is separable, since we may restrict our arguments to the smallest subspace of X containing the range of f . In the case X = R, it is well known that a generalized solution, the Perron solution, exists. This solution is the unique function which is bounded, harmonic and satisfies the boundary values at least on the set ∂regΩ of all regular points of ∂Ω, c.f. [AG01, Chapter 6.6], [Hel09, Chapter 2] or [Kel66]. The set ∂regΩ can be described by potential theoretic means, e.g. a point z ∈ ∂Ω is regular if and only if Ωc is not thin at z which in turn can be described via Wiener's criterion [AG01, Date: May 3, 2019. 2010 Mathematics Subject Classification. 31B20,31C05,46E40. Key words and phrases. Perron solution, vector-valued functions, Banach space, harmonic functions, generalized solution, Dirichlet problem, Poisson problem, Heat equation, irregular domain. 1 PERRON SOLUTIONS 2 Theorems 7.5.1 and 7.7.2]. The set of all bounded harmonic functions will be denoted by Hb(Ω, X). Theorem 1.1. Let Ω ⊂ Rd be open and bounded and let X be a real Banach space. For every f ∈ C(∂Ω, X) there exists a unique function Hf , called the Perron solution, which satisfies (Hf ∈ Hb(Ω, X) limξ→z Hf (ξ) = f (z) for all z ∈ ∂regΩ. Remark 1.2. The real-valued Dirichlet problem has a classical solu- tion for every boundary data if and only if ∂regΩ = ∂Ω. It follows that if the Dirichlet problem has a solution for every real-valued con- tinuous boundary data, then it also has a solution for every X-valued continuous boundary data. To prove this we will need some auxiliary results. The (d − 1)- dimensional Hausdorff measure will be denoted by σd−1. Theorem 1.3. [Are16, Lemma 5.1, Proposition 5.3 & Theorem 5.4] (i) Let u : Ω → X be locally bounded and suppose that there exists a separating subset W ⊂ X ′ such that hu, x′i := hu(·), x′i is har- monic for all x′ ∈ W . Then u is harmonic. (ii) If u ∈ H(Ω, X) then Poisson's Integral Formula r2 0 − x − x02 1 u(x) = r0σd−1(B(0, 1))Z∂B(x0,r0) x − sd u(s) dσd−1(s) holds for all x0 ∈ Ω and r0 > 0 such that B(x0, r0) ⊂⊂ Ω. (iii) Let (ui : Ω → X)i∈I be a bounded net of harmonic functions. Suppose that (ui(x))i∈I converges for each x ∈ Ω, then (ui)i∈I converges uniformly on compact subsets of Ω and the limit is a harmonic function. If K ⊆ X then conv(K) will denote the closed convex hull of K. Note that if K is compact, then so is conv(K), [AB06a, Theorem 5.35]. Proposition 1.4 (Maximum Principle for vector-valued functions). Let u ∈ H(Ω, X) ∩ C(Ω, X). Then for all ξ ∈ Ω we have (i) u(ξ) ∈ conv(u(∂Ω)) (ii) ku(ξ)k ≤ maxz∈∂Ω ku(z)k. Proof. It suffices to show (i). Suppose that u(x) /∈ conv(u(∂Ω)) =: M, then by the Hahn-Banach Theorem there exists a functional x′ ∈ X ′ such that Re hu(x), x′i > supm∈M Re hm, x′i. This contradicts the Maximum Principle for real-valued functions. (cid:3) Remark 1.5. The proof of Proposition 1.4 shows that it is actually enough to have u ∈ H(Ω, X) ∩ Cb(Ω, X) such that u(ξ) ⇀ f (z) as ξ → z ∈ ∂Ω. PERRON SOLUTIONS 3 The last tool we will need for the proof is the following density result. Lemma 1.6. Let K be a compact space and X be a Banach space. If W ⊂ C(K, R) and Y ⊂ X are dense in the respective spaces, then the set (f ∈ C(K, X), f = n Xj=1 fj ⊗ xj, fj ∈ W, xj ∈ Y) is dense in C(K, X). i=1 Ui = K. Let (ϕi)n subordinate to the collection (Ui)n kϕi − ψik∞ < εn−1kf k−1 i=1 ψi ⊗ xi. One easily computes that kf − gk < 3ε. n=1 gn ⊗ xn Proof. Let f ∈ C(K, X) and ε > 0. For every k ∈ K there exists an open set U ∋ k such that kf (k) − f (l)k ≤ ε for all l ∈ U. Since K is compact we may choose a finite number of ki ∈ K and Ui as i=1 be a partition of unity i=1. Further choose ψi ∈ W such that ∞ , xi ∈ Y such that kf (ki) − xik < ε and define (cid:3) above such that Sn g :=Pn Proof of Theorem 1.1. First suppose that f is of the form f =PN where gn ∈ C(∂Ω, R) and xn ∈ X. The function Hf :=PN n=1 Hgn ⊗ xn satisfies the claim. For an arbitrary function f ∈ C(∂Ω, X) there ex- ists a sequence of functions fn of the above form which converges to f . By the maximum principle the Perron solutions Hfn form a Cauchy sequence in Cb(Ω, X) and hence converge uniformly on compact sets to a harmonic function Hf . Let z ∈ ∂Ω be regular, then kHf (x) − f (z)k ≤kHf − Hfnk∞ + kHfn(x) − fn(z)k + kfn(z) − f (z)k. For large n ∈ N the first and last term will be smaller than a given ε > 0, thus lim sup x→z kHf (x) − f (z)k ≤ 2ε + lim x→z kHfn(x) − fn(z)k = 2ε. Letting ε → 0 yields existence. By composing Hf with an arbitrary functional x′ ∈ X ′ the uniqueness of Hf follows from the uniqueness of the real-valued Perron solution and the Hahn-Banach Theorem. (cid:3) The Dirichlet problem is one of the oldest problems in partial dif- ferential equations and various different constructions of the Perron solution have been given. For the remainder of this section and during the next section we want to describe how several of these constructions work in the vector-valued case as well. Corollary 1.7 (Wiener's construction of the Perron solution). Let f ∈ C(∂Ω, X). For any continuous extension F of f to the whole of Ω and every sequence (ωn) of Dirichlet regular open subsets of Ω such that ωn ⊂⊂ ωn+1 and Sn ωn = Ω we have: If Hn denotes the solution of the PERRON SOLUTIONS 4 Dirichlet problem for ωn with boundary data F∂ωn, then Hn converges to Hf uniformly on compact subsets. Proof. The result is well known in the real-valued case, see e.g. [Kel66, Theorems I & II]. By the Maximum Principle we have that Hn(x) ∈ conv(F (∂ωn)) ⊆ conv(F (Ω)). Hence Hn is uniformly bounded by a constant independent of n. Poisson's Integral Formula implies that (Hn)n∈N is equicontinuous. Further conv(F (Ω)) is compact. The Arzela- Ascoli Theorem implies that each subsequence of (Hn)n∈N has a sub- sequence (Hnk)k∈N converging to some function v on compact subsets. The real-valued case implies that hHnk, x′i → Hhf,x′i and hence v = Hf . As the subsequence was chosen arbitrarily the claim follows. (cid:3) Wiener's construction is justified since an extension F and an ex- hausting sequence of regular sets always exist, c.f. [Dug51, Theorem 4.1] and [AG01, Corollary 6.6.13]. The next construction is due to Poincar´e and can be found in [Hil05, Theorem 2]. We will omit the proof since it works similar to the proof of Theorem 1.1. Proposition 1.8 (Poincar´e's construction of the Perron solution). Let f ∈ C(∂Ω, X) and let F be a continuous extension of f to the whole of Ω. Let Bi be a sequence of non-trivial balls such that Bi ⊂⊂ Ω and Si Bi = Ω and let in be a sequence such that every natural number appears infinitely often in in. For a continuous function u and a ball B in Ω we let uB be the function defined by u outside of B and the solution to the Dirichlet problem in B with boundary data u∂B inside B. Define a sequence un inductively via u0 = F and un = (un−1)Bin . Then the sequence un converges to Hf uniformly on compact subsets. Another possible way of constructing the Perron solution is via har- monic measures. We summarize the construction of these measures and refer to [AG01, Chapter 6.4] for more information. For every ξ ∈ Ω the mapping C(∂Ω, R) → R f 7→ Hf (ξ) is well-defined, linear and positive. It follows from the Riesz Represen- tation Theorem that there exists a unique probability measure µξ on the Borel algebra B(∂Ω) such that (1) Hf (ξ) =Z∂Ω f dµξ. Corollary 1.9 (Construction via harmonic measures). Let f ∈ C(∂Ω, X). The function Hf can be defined via (1). Proof. Let x′ ∈ X ′. Since x′ commutes with integration the claim holds due to the uniqueness of the Perron solution. (cid:3) PERRON SOLUTIONS 5 Remark 1.10. Another way to construct the Perron solution is to use Hilbert space theory in the space H 1(Ω, R). This construction can be found in [Hil05, Theorem 1], [DL90, II,§7, 4.] or [AD08]. Obviously, this approach fails if X is not a Hilbert space. However, if X happens to be a Hilbert space, the Perron solution can be constructed via Hilbert space methods as in the real-valued case with minor changes and we will omit to carry out the proof. 2. Perron's Method on Banach Lattices The classical method to obtain the Perron solution is to construct it as the pointwise supremum of subsolutions, see e.g. c.f. [AG01, Chapter 6.6], [Hel09, Chapter 2] or [Kel66, Theorem II]. This crucially depends on the order of R and hence makes no sense in general Banach spaces. Thus throughout this section let X be a Banach lattice. We will use the partial ordering on X to generalize Perron's approach. As in the real-valued case a function v ∈ C(Ω, X) is called subhar- monic if for all ξ ∈ Ω there exists R > 0 such that B(ξ, R) ⊂⊂ Ω and we have v(ξ) ≤ −Z∂B(ξ,r) v dσd−1 for all 0 < r < R. Testing with positive functionals and using the real- valued case [ABR01, Chapter 11, Exercise 5] reveals that this inequality actually holds for all 0 < r < dist(ξ, ∂Ω). Let f ∈ C(∂Ω, X). Then a continuous subsolution of the Dirichlet problem with boundary data f is a function v ∈ C(Ω, X) that is subharmonic and satisfies v(z) ≤ f (z) for all z ∈ ∂Ω. Analogously one defines superharmonic functions and continuous supersolutions. The sets of continuous sub-/supersolutions will be denoted by CS − f respectively. f and CS + Remark 2.1. The above definition is common in books on partial differential equations, such as [GT01] and [ABR01], where they are subsolutions". In potential theory subsolutions are defined just called " in a weaker sense, see , [AG01] or [Hel09]. For this reason we chose the name continuous subsolution". " If X = R then f ∈ C(∂Ω, R) is automatically bounded and hence v− ≡ min f and v+ ≡ max f are continuous sub-/supersolutions. In an arbitrary Banach lattice, v− and v+ can only exist if f is bounded there exist s, S ∈ X such that s ≤ f (ξ) ≤ S. The in order, i.e. set of all order bounded and continuous functions will be denoted by Cob(∂Ω, X). We give a counterexample. Example 2.2. Let (zn)n∈N be a pairwise distinct sequence of points in ∂Ω converging to z. Set f (zn) := en n ∈ ℓ1 and f (z) := 0, then f is continuous. Using Dugundi's Extension Theorem [Dug51, Theorem PERRON SOLUTIONS 6 4.1] we may extend f to a function f ∈ C(∂Ω, ℓ1) that is not bounded from above in the order of ℓ1. Hence there is no guarantee that CS + f 6= ∅. For f ∈ Cob(∂Ω, X) with lower bound s and upper bound S the sets f and CS + f are nonempty as they contain the functions v− ≡ s and CS − v+ ≡ S respectively. Proposition 2.3 (Maximum Principle for lattices). Let u, v ∈ C(Ω, X) and let v be subharmonic and u harmonic. If v ≤ u on ∂Ω then v ≤ u in Ω. Proof. Let x′ ∈ X ′ + then hv, x′i and hu, x′i are real-valued subharmonic respectively harmonic functions satisfying hv, x′i ≤ hu, x′i on ∂Ω. By the real-valued maximum principle we have that hv, x′i ≤ hu, x′i in Ω. Since X ′ (cid:3) + determines positivity the claim follows immediately. Let f ∈ C(∂Ω, X) and suppose that CS − f are non-empty. v(ξ) and the Suppose further that the pointwise supremum supv∈CS − v(ξ) exist for all ξ ∈ Ω. Then we will denote pointwise infimum inf v∈CS + them by H f and H f . The following obvious result is the motivation for Perron's method. f f and CS + f Proposition 2.4. Suppose that the Dirichlet problem with boundary data f ∈ C(∂Ω, X) admits a classical solution Hf , then H f = H f = Hf . Proof. Since Hf is harmonic and takes on the prescribed boundary values it follows that Hf ∈ CS − f ∩CS + f . The Maximum Principle implies that Hf = sup CS − f = inf CS + f . (cid:3) In the real-valued case we have Theorem 2.5. Let f ∈ C(∂Ω, R), then H f = H f = Hf . We will give a proof of this theorem later on. Even if the boundary data is not order bounded it might still be true that CS ± f 6= ∅. For example this is the case if the domain is Dirichlet regular. The following example covers a variety of domains which still allow Perron's method to work. It allows us to consider such common examples of irregular sets as the punctured disk BC(0, 1)\{0}. Example 2.6. Let Ω∗ be a Dirichlet regular domain and let z1, . . . , zk ∈ Ω∗. If Ω := Ω∗\{z1, . . . , zk} then for every f ∈ C(∂Ω, X) the sets CS ± f are non-empty. To see this note that {z1, . . . , zk} is the set of irregular points of Ω. Thus the uniqueness of the Perron solution implies that Hf = Hf ∗ for f ∗ := f∂Ω∗. Hence Hf (zj) is well defined and Hf ∈ PERRON SOLUTIONS 7 C(Ω, X). We define k v±(ξ) := Hf (ξ) ± Hf (zj) − f (zj), Xj=1 then it is easy to see that v± ∈ CS ± f . The construction shows that we may also withdraw an infinite number of points in a regular domain as long as every connected component contains only a finite number of these points. In some cases, the Banach lattice itself has the property that sub- and supersolutions exist. Example 2.7. Let X be an AM-space, see e.g. [AB06b, Definition 4.20], then every compact set is order bounded [AB06b, Theorem 4.30]. Hence for every continuous boundary data f the sets CS ± f are non- empty. Let u ∈ C(Ω, X) be subharmonic and let B := B(x, r) ⊂⊂ Ω. We define the harmonic lifting via uB :=(u Hu∂B on Ω\B on B. The Maximum Principle implies that uB ≥ u and it follows that uB is still subharmonic. Further if v ∈ C(Ω, X) is another subharmonic function, then the function u ∨ v is subharmonic as well. Let F ⊂ C(Ω, X) be a set of subharmonic functions. We say that F is a Perron family if (a) F is upwards directed (a sufficient condition would be that F is closed under taking the pointwise supremum of two functions) (b) uB ∈ F for every u ∈ F and every ball B := B(x, r) ⊂⊂ Ω. A Banach lattice X is called order complete if every order bounded subset of X has a supremum and an infimum. Common examples are Lp-spaces (p < ∞), a counterexample is the space C([0, 1]). A functional x′ ∈ X ′ + is called order continuous if for every upwards directed set A ⊂ X which has a supremum, we have hsup A, x′i = suphA, x′i. Proposition 2.8. Let X be a Banach lattice which is order complete and assume that the order continuous functionals separate. Suppose that F ⊂ C(Ω, X) is a Perron family which is pointwise order bounded and locally bounded. Then the function is well defined and harmonic. ξ 7→ sup v∈F v(ξ) Proof. The case X = R is well known, see [GT01, Theorem 2.12]. In general, the supremum exists since X is order complete. Now let PERRON SOLUTIONS 8 + be order continuous, then hF , x′i is a Perron family as well and x′ ∈ X ′ we have that hsupv∈F v, x′i = supv∈F hv, x′i. The case X = R shows that the latter is harmonic. By Theorem 1.3(i) it follows that supv∈F v is harmonic. (cid:3) Example 2.9. (a) If X has order continuous norm, see [AB06b, Sec- tion 4.1], then X automatically satisfies the conditions of Proposi- tion 2.8. Examples are reflexive Banach lattices and also L1. αi. γi ≥ hx, x′ (b) If X is a Banach lattice, then X ′ satisfies the conditions of Proposi- α) be upwards directed in X ′. We define tion 2.8. To see this let (x′ x′ : X+ → R via x 7→ supαhx, x′ Immediately x′ is sublinear on X+. But for every α, β there exists γ such that x′ β ≤ x′ α, x′ γ, from which we obtain that hx + y, x′ αi + hy, x′ βi for every x, y ∈ X+. Taking the suprema on both sides we obtain that x′ is also superlinear and hence linear on X+. Extend x′ to all of X via x 7→ hx+, x′i − hx−, x′i to obtain a linear functional. From the defi- nition of x′ one immediately has that x′ = supα x′ α. Further one has that every x ∈ X+ ⊂ X ′′ is a positive order continuous functional and X separates X ′ naturally. To show the order completeness of X ′ let A ⊂ X ′ be any subset which has an upper bound. Let A be the set of all finite suprema of elements in A, then A is upwards directed. It follows from the first considereation that sup A exists and one immediately sees that sup A = sup A. If f ∈ C(∂Ω, X) such that CS − f is nonempty, then CS − f and −CS + are Perron families which are bounded from above. So we obtain f Theorem 2.10. Suppose that X is order complete and that the order continuous functionals separate. Let f ∈ C(∂Ω, X) such that CS ± f 6= ∅. Then H f and H f exist and are harmonic. Suppose that CS ± f 6= ∅ for every boundary data f ∈ C(∂Ω, X) and that X is as in Theorem 2.10. The question we want to study now is whether in this situation H f = H f = Hf . Consider the operators T −f 7→ H f and T +f 7→ H f . It is easy to see that they are super-/ sublinear. By what we have seen so far, they map into the bounded harmonic functions Hb(Ω, X) and satisfy T −f = T +f = Hf if a clas- sical solution of the Dirichlet problem with boundary data f exists. This suggests an operator theoretic approach. The operators T − and T + are monotone, i.e. if f ≤ g, then T −f ≤ T −g. For linear operators we have Theorem 2.11 (Keldysh). Let T be an operator that maps C(∂Ω, R) into the bounded harmonic functions on Ω such that T (i) is linear (ii) positive (iii) and maps f to the classical solution if it exists. PERRON SOLUTIONS Then T f = Hf for all f ∈ C(∂Ω, R) Proof. See [Lan72, Theorem 4.11]. 9 (cid:3) Remark 2.12. The original proof is available in Russian only. For a reference, see [Lan72]. Landkof's proof differs from the original one. A simplified proof using Keldysh's ideas was given by Brelot [Bre61] in French. We will reproduce this proof in Theorem 2.22. Corollary 2.13 (Keldysh's Theorem for lattices). Let X be a Banach lattice and let T : C(∂Ω, X) → Hb(Ω, X) such that T is (i) is linear (ii) positive (iii) and maps f to the classical solution if it exists. Then T f = Hf for all f ∈ C(∂Ω, X) Proof. We first show that for x ∈ X+ and f ∈ C(∂Ω, R) we have T (f ⊗ x) = Hf ⊗x. Since Hf ⊗x = Hf ⊗ x it is enough to show that hT (f ⊗ x), x′i = Hf for each x′ ∈ X ′ + such that hx, x′i = 1. Choose such x′ and define the operator S : C(∂Ω, R) → H(Ω, R) f 7→ hT (f ⊗ x), x′i. It is immediately clear that S satisfies the conditions of Keldysh's The- orem and hence hT (f ⊗ x), x′i = Hf . obtain T f = Hf for any function of the form f =PN Next we may consider arbitrary x = x+ − x− by linearity and further n=1 fn ⊗ xn. Now let f ∈ Cob(∂Ω, X) be arbitrary, then there exist functions fn of the above form such that fn → f . Since T is positive, linear and its range is contained in the Banach lattice Cb(Ω, X) it is also continuous and hence the claim follows for f . (cid:3) Since T + and T − are not linear, we cannot immediately apply Keldysh's Theorem and need to show some further results. An ordered vector space W is said to have the interpolation property if for every two sets A, B ⊂ W such that a ≤ b for all a ∈ A, b ∈ B there exists w ∈ W such that a ≤ w ≤ b for all a ∈ A, b ∈ B. A function p mapping a vector space V into W is called sublinear if p(λv) = λp(v) and p(v1 + v2) ≤ p(v1) + p(v2) for all λ ≥ 0 and v, v1, v2 ∈ V . Theorem 2.14 (Hahn-Banach-Kantorowicz). Let V be a vector space and U ⊂ V a subspace. Let W be and ordered vector space which has the interpolation property. Let ϕ : U → W PERRON SOLUTIONS 10 be linear and p : V → W be sublinear such that ϕ ≤ p on U. Then ϕ can be extended to V such that ϕ ≤ p on V . Proof. Using the interpolation property, this can be established analo- gously to the real-valued result, see [Bre11, Theorem 1.1]. (cid:3) Corollary 2.15. Given p as in the Hahn-Banach-Kantorowicz Theo- rem, we have p(v) = max{ϕ(v), ϕ : V → W linear, ϕ ≤ p} for every v ∈ V . Proof. For v ∈ V define ϕ(λv) := λp(v) for all λ ∈ R and extend ϕ to V via the Hahn-Banach-Kantorowicz Theorem. The claim follows immediately. (cid:3) We show that one can apply the above considerations to the case where W is the space of harmonic and bounded functions. Lemma 2.16. Let A, B ⊂ Hb(Ω, X) such that a ≤ b for all a ∈ A, b ∈ B. If X is order complete and the order continuous functionals separate, then there exists a function h ∈ Hb(Ω, X) such that a ≤ h ≤ b holds for all a ∈ A, b ∈ B. Proof. Consider the set A∗ := {a ∈ Cb(Ω, X), a subharmonic, a ≤ b for all b ∈ B} ⊃ A. One easily sees that A∗ is a Perron family and hence has a harmonic supremum h by Proposition 2.8. The function h satisfies the claim. (cid:3) Remark 2.17. It follows from the preceding lemma, that Hb(Ω, X) is an order complete vector lattice. We can also show that Hb(Ω, R) is even a Banach lattice. Since this is not necessary for our considerations, we have deferred this discussion to the Appendix A. Proposition 2.18. Let p : C(∂Ω, X) → Hb(Ω, X) such that (i) p is sublinear (ii) p is monotone, i.e. p(f ) ≤ p(g) if f ≤ g (iii) p maps f to the classical solution of the Dirichlet problem if it exists. Then p(f ) = Hf for all f ∈ C(∂Ω, X) Proof. Consider the set M := {ϕ : C(∂Ω, X) → Hb(Ω, X) linear , ϕ ≤ p}. We will show that M actually consists of only one element, namely f 7→ Hf . The claim then follows since p(f ) = maxϕ∈M ϕ(f ). Let ϕ ∈ M. Then ϕ is positive. In fact, let f ≤ 0. Then p(f ) ≤ p(0) = 0 PERRON SOLUTIONS 11 and hence ϕ(f ) ≤ 0. Now let f ∈ C(∂Ω, X) such that Hf is a classical solution. Then ϕ(f ) ≤ Hf . On the other hand −ϕ(f ) = ϕ(−f ) ≤ p(−f ) = H−f = −Hf . Hence ϕ(f ) = Hf . In total, ϕ satisfies the conditions of Keldysh's Theorem which finishes the proof. (cid:3) Theorem 2.19. Assume that X is order complete, the order contin- uous functionals separate and that CS ± f 6= ∅ for every f ∈ C(∂Ω, X). Then H f = H f = Hf for all f ∈ C(∂Ω, X). Proof. The mapping T + satisfies the assumptions of the preceding proposition. The reasoning for T − is analogous. (cid:3) Examples 2.6, 2.7 and 2.9 describe situations in which the conditions of Theorem 2.19 are fulfilled. We also obtain the Proof of Theorem 2.5. If X = R, the assumptions of Theorem 2.19 are satisfied. (cid:3) Open Problem 2.20. Is there a function f ∈ C(∂Ω, X) for some Ω ⊂ Rd open and bounded and some Banach lattice X such that CS ± f = ∅? A negative answer to the problem above would imply that Perron's method works whenever the Banach lattice is order complete and has the property that the order continuous functionals separate the space. To avoid this open question we will now demand a stronger continuity assumption of the boundary values and reevaluate the operator the- oretic approach of Theorem 2.19. Let X be a Banach lattice which is order complete and whose order continuous functionals separate. A function f : ∂Ω → X is called order boundedly continuous at z0 ∈ ∂Ω if for every ε > 0 there exists y ∈ X+ with kyk ≤ ε and δ > 0 such that f (z) − f (z0) ≤ y for all z ∈ ∂Ω for which z − z0 ≤ δ. The set of all order boundedly continuous functions on ∂Ω will be denoted by OBC(∂Ω, X). Example 2.21. If X is an AM-space, then every continuous function f ∈ C(∂Ω, X) is automatically an element of OBC(∂Ω, X). Indeed, for z0 ∈ ∂Ω and ε > 0 let δ > 0 such that kf (z) − f (z0)k < ε whenever z − z0 < δ. By [AB06b, Theorem 4.30] there exists y := sup{f (z) − f (z0), z−z0 < δ} and the proof of this theorem can easily be modified to show that kyk < ε. Hence f is order boundedly continuous at z0. A simple compactness argument shows that every f ∈ OBC(∂Ω, X) is order bounded and hence the sets CS + f are nonempty. It follows from Proposition 2.10 that the operators H f and H f are well defined. We now prove Keldysh's Theorem in this case. f and CS − PERRON SOLUTIONS 12 Theorem 2.22 (Keldysh's Theorem for OBC-functions). Let T : OBC(∂Ω, X) → Hb(Ω, X) such that (a) T is linear (b) T is positive (c) T f = Hf if Hf is a classical solution. Then T f = Hf for all f ∈ OBC(∂Ω, X). Proof. Let f ∈ OBC(∂Ω, X), z0 ∈ ∂Ω regular, ε > 0. There exists a δ > 0 and a y ∈ X+ such that f (z) − f (z0) ≤ y if z ∈ ∂Ω such that z − z0 ≤ δ. By [Bre61, Lemme 1] there exists a harmonic function w in Ω which is non negative, continuous up to the boundary of Ω and satisfies w(z0) < ε as well as w(z) ≥ 1 for all z ∈ ∂Ω such that z − z0 > δ. Let s := 2 sup∂Ω f , then we have that f ≤ f (z0) + y + w ⊗ s. Indeed if z − z0 ≤ δ, then this follows from the choice of δ and the positivity of w ⊗ s. Otherwise it follows from the choice of s, the properties of w and the positivity of y. Using the properties of T we obtain and hence for every x′ ∈ X ′ + we have T f ≤ f (z0) + y + w ⊗ s lim sup ξ→z0 ξ∈Ω hT f (ξ), x′i ≤ hf (z0), x′i + εkx′k + lim ξ→z0 ξ∈Ω w(ξ)hs, x′i ≤ hf (z0), x′i + εkx′k + εhs, x′i. Since ε > 0 was arbitrary, it follows that lim sup hT f (ξ), x′i ≤ hf (z0), x′i and analogously ξ→z ξ∈Ω lim inf ξ→z ξ∈Ω hT f (ξ), x′i ≥ hf (z0), x′i. These two inequalities show that hT f (ξ), x′i → hf (z0), x′i and by the uniqueness of the Perron solution this implies that T f = Hf . (cid:3) From here on we can work analogously to the proof of Theorem 2.19 and obtain Theorem 2.23. Let X be a Banach lattice which is order complete such that the order continuous functional separate. Then H f = H f = Hf for every f ∈ OBC(∂Ω, X). For the rest of this section, we want to consider Perron's method on a fixed function rather than the operator theoretic approach above where we had to consider all functions on the boundary at once. We will only consider order bounded functions on special Banach lattices PERRON SOLUTIONS 13 and reconstruct special subsolutions from the real-valued case. As a result we will have Theorem 2.24. Let X be an AM-space with unit (see [AB06b, p. 195]) or X = ℓp where 1 ≤ p ≤ ∞. Then for every f ∈ Cob(∂Ω, X) we have that H f = H f = Hf . Note that in the case of an AM-space every bounded function is already order bounded, see Example 2.7. But Theorem 2.19 does not apply since X is not order complete in general, e.g. take X = C([0, 1]). Lemma 2.25. Let X be a Banach lattice and f ∈ C(∂Ω, X). Suppose that W ⊂ X ′ + determines positivity and that for all x′ ∈ W, ε > 0, v ∈ CS − f satisfying hv(ξ), x′i ≥ v(ξ) − ε. Then H f = uf . The analogous result for H f holds. hf,x′i, ξ ∈ Ω there exists a function v ∈ CS − Proof. By symmetry it is enough to show the claims for H f . Let ξ ∈ Ω, ε ≥ 0. By Theorem 2.5 we may choose v ∈ CS − that hf,x′i such v(ξ) ≥ Hhf,x′i(ξ) − ε = hHf (ξ), x′i − ε. Then the function v corresponding to v satisfies hv(ξ), x′i ≥ hHf (ξ), x′i− 2ε. By the Regular Maximum Principle we have that Hf (ξ) ≥ w(ξ) for every w ∈ CS − f }, then hs, x′i ≥ hv(ξ), x′i ≥ hHf (ξ), x′i − 2ε. Since ε > 0 and x′ ∈ W were arbitrary we obtain that s ≥ Hf (ξ). (cid:3) f . Let s be any upper bound of {w(ξ), w ∈ CS − Proof of Theorem 2.24. We will show that in both cases we can con- struct subsolutions as in Lemma 2.25. First let X be an AM-space with unit. By Kakutani's Theorem [AB06b, Theorem 4.29] we may assume that X = C(K) for some compact topological space K. Replacing f by f + kf k∞ we may assume that f ≥ 0. The family (δa)a∈K determines positivity. Let a ∈ K, ε > 0 and v ∈ CS − f (·)(a). For every z ∈ ∂Ω we have that v(z) − ε < f (z)(a). Since the function (z, k) 7→ f (z)(k) is continuous and ∂Ω is compact there exists a neighbourhood U ∗ of a such that f (z)(k) > v(z) − ε for all z ∈ ∂Ω and all k ∈ U ∗. By Tietze's Extension Theorem there exists a function g ∈ C(K) such that if k = a = 1, if k /∈ U ∗ = 0, ∈ [0, 1], otherwise . g(k)  Then the function ξ 7→ v(ξ)(k) := (v(ξ) − ε)g(k) satisfies the demands. Now let X = ℓp and let s be an upper bound of f ∈ Cob(∂Ω, X). We choose the coordinate functionals en as separating subset of X ′. Let PERRON SOLUTIONS fn := hf, eni and let v ∈ CS + f . Set v(ξ)k :=(v if k = n sk if k 6= n. One easily sees that v satisfies the demands. 14 (cid:3) 3. Application: The Poisson Problem for Elliptic Operators Again, let X be a Banach space and Ω ⊂ Rd be open and bounded. Recall that a function u ∈ C(Ω, X) is uniformly Holder continuous if there exist constants C > 0 and 0 < α < 1 such that ku(ξ) − u(η)k ≤ Cξ − ηα holds for all ξ, η ∈ Ω. The set of all such mappings will be denoted by C α(Ω, X). A function u ∈ C(Ω, X) will be called locally Holder contin- uous if for every ω ⊂⊂ Ω there exists αω such that u ∈ C αω (ω, X). Fur- ther we denote by C 2,α(Ω, X) the set of all functions u ∈ C 2(Ω, X) such that the second derivatives of u are elements of C α(Ω, X). Equipped with the norms kukα := kukC(Ω,X) + sup ξ,η∈Ω ku(ξ) − u(η)k ξ − ηα respectively kuk2,α := kukC 2(Ω,X) + Xβ=2 sup ξ,η∈Ω kDβu(ξ) − Dβu(η)k ξ − ηα the spaces C α(Ω, X) and C 2,α(Ω, X) are Banach spaces. In this section we consider a second order differential operator L := d Xi,j=1 aij(·)Dij + d Xi=1 bi(·)Di + c(·) whose coefficients aij, bi and c are in C α(Ω, R). Further assume that c ≤ 0 and that the aij are symmetric and satisfy a strict ellipticity condition ξiaij(x)ξj ≥ λξ2 d Xi,j=1 for some λ > 0. For f ∈ C(∂Ω, X) and g ∈ C(Ω, X) consider the Poisson problem for L given by (Lu = g u = f in Ω on ∂Ω PERRON SOLUTIONS 15 A classical solution to this problem is a function u ∈ C 2(Ω, X) ∩ C(Ω, X) which satisfies the above. We want to apply the results of the previous sections to show Theorem 3.1. Suppose that Ω has a C 2,α-boundary and let L be as above. Assume that g ∈ C α(Ω, X) and f = F∂Ω, where F ∈ C 2,α(Ω, X). Then there exists a unique classical solution u to the Poisson problem. Moreover u ∈ C 2,α(Ω, X). Before we can show this we need to consider the special case L = ∆. The fundamental solution for the (real-valued) Dirichlet problem is given by Γ(x − y) =( 1 2π log x − y, d = 2 d(2−d)B(0,1) x − y2−d, d > 2. 1 The Newtonian Potential of g ∈ L∞(Ω, X) is teh function defined via w := Γ ∗ g. Analogously to the real-valued case, see e.g. [GT01, Lemmas 4.1 & 4.2], we obtain Lemma 3.2. Let Ω ⊂ Rd be open and bounded. Let g ∈ C(Ω, X) be bounded and locally Holder continuous and let w be its Newtonian potential. Then w ∈ C 2(Ω, X) ∩ C(Ω, X) with ∆w = g. Assume now that g is bounded and locally Holder continuous. Using this lemma we immediately see that the Poisson Problem for L = ∆ is equivalent to the problem (∆v = 0 v = f − w on ∂Ω in Ω where v := u−w. Taking v as the Perron solution of the above Dirichlet Problem and adding w we obtain Theorem 3.3. Let Ω ⊂ Rd be open and bounded. For every bounded and locally Holder continuous function g ∈ C(Ω, X) and f ∈ C(∂Ω, X) there exists a unique function uf,g ∈ C 2(Ω, X) satisfying ∆uf,g = g lim x→z uf,g(x) = f (z) for all regular z ∈ ∂Ω referred to as the Perron solution of the Poisson Problem with L = ∆. In particular: The Poisson Problem for L = ∆ and any fixed g as above has a classical solution for every boundary data f ∈ C(∂Ω, X) if and only if every z ∈ ∂Ω is regular. Note that if Ω has a C 2,α-alpha boundary, then every z ∈ ∂Ω is re- gular, see e.g. [GT01, Problem 2.11]. PERRON SOLUTIONS 16 We now want to apply Schauder's Continuity Method to derive The- orem 3.1 from Theorem 3.3. The necessary Schauder Estimates will be derived from the real-valued case rather than proven in detail all over. Theorem 3.4. Under the assumptions of Theorem 3.1 a classical solu- tion u of the Poisson problem, if it exists, is an element of C 2,α(Ω, X) and there exists a constant C -- dependent on Ω, λ, α and the C α-norms of the coefficients of L -- such that kuk2,α ≤ C(kukC(Ω) + kF k2,α + kgkα). Further we have the maximum principle kukC(Ω,X) ≤ kf kC(∂Ω,X) + CkgkC(Ω,X). In particular this holds if L = ∆. Proof. The case X = R is well known, see e.g. [GT01, Theorems 6.6 and 6.14, Corollary 3.8]. Now consider a classical solution u of the Poisson Problem. For every x′ ∈ X ′ the function v := hu, x′i is a solution the the corresponding real-valued Poisson Problem (Lv = hg, x′i v = hf, x′i in Ω on ∂Ω and hence enjoys the claimed properties. Taking the supremum over all normed x′ it follows that kDiju(ξ) − Diju(η)k ξ − ηα ≤ C(kukC(Ω,X) + kf k2,α + kgkα) hence Diju ∈ C α(Ω, X). For the first claim it remains to show that the first derivatives Diu are continuous up to the boundary. The Uniform Boundedness Principle ensures that they are bounded. It follows that Diu ∈ W 1,p(Ω, X) for every p ≥ 1 and hence they are uniformly Holder continuous by Morrey's Embedding Theorem, see [AK18, Theorem 5.2]. We now show the estimate for the C 2,α-norm of u. Choose ξβ ∈ Ω and x′ β ∈ BX ′(0, 1) such that kDβukC(Ω,X) = hDβu(ξβ), x′ βi PERRON SOLUTIONS 17 holds for every multi-index β with β = 1, 2. Using these and the above estimate we compute kDβu(ξ) − Dβu(η)k ξ − ηα sup ξ,η∈Ω kDβukC(Ω,X) + Xβ=2 kDβhu, x′ βikC(Ω,R) kuk2,α =kukC(Ω,X) + Xβ=1,2 =kukC(Ω,X) + Xβ=1,2 + Xβ=2 ≤kukC(Ω,X) + Xβ=1,2 + Xβ=2 sup ξ,η∈Ω sup ξ,η∈Ω kDβu(ξ) − Dβu(η)k ξ − ηα khu, x′ βik2,α kDβu(ξ) − Dβu(η)k ξ − ηα ≤C(kukC(Ω,X) + kf k2,α + kgkα) where C is a multiple of the original C only dependant on the dimension of Ω. The last estimate follows similarly from the real case. (cid:3) Proof of Theorem 3.1. Uniqueness follows from the maximum princi- ple. In order to show existence it suffices to consider the case f = 0 since the Poisson Problem is equivalent to (Lv = h in Ω v = 0 on ∂Ω with v ∈ C 2,α(Ω, X) and h ∈ C α(Ω, X). In fact, choose h = g − LF , then u = v + F is the solution to the original problem. Let L0 := ∆ and L1 := L. Then the operator Lt := (1 − t)L0 + tL1 satisfies the same requirements as L where the constants and the C α-bounds for the coefficients can be chosen independently of t. We have that Lt ∈ L({u ∈ C 2,α(Ω, X), u∂Ω = 0}, C α(Ω, X)). Now if Ltut = f , then the estimates in Theorem 3.4 show that the Continuity Method [GT01, Theorem 5.2] is applicable. Since Ω is regular, Theorem 3.3 shows that L0 is bijective. Hence L1 is bijective as well from which we obtain the result. (cid:3) For the remainder of this section we will take a quick look at the reg- ularity of the solutions. Analogously to the Schauder estimates, we will show that the regularity carries over from the real to the vector-valued case. The Banach spaces C k,α(Ω, X) (k ∈ N0) are defined analogously to the space C 2,α(Ω, X). We start with the interior regularity: Proposition 3.5. Let k ∈ N0 and 0 < α < 1. Suppose u ∈ C 2(Ω, X) satisfies Lu = g for some g ∈ C k,α(Ω, X). Suppose further that the coefficients of L are in C k,α(Ω, R) and that they are bounded in this PERRON SOLUTIONS 18 space. Then u ∈ C k+2,α(Ω0, X) for every Ω0 ⊂⊂ Ω1 ⊂⊂ Ω and we have the estimate (2) kukC k+2,α(Ω0,X) ≤ C(kukC(Ω1,X) + kgkC k,α(Ω1,X)), where C = C(L, Ω1, Ω0, α, d). In particular: If we replace C k,α by C ∞, then also u ∈ C ∞(Ω, X). Proof. The case X = R is well known, see [GT01, Theorem 6.17 and If h < dist(Ω1, ∂Ω), then the Problem 6.1]. Let ρh be a mollifier. functions uh := ρh ∗ u and gh := ρh ∗ g satisfy Luh = gh in Ω1. We have uh ∈ C ∞(Ω1, X) and hence we may compute the norm of uh in C k+2,α(Ω0, X). Proceeding as in the proof of Theorem 3.4 we obtain the estimate (2) for uh and gh. Since uh → u uniformly in Ω1 and gh → g in C k,α(Ω1, X) it follows that uh is Cauchy in C k+2,α(Ω0, X) and hence converges in this space. We have that uh → u in C 2,α(Ω0, X) and hence u ∈ C k+2,α(Ω0, X) and the estimate (2) follows as well. (cid:3) Corollary 3.6. In the setting of Theorem 3.1 assume that the boundary of Ω is of class C k+2,α. Assume further that g ∈ C k,α(Ω, X) and the coefficients of L are bounded in C k,α(Ω, R). Assume finally that F ∈ C k+2,α(Ω, X). Then the solution u to the Poisson problem is an element of C k+2,α(Ω, X) as well. In particular: If we replace C k,α and C k+2,α by C ∞ then u ∈ C ∞(Ω, X). Proof. Again the real-valued case is well known, see [GT01, Theorem 6.19]. Now let X be arbitrary. From Proposition 3.5 we have that u ∈ C k+2(Ω, X). If we can show that the (k + 2) − nd derivatives of u are Holder-continuous, the result follows easily. To see this, note that by the real-valued case we have that hu, x′i ∈ C k+2,α(Ω, R) for every x′ ∈ X ′. Let v be any (k + 2) − nd derivative of u. The quotient v(ξ)−v(η) is weakly bounded and hence also bounded for all ξ, η ∈ Ω. ξ−ηα Thus v ∈ C α(Ω, X). (cid:3) 4. Application: Heat Equation on Irregular Domains As a second application of our work we want to investigate the heat equation with Dirichlet boundary data d dt u = ∆u u(t, ·) ∈ C0(Ω, X) for all t > 0 u(0, ·) = u0.   This problem is not well-posed if the domain Ω is irregular for the Dirichlet problem, see [AB99]. Hence we want to impose boundary conditions which incorporate the nature of the underlying Dirichlet problem. We consider the problem PERRON SOLUTIONS 19 d dt u = ∆u u(t, ·) ∈ Cb(Ω, X) for all t > 0 limξ→z,ξ∈Ω u(t, ξ) = 0 for all regular z ∈ ∂Ω and all t > 0 u(0, ·) = u0. (3)   We will treat this problem as an abstract Cauchy problem. For an overview of holomorphic semigroups and their generators, we refer to [Are04, Chapter 2] and [ABHN11, Chapter 3.7]. Consider the Banach space Cb,0,reg(Ω, X) := {u ∈ Cb(Ω, X), lim ξ→z u(ξ) = 0 for all z ∈ ∂regΩ}. On Cb,0,reg(Ω, X) we consider the Perron-Dirichlet-Laplacian ∆PD given by D(∆PD) := {u ∈ Cb,0,reg(Ω, X), ∆u ∈ Cb,0,reg(Ω, X)} ∆PDu := ∆u, i.e. ∆PD is the distributional Laplacian with maximal domain in Cb,0,reg(Ω, X). The main theorem of this section is Theorem 4.1. The operator ∆PD is the generator of a bounded holo- morphic semigroup on Cb,0,reg(Ω, X). Corollary 4.2. The problem (3) has a unique mild solution for all u0 ∈ D(∆PD). We will need some preparations for the proof. Lemma 4.3. ∆PD is closed and 0 ∈ ρ(∆PD). Proof. One easily sees that ∆PD is a closed operator. From the unique- ness of the Perron-solution it follows that ∆PD is injective. It remains to show that ∆PD is surjective. Let v ∈ Cb,0,reg(Ω, X) and let v be its extension to Rd by 0. As in the scalar-valued case, see e.g. [DL90, II,§3] one sees that the Newtonian potential w of v is continuous and satisfies ∆w = v in the sense of distributions. Let ϕ := w∂Ω, then u := H−ϕ + w ∈ Cb,0,reg(Ω, X) satisfies ∆u = v. (cid:3) For the m-dissipativity of the Perron-Dirichlet-Laplacian, we will need a stronger maximum principle. Proposition 4.4. [AG01, Theorem 5.2.6 (i)] Let u ∈ Cb(Ω ∪ ∂regΩ, R) be subharmonic. Then supξ∈Ω u(ξ) = supz∈∂regΩ u(z). In particular if u is even harmonic, then inf z∈∂regΩ u(z) ≤ u(ξ) ≤ supz∈∂regΩ u(z). Corollary 4.5. Let u ∈ Cb(Ω ∪ ∂regΩ, R). Suppose that there exist λ, M ≥ 0 such that ∆u − λu ≥ 0 (in the sense of distributions) and u ≤ M on ∂regΩ. Then u ≤ M. PERRON SOLUTIONS 20 Proof. Let ω := {ξ ∈ Ω, u(ξ) ≥ 0} and define In ω we have v(ξ) :=(u(ξ), 0, if ξ ∈ ω otherwise. ∆u ≥ ∆u − λu ≥ 0, hence u is subharmonic in ω. Moreover, v is constant in (ωc)◦ and hence subharmonic. On the boundary ∂ω we have that v is constantly zero and hence v(z) ≤ −RB(z,r) v for all z ∈ ∂ω and r > 0 sufficiently small, showing that v is subharmonic in Ω. Note that v can be continuously extended on the regular boundary points of Ω and that it satisfies v ≤ M on ∂regΩ. Hence by Proposition 4.4, we have that u ≤ v ≤ M. (cid:3) Proposition 4.6. ∆PD is m-dissipative. Proof. Let t > 0, u ∈ D(∆PD) and define M := ku − t∆PDuk∞. For x′ ∈ BX ′(0, 1) let v := Rehu, x′i. Then (v − M) − t∆(v − M) ≤ 0 and v − M ≤ 0 on ∂regΩ. Corollary 4.5 implies that v ≤ M. It follows that kuk∞ ≤ ku − t∆PDuk∞, and hence ∆PD is dissipative. We know that 0 ∈ ρ(∆PD), thus ∆PD is m-dissipative. (cid:3) Proposition 4.7. The function z 7→ G(z) ∈ L(L∞(Rd, X)) where (G(z)f )(ξ) := (4πz)−d/2ZRd f (ψ) exp(−(ξ − ψ)2/4z) dψ for all f ∈ L∞(Rd, X), ξ ∈ Rd, is a bounded holomorphic semigroup on Σθ for every θ < π/2. Its generator is the distributional Laplacian with maximal domain in L∞(Rd, X), that is D(∆∞) := {f ∈ L∞(Rd, X), ∆f ∈ L∞(Rd, X)} ∆∞f := ∆f. [Are04, 2.4]. For the Proof. The case X = C is well known, see e.g. general case note that kG(·)k is obviously bounded by 1 and that test- ing with x′ ∈ X ′ shows that G is a semigroup. The functionals on L(L∞(Rd, x)) given by T 7→ hx′ ◦ T f, µi where f ∈ L∞(Rd, X), x′ ∈ X ′ and µ ∈ L∞(Rd, R)′ are separating. Since for scalar-valued functions the semigroup G is holomorphic, it follows that z 7→ hx′ ◦ G(z)f, µi = PERRON SOLUTIONS 21 hG(z)(x′ ◦ f ), µi is holomorphic. By [Are16, Theorem 2.1] the semi- group G is holomorphic on X-valued functions as well. To identify the generator we fix λ > 0 = ω(G), f ∈ L∞(Rd, X) and define R(λ)f :=Z ∞ 0 e−λtG(t)f dt. Note that h∆h, x′i = ∆hh, x′i holds for every h ∈ L1 loc(Rd, X) and every x′ ∈ X ′ and that closed operators commute with integration. Thus testing with x′ shows that (λ − ∆∞)R(λ)f = f and that λ − ∆∞ is injective. Since ∆∞ is obviously closed we obtain λ ∈ ρ(∆∞) and R(λ, ∆∞) = R(λ). This shows that ∆∞ is the generator of G(·). (cid:3) Proof of Theorem 4.1. Since ∆∞ generates a bounded holomorphic semi- group it follows that kR(λ, ∆∞)k ≤ M λ for some M ≥ 0 whenever Re λ > 0. Let Re λ > 0, then by m-dissipativity λ ∈ ρ(∆PD). Let f ∈ Cb,0,reg(Ω, X) and denote by f ∈ L∞(Rd, X) the extension of f by 0. Let g := R(λ, ∆∞) f and g := R(λ, ∆PD)f . As in the real- [DL90, II,§3], we see that g ∈ C(Rd, X) and hence valued case, c.f. v := g − g ∈ Cb(Ω ∪ ∂regΩ, X) satisfies λv − ∆v = 0 and v = −g on ∂regΩ. The Regular Maximum Principle shows that and hence kvk∞ ≤ sup ∂regΩ kgk ≤ M λ kf k∞ kgk∞ ≤ 2M λ kf k∞, which finishes the proof. (cid:3) Remark 4.8. (a) If Ω is a regular domain, then Cb,0,reg(Ω, X) is noth- ing but C0(Ω, X). It follows from the density of the test functions that the semigroup generated by ∆PD is strongly continuous. (b) In general, the semigroup generated by ∆PD is not strongly con- Indeed, consider d = 2, Ω = B(0, 1)\{0} and X = R. tinuous. We show that every u ∈ D(∆PD) can be continuously extended to B(0, 1) which implies that D(∆PD) is not dense in Cb,0,reg(Ω, R). To see this consider a function u ∈ D(∆PD). We may extend u to R2 by 0 outside of Ω and consider the tempered distribution Tu defined that the distribution ∆Tu − T∆u is a distribution of order at most via hTu, ϕi = RR2 uϕ and proceed analogously for ∆u. It follows 2 which is supported in {0}, i.e. ∆Tu − T∆u =Pα≤2 aα∂αδ0. Let v and w be the solutions to ∆v = T∆u and ∆w = Pα≤2 aα∂αδ0, then Tu = v + w up to a perturbation by a harmonic function. By elliptic regularity we find that v is continuous even in 0. Further PERRON SOLUTIONS 22 w -- up to a perturbation by an analytic function -- is given by w(x, y) = a0 log r + a1 + a11 r2 − 2x2 r4 x r2 + a2 + a22 y r2 r2 − 2y2 r4 + a12 −2xy r4 , where r =px2 + y2. It is easy to see, that either w is unbounded or w = 0. Since u is bounded, we obtain that the latter is true and hence u can be continuously extended in {0}. Appendix A. The vector lattice Hb An ordered vector space which has the interpolation property is au- tomatically an order complete vector lattice, i.e. every set bounded from above has a supremum. To see this let A be a set bounded from above and B be the set of its upper bounds. Since a ≤ b for all a ∈ A, b ∈ B the interpolation property yields an element w satisfying a ≤ w ≤ b for all a ∈ A, b ∈ B, i.e. w is the least upper bound of A. In particular in the setting of Lemma 2.16 the vector space Hb(Ω, X) is an order complete vector lattice with respect to the order on C(Ω, X). Note however that it is not a sublattice of C(Ω, X) since the pointwise supremum of two harmonic functions is not harmonic in general. For two vectors u, v ∈ Hb(Ω, R) we denote their supremum in the vector lattice Hb(Ω, R) by u ∨H v. In the case X = R, the fact that Hb(Ω, R) is an order complete vector lattice can also be found in a more general form in [Bre67, Part IV, Theorem 11]. Moreover: Theorem A.1. Hb(Ω, R) is an order complete Banach lattice. For the proof we will need Lemma A.2. Let u, v ∈ Hb(Ω, R). Choose ωn ⊂⊂ ωn+1 ⊂⊂ Ω such that Sn ωn = Ω and such that ωn is regular for each n ∈ N. For all ξ ∈ ∂ωn set fn(ξ) := u(ξ) ∨R v(ξ), where u(ξ)∨Rv(ξ) denotes the supremum in R. Let hn be the solution to the Dirichlet problem in ωn with boundary data fn. Then hn → u ∨H v uniformly on compact sets. Proof. The fact that Hb(Ω, R) is an order complete vector lattice was discussed in the introduction of the appendix. It remains to show that hn converges to u ∨H v. First let K ⊂ Ω be compact. We may without loss of generality assume that K ⊂ ω1. For n ∈ N and ξ ∈ ∂ωn we have hn(ξ) = u(ξ) ∨R v(ξ) ≤ (u ∨H v) (ξ). Note that the functions on the left hand side and on the right hand side are harmonic while the one in the middle is subharmonic. The PERRON SOLUTIONS 23 maximum principle implies that u(ξ) ∨R v(ξ) ≤ hn(ξ) ≤ (u ∨H v) (ξ) for all ξ ∈ ωn. In particular this holds for ξ ∈ K. The Poisson inte- gral formula implies that hn is equicontinuous, hence the Arzela-Ascoli Theorem yields a subsequence of hn which converges uniformly on K to a function h ∈ Hb(K ◦, R). Next take Kn := ωn then using a diagonal argument we find a subse- quence of hn which converges to h ∈ Hb(Ω, R) uniformly on every Kn. The above inequality shows that h = u ∨H v. A subsequence argument shows that hn → u ∨H v. (cid:3) One can easily use Lemma A.2 to show that this construction works analogously for other lattice operations: Corollary A.3. In the formulation of Lemma A.2 choose fn(ξ) := u(ξ)R, where u(ξ)R denotes the absolute value in R. Then hn → uH uni- formly on compact sets, where uH denotes the absolute value in the lattice Hb(Ω, R). Proof of Theorem A.1. It remains to show that k uH k∞ = kuk∞. The ≥" is immediate. On the other hand if hn is chosen as in estimate " Corollary A.3, then the maximum principle implies that 0 ≤ hn(ξ) ≤ max ξ∈∂ωn u(ξ)R ≤ kuk∞. Since hn(ξ) converges to u(ξ)H the estimate ≤" follows. " (cid:3) References [AB99] Wolfgang Arendt and Philippe B´enilan, Wiener regularity and heat semigroups on spaces of continuous functions, Topics in nonlinear anal- ysis, Progr. Nonlinear Differential Equations Appl., vol. 35, Birkhauser, Basel, 1999, pp. 29 -- 49. MR 1724790 Charalambos D. Aliprantis and Kim C. Border, Infinite dimensional analysis, third ed., Springer, Berlin, 2006, A hitchhiker's guide. MR 2378491 [AB06a] [AB06b] Charalambos D. Aliprantis and Owen Burkinshaw, Positive operators, Springer, Dordrecht, 2006, Reprint of the 1985 original. MR 2262133 [ABHN11] Wolfgang Arendt, Charles J. K. Batty, Matthias Hieber, and Frank Neubrander, Vector-valued Laplace transforms and Cauchy problems, second ed., Monographs in Mathematics, vol. 96, Birkhauser/Springer Basel AG, Basel, 2011. MR 2798103 Sheldon Axler, Paul Bourdon, and Wade Ramey, Harmonic function theory, second ed., Graduate Texts in Mathematics, vol. 137, Springer- Verlag, New York, 2001. MR 1805196 [ABR01] [AD08] Wolfgang Arendt and Daniel Daners, The Dirichlet problem by vari- ational methods, Bull. Lond. Math. Soc. 40 (2008), no. 1, 51 -- 56. MR 2409177 PERRON SOLUTIONS 24 [AG01] David H. Armitage and Stephen J. Gardiner, Classical potential theory, Springer Monographs in Mathematics, Springer-Verlag London, Ltd., London, 2001. MR 1801253 [AK18] Wolfgang Arendt and Marcel Kreuter, Mapping theorems for Sobolev spaces of vector-valued functions, Studia Math. 240 (2018), no. 3, 275 -- 299. MR 3731026 [Are04] Wolfgang Arendt, Semigroups and evolution equations: functional cal- culus, regularity and kernel estimates, Evolutionary equations. Vol. I, Handb. Differ. Equ., North-Holland, Amsterdam, 2004, pp. 1 -- 85. MR 2103696 [Are16] [Bre61] [Bre67] [Bre11] [DL90] [Dug51] [GT01] [Hel09] [Hil05] [Kel66] [Lan72] , Vector-valued holomorphic and harmonic functions, Concr. Oper. 3 (2016), 68 -- 76. MR 3499920 Marcel Brelot, Sur un th´eor`eme de prolongement fonctionnel de Keldych concernant le probl`eme de Dirichlet, J. Analyse Math. 8 (1960/1961), 273 -- 288. MR 0125245 , Lectures on potential theory, Notes by K. N. Gowrisankaran and M. K. Venkatesha Murthy. Second edition, revised and enlarged with the help of S. Ramaswamy. Tata Institute of Fundamental Research Lec- tures on Mathematics, No. 19, Tata Institute of Fundamental Research, Bombay, 1967. MR 0259146 Haim Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011. MR 2759829 Robert Dautray and Jacques-Louis Lions, Mathematical analysis and numerical methods for science and technology. Vol. 1, Springer-Verlag, Berlin, 1990, Physical origins and classical methods, With the col- laboration of Philippe B´enilan, Michel Cessenat, Andr´e Gervat, Alain Kavenoky and H´el`ene Lanchon, Translated from the French by Ian N. Sneddon, With a preface by Jean Teillac. MR 1036731 J. Dugundji, An extension of Tietze's theorem, Pacific J. Math. 1 (1951), 353 -- 367. MR 0044116 David Gilbarg and Neil S. Trudinger, Elliptic partial differential equa- tions of second order, Classics in Mathematics, Springer-Verlag, Berlin, 2001, Reprint of the 1998 edition. MR 1814364 Lester L. Helms, Potential theory, Universitext, Springer-Verlag Lon- don, Ltd., London, 2009. MR 2526019 Stefan Hildebrandt, On Dirichlet's principle and Poincar´e's m´ethode de balayage, Math. Nachr. 278 (2005), no. 1-2, 141 -- 144. MR 2111805 Mstislaw V. Keldys, On the solvability and the stability of the dirichlet problem, Amer. Math. Soc. Translations(2) 51 (1966), 1 -- 73. Naum S. Landkof, Foundations of modern potential theory, Springer- Verlag, New York-Heidelberg, 1972, Translated from the Russian by A. P. Doohovskoy, Die Grundlehren der mathematischen Wissenschaften, Band 180. MR 0350027 Marcel Kreuter, Institute of Applied Analysis, Ulm University, 89069 Ulm, Germany E-mail address: [email protected]
1603.06198
2
1603
2016-04-17T11:43:51
A Selection Theorem for Banach Bundles and Applications
[ "math.FA" ]
It is shown that certain lower semi-continuous maps from a paracompact space to the family of closed subsets of the bundle space of a Banach bundle admit continuous selections. This generalization of the theorem of Douady, dal Soglio-Herault, and Hofmann on the fullness of Banach bundles has applications to establishing conditions under which the induced maps between the spaces of sections of Banach bundles are onto. Another application is to a generalization of the theorem of Bartle and Graves for Banach bundle maps that are onto their images. Other applications of the selection theorem are to the study begun by Behrends and continued by Gierz of the M-ideals of the space of bounded sections. A class of Banach bundles that generalizes the class of locally trivial bundles is introduced and some properties of the Banach bundles in this class are discussed.
math.FA
math
A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS ALDO J. LAZAR Abstract. It is shown that certain lower semi-continuous maps from a para- compact space to the family of closed subsets of the bundle space of a Banach bundle admit continuous selections. This generalization of the the theorem of Douady, dal Soglio-Herault, and Hofmann on the fullness of Banach bundles has applications to establishing conditions under which the induced maps be- tween the spaces of sections of Banach bundles are onto. Another application is to a generalization of the theorem of Bartle and Graves [3] for Banach bundle maps that are onto their images. Other applications of the selection theorem are to the study of the M-ideals of the space of bounded sections begun in [4] and continued in [9]. A class of Banach bundles that generalizes the class of trivial Banach bundles is introduced and some properties of these Banach bundles are discussed. Date: April 17, 2016. 1991 Mathematics Subject Classification. Primary 46B20; Secondary 55R65, 58B05. Key words and phrases. Banach bundles, Banach bundle maps, spaces of sections of Banach bundles, lower semi-continuous set-valued maps. 1 2 ALDO J. LAZAR 1. Introduction and basic definitions and notations The theory of Banach bundles, sometimes presented as continuous fields of Ba- nach spaces, helped to advance the study of unitary representations of locally com- pact groups and of C∗-algebras. An early example of its usefulness and interest in these fields can be found in [7] and for later developments we mention [8]. Other beneficial connections of this theory are with the theory of Banach lattices. A short exposition on this aspect is given in the comprehensive monograph of Gierz [9] and more recent results appeared in [12]. The theory of Banach bundles provides an association of topology, the geometry of Banach spaces and operator theory, as exemplified in [9, Chapters 15 and 16] and [13]. We prove in this paper a generalization of Michael's well known selection theo- rem, [17, Theorem 3.2"], in the context of Banach bundles. This result, Theorem 2.9, also generalizes the theorems of Douady, L. dal Soglio-Herault, and K. H. Hof- mann, see [8, Appendix C], on the existence of sufficiently many continuous cross- sections in a Banach bundle. A particular case of Theorem 2.9 is [9, Proposition 15.13]. In Section 3 we consider Banach bundle maps (definition will follow) and the induced maps between the spaces of sections. We apply the selection theorem to establishing some conditions under which these later maps are onto their images. This investigation led to examining when a map between two Banach bundles is open. The section ends with the generalization of a theorem of Bartle and Graves on the existence of continuous right inverses for maps between Banach spaces. Here the context is of course maps between Banach bundles and we follow Michael [17] in using the selection theorem for deriving the existence of such an inverse. Alfsen and Effros developed in [1] a structure theory for Banach spaces in which certain subspaces named M-ideals play a crucial role. The M-ideals of the Banach space of sections of a Banach bundle whose base space is compact Hausdorff were investigated in depth by Behrends [4] and Gierz [9]. In Section 4 we extend this investigation to the case when the base space is locally compact Hausdorff and the sections vanish at infinity. A certain class of continuous Banach bundles is introduced in Section 5; the bundles in this class are called locally uniform. Every locally trivial Banach bundle A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 3 is locally uniform but the converse is false. The locally uniform Banach bundles enjoy some nice properties. For instance we show in Proposition 5.7 that the quo- tient of a locally uniform Banach bundle by a locally uniform Banach subbundle is a continuous Banach bundle. Some topological properties of the bundle space are treated in the Appendix. Conditions which insure that the bundle space is a Baire space are given. Also its paracompactness and its metrizability are discussed there. Throughout of this paper the field of scalars is denoted by K; it can be the field of the real numbers or the field of the complex numbers. A paracompact space is always considered to be Hausdorff. A topological space is locally paracompact if it has a cover with open subsets whose closures are paracompact. A map Φ from a topological space T1 to the family of subsets of a topological space T2 is called lower semi-continuous if for every open non-void subset O of T2 the set {t ∈ T1 O ∩ Φ(t) 6= ∅} is open in T1. For a Banach space X, x ∈ X, and r a positive number we denote B(x, r) := {y ∈ X ky − xk < r} and by Xr the closed ball of X whose center is at the origin with radius r. A closed subspace Y of the Banach space X is called an M-ideal of X if its polar Z1 in the dual of X has a closed complement Z2 such that kz1 + z2k = kz1k + kz2k for every zi ∈ Zi, i = 1, 2. A closed subspace Y of X is an M-ideal if and only if ithas the 3-ball property: if ∩3 i=1B(xi, ri) 6= ∅ and B(xi, ri) ∩ Y 6= ∅, 1 ≤ i ≤ 3, then Y ∩ ∩3 i=1B(xi, ri) 6= ∅. By a Banach bundle ξ := (E, p, T ) we shall mean what in [5, pp. 8,9] is called an (H)Banach bundle. That is x → kxk is upper semi-continuous on the bundle space E, see also [9, p.21]. We shall always suppose that the base space T is Hausdorff. The fiber p−1(t) over t ∈ T will be sometimes denoted E(t). The origin of the Banach space E(t) is denoted 0t but if there can be no confusion just 0 will be used. The bundle ξ is called a continuous Banach bundle if x → kxk is continuous on E. The bundle space of a continuous Banach bundle is always Hausdorff, see [9, Proposition 16.4]. If E ′ is a subset of E such that p E ′ is onto T and open and each p−1(t) ∩ E ′ is a closed subspace of E(t) then ξ′ := (E ′, p E ′ , T ) is a Banach bundle called a Banach subbundle of ξ. One can define an equivalence relation on E as follows: x1, x2 are equivalent if p(x1) = p(x2) and x1 − x2 belongs to 4 ALDO J. LAZAR p−1(p(x1)). The quotient space, denoted E/E ′, is the union of all the quotient Banach spaces E(t)/E ′(t), t ∈ T . If we denote by p the obvious map of E/E ′ onto T then η := (E/E ′, p, T ) is a Banach bundle by [9, Chapter 9]. The quotient map q : E → E/E ′ is open, see [9, 9.4]. Let ξ := (E, p, T ) be a Banach bundle. A continuous function f : T → E is called a section of ξ if p(f (t)) = t for every t ∈ T . The linear space of all the sections of ξ is denoted Γ(ξ); its subspace of all the bounded sections is denoted Γb(ξ). This is a Banach space, the norm of f ∈ Γb(ξ) being kf k := supt∈T kf (t)k. If T is locally compact Hausdorff then the subspace Γ0(ξ) of Γb(ξ) consisting of all the sections that vanish at infinity is of interest. A full Banach bundle is a bundle such that for each x ∈ E there exists a section f ∈ Γ(ξ) satisfying f (p(x)) = x. In this definition one can replace Γ(ξ) by Γb(ξ) as observed in [5, p. 14] and even by Γ0(ξ) if T is locally compact. For f ∈ Γ(ξ), V an open subset of T , and a > 0 the open subset U (f, V, a) := {x ∈ E p(x) ∈ V, kx − f (p(x))k < a} of E is called a tube. If ξ is a full Banach bundle then the family of all the tubes forms a base for the topology of E by [5, p. 10]. A Banach bundle map from ξ1 := (E1, p1, T ) to ξ2 := (E2, p2, T ) is a continuous map ϕ : E1 → E2 such that ϕ(p−1 of ϕ to each fiber p−1 2 (t) for every t ∈ T and the restriction 1 (t) is linear. The map ϕ is called an (isometric) isomorphism 1 (t)) ⊂ p−1 if ϕ(E1) = E2 and ϕ is an isometry on every fiber. 2. The selection theorem In this section we shall state and prove the selection theorem for set valued maps into a Banach bundle and we shall discuss some of its applications. The proof mimics closely the proofs given in [8, Appendix C] and [17]. A sequence of lemmas prepares the proof of the theorem itself. From now on in this section ξ := (E, p, T ) is a fixed Banach bundle. Following [8, Appendix C] we shall say that a subset U of E is ε-thin, ε > 0, if kx − x′k < ε whenever x, x′ ∈ E and p(x) = p(x′). Lemma 2.1. Let t ∈ T , x1, x2 ∈ p−1(t) with kx1 − x2k < ε. There is an open ε-thin set U that contains the segment [x1, x2]. A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 5 Proof. With D := {(x, y) ∈ E × E p(x) = p(y)}, the set {(x, y) kx − yk < ε} is open since (x, y) → kx − yk is upper semi-continuous. There is an open subset O of E × E such that [x1, x2] × [x1, x2] ⊂ {(x, y) kx − yk} = O ∩ D. By [18, p. 171] there is an open set U ⊂ E such that [x1, x2] × [x1, x2] ⊂ U × U ⊂ O. Obviously U is ε-thin. (cid:3) We enlarge now our setting with a completely regular topological space S and a continuous open map π of S onto T . A function f : S → E is called admissible if p ◦ f = π. Let ε > 0. A function f : S → E is called ε-continuous at s ∈ S if there are a neighbourhood V of s and an ε-thin neighbourhood U of f (s) such that f (V ) ⊂ U . If f is ε-continuous at all the points of S then it is called shortly ε-continuous. Lemma 2.2. If an admissible function f is ε-continuous for all ε > 0 then it is continuous. Proof. Let s ∈ S and and suppose that Vn is a neighbourhood of s and Un is an 1/n-thin neighbourhood of f (s), with f (Vn) ⊂ Un, n ∈ N. We have f (Vn ∩ Vn+1) ⊂ Un ∩ Un+1 so there is no loss of generality if we suppose that {Un} is decreasing. Let now W be a neighbourhood of f (s). By [5, p. 10] there are a natural number n and a neighbourhood V of π(s) such that Un ∩ p−1(V ) ⊂ W . The neighbourhood V ∩ π(Vn) of p(f (s)) = π(s) satisfies Un ∩ p−1(V ∩ π(Vn)) ⊂ W . The neighbourhood Vn ∩ π−1(V ) os s satisfies f (Vn ∩ π−1(V )) ⊂ W . Indeed, if s′ ∈ Vn ∩ π−1(V ) ⊂ Vn then f (s′) ∈ Un. Moreover, p(f (s−1)) = π(s′) ∈ π(Vn) ∩ V ; hence f (s′) ∈ p−1(π(Vn) ∩ V ) ⊂ p−1(V ). Thus f (s′) ∈ Un ∩ p−1(V ) ⊂ W and the continuity of f at s is established. (cid:3) Lemma 2.3. Let G be an open subset of S and {Ui}n that p(Ui) = π(G), 1 ≤ i ≤ n. Suppose that {ϕi}n i=1 be open subsets of E such i=1 are continuous scalar functions 6 ALDO J. LAZAR on G with ϕ1(s) 6= 0 for each s ∈ G. Then the set W := {X ϕi(s)xi s ∈ G, xi ∈ Ui, p(xi) = π(s), 1 ≤ i ≤ n} is open in E. Proof. Consider the set Z := {(s, x1 . . . , xn) s ∈ G, xi ∈ E, p(xi) = π(s), 1 ≤ i ≤ n} endowed with the relative topology inherited from G × p−1(π(G))n. A basis for this topology consists of all the sets of the form O := (V × V1 × . . . Vn) ∩ Z where V ⊂ G, Vi ⊂ p−1(π(G)), 1 ≤ i ≤ n, are open. Thus (1) O := {(s, x1, . . . xn) s ∈ V, xi ∈ Vi, p(xi) = π(s), 1 ≤ i ≤ n}. is a homeomorphism. Indeed it is continuous and its inverse maps (s, y1, . . . yn) The map α of Z onto itself given by α(s, x1, . . . xn) := (s,P ϕi(s)xi, x2, . . . xn) to (s, 1/ϕ1(s)(y1 − Pn i=2 ϕi(s)yi), y2, . . . yn). The map β of Z into E given by β(s, x1, . . . xn) := x1 is open. Indeed, if O is as in (1) then β(O) = V1 ∩ p−1(π(V ) ∩ ∩n i=1p(Vi)) that is open. Now W = β(α((G × U1, . . . Un) ∩ Z)) is obviously open. (cid:3) Lemma 2.4. Suppose ϕi, 1 ≤ i ≤ n, are continuous functions from S to [0, 1] with P ϕi = 1. Let ε > 0. If fi, 1 ≤ i ≤ n are admissible and ε-continuous at s0 ∈ S then f :=P ϕifi is admissible and ε-continuous at s0. Proof. We have to prove only that f is ε-continuous at s0. There is no loss of generality if we suppose ϕ1(s0) 6= 0. There exist ε-thin open neighbourhoods Ui of fi(s0) and open neighbourhoods Vi of s0 such that fi(Vi) ⊂ Ui, 1 ≤ i ≤ n. By taking ∩n i=1Vi and further reducing it if needed one gets an open neighbourhood V of s0 such that fi(V ) ⊂ Ui, 1 ≤ i ≤ n, and ϕ(s) > 0 if s ∈ V . From fi(V ) ⊂ Ui and the admissibility of fi we get fi(V ) ⊂ p−1(π(V ))∩Ui and p(p−1(π(V ))∩Ui) = π(V ), 1 ≤ i ≤ n. The set p−1(π(V )) ∩ Ui is an open neighbourhood of fi(s0). The set W := {X ϕi(s)xi s ∈ V, xi ∈ p−1(π(V )) ∩ Ui, p(xi) = π(s), 1 ≤ i ≤ n} is open in E by Lemma 2.3 and f (s0) ∈ W . By the above f (V ) ⊂ W and it remains to show that W is ε-thin. To this end, let x′, x′′ ∈ W with p(x′) = p(x′′) = π(s) for A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 7 some s ∈ V . Then x′ =P ϕi(s)x′ i ∈ Ui, p(x′′ with x′′ i i ) = π(s), 1 ≤ i ≤ n. Each Ui is ε-thin so kx′ − x′′k < ε and we i) = π(s) and x′′ =P ϕi(s)x′′ i ∈ Ui, p(x′ i with x′ conclude that W is ε-thin. (cid:3) Lemma 2.5. Let x0 ∈ E, s0 ∈ S with π(s0) = p(x0). There exists an ε-continuous admissible function f : S → E such that f (s0) = x0. Proof. Let U be an ε-thin open neighbourhood of x0 and set V := π−1(p(U )) which is an open neighbourhood of s0. Define g : V → E so that g(s0) = x0 and g(s) ∈ U with p(g(s)) = π(s) for s ∈ V . Let W be an open neighbourhood of s0 such that W ⊂ V and ϕ : S → [0, 1] a continuous function such that ϕ(s0) = 1 and ϕ S\W ≡ 0. Set (2) f (s) :=(ϕ(s)g(s), 0, if s ∈ V , if s /∈ V . Then on V , f = ϕg + (1 − ϕ)0. Thus f is ε-continuous on V by Lemma 2.4 and trivially ε-continuous on S \ W . Clearly f is an admissible function. (cid:3) From here on, until the end of the proof of Theorem 2.9, S is a paracompact space and Φ is a lower semi-continuous map from S to the family of non-void closed subsets of E such that Φ(s) is a convex subset of p−1(π(s)) for every s ∈ S. Lemma 2.6. Given s0 ∈ S and ε > 0 there are an ε-continuous admissible function f and a neighbourhood V of s0 such that B(f (s), ε) ∩ Φ(s) 6= ∅ for every s ∈ S. Proof. Let x0 ∈ Φ(s0) and f be an ε-continuous admissible function such that f (s0) = x0 as given by Lemma 2.5. There are an ε-thin neighbourhood U of x0 and a neighbourhood V1 of s0 such that f (V1) ⊂ U . The set V2 := {s ∈ S U ∩Φ(s) 6= ∅} is an open neighbourhood of s0. Set V := V1 ∩ V2. Then, for each s ∈ V there is x ∈ U ∩ Φ(s) and kf (s) − xk < ε since f (s) ∈ U and U is an ε-thin set. (cid:3) Now we have reached the main approximation lemma. Its statement consists of two assertions with similar proofs. Lemma 2.7. Let ε > 0. 8 ALDO J. LAZAR (i) There is an ε-continuous admissible function g such that B(g(s), ε) ∩ Φ(s) 6= ∅ for every s ∈ S. (ii) Given an ε-continuous admissible function f such that B(f (s), ε)∩Φ(s) 6= ∅ for every s ∈ S there is an ε/2-continuous admissible function g such that kf (s) − g(s)k < 2ε and B(g(s), ε/2) ∩ Φ(s) 6= ∅ for every s ∈ S. Proof. We begin with the proof of (ii). Given f as above, the first step of the proof is to show that for every s ∈ S there are an open neighbourhood V and an ε/2-continuous admissible function h such that B(h(s′), ε/2) ∩ Φ(s′) 6= ∅ and kf (s′) − h(s′)k < 2ε for each s′ ∈ V . Let s ∈ S and choose x ∈ B(f (s), ε) ∩ Φ(s). There is an ε-thin open neighbourhood U1 of f (s) and an open neighbourhood V1 of s such that f (V1) ⊂ U1. Let y = 3/4x + 1/4f (s). Lemma 2.5 yields an ε/4- continuous admissible function h such that h(s) = y. We are going to show that h satisfies our claim. There are an ε/4-thin open neighbourhood U2 of y and an open neighbourhood V2 of s such that h(V2) ⊂ U2. By using Lemma 2.1 we get an ε/4-thin open set U3 that contains the segment [y, x] and an open 3ε/4-thin set U4 that contains the segment [y, f (s)]. Denote V ′ := {s′ U3 ∩ Φ(s′) 6= ∅}, an open neighbourhood of s since x ∈ U3 ∩ Φ(s). Set now V := π−1(p(U2 ∩ U3 ∩ U4)) ∩ π−1(p(U1 ∩ U4)) ∩ V1 ∩ V2 ∩ V ′. We have s ∈ π−1(p(U2 ∩ U3 ∩ U4)) since y ∈ U2 ∩ U3 ∩ U4 and p(y) = p(f (s)) = π(s). Moreover, f (s) ∈ U1 ∩ U4 hence s ∈ π−1(p(U1 ∩ U4)). We conclude that V is an open neighbourhood of s. If s′ ∈ V ⊂ V1 then f (s′) ∈ U1 and π(s′) ∈ p(U1 ∩ U4). Thus there is x1 ∈ U1 ∩ U4 with p(x1) = π(s′) and (3) kf (s′) − x1k < ε since U1 is ε-thin. For s′ ∈ V we have π(s′) ∈ p(U2 ∩ U3 ∩ U4) hence there is x2 ∈ U2 ∩ U3 ∩ U4 such that p(x2) = π(s′). From s′ ∈ V ⊂ V2 we get h(s′) ∈ U2 and (4) kh(s′) − x2k < ε/4. The set U4 is 3/4ε-thin so (5) kx1 − x2k < 3/4ε. A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 9 Finally, if s′ ∈ V ⊂ V ′ then there is x3 ∈ U3 ∩ Φ(s′) and we have (6) kx2 − x3k < ε/4 since U3 is ε/4-thin. From (4) and (6) we get kh(s′) − x3k ≤ kh(s′) − x2k + kx2 − x3k < ε/2. Therefore, if s′ ∈ V then B(h(s′), ε/2) ∩ Φ(s′) 6= ∅. Also, if s′ ∈ V we get from (4), (5) and (3) that kh(s′) − f (s′)k ≤ kh(s′) − x2k + kx2 − x1k + kx1 − f (s′)k < ε/4 + 3ε/4 + ε = 2ε and the claim is established. As mentioned above, what has been done up to this point is irrelevant for (i). To obtain the proof of (i) from what follows one has to ignore every affirmation about the function f . We know now that there exist an open covering {Vα}α∈A of S and admissible ε/2-continuous functions hα : S → E such that B(hα(s), ε/2) ∩ Φ(s) 6= ∅ and kf (s) − hαk < 2ε for every s ∈ Vα. For proving (i) one obtains this from Lemma 2.6. We may and shall suppose that the covering of S is locally finite. Let {ϕα} be a partition of unity subordinated to {Vα}. and define g := Pα ϕαhα. Each point s ∈ S has a neighbourhood Os such that there exists a finite set Fs ⊂ A with the property that Os ∩ Vα = ∅ is α /∈ Fs. Thus g is well defined, admissible and ε/2-continuous by Lemma 2.4. We have B(g(s), ε/2) ∩ Φ(s) 6= ∅ for every s ∈ S. Indeed, choose zα ∈ B(hα(s), ε/2) ∩ Φ(s) for α ∈ Fs. Then z :=Pα∈Fs zα ∈ Φ(s) and kg(s) − zk ≤ Xα∈Fs ϕα(s)khα(s) − zαk < ε/2. Furthermore, kg(s)−f (s)k ≤ k Xα∈Fs ϕα(s)hα(s)−Xα∈Fs ϕα(s)f (s)k ≤ Xα∈Fs and the proof is complete. ϕα(s)khα(s)−f (s)k < 2ε (cid:3) Lemma 2.8. Let {fn} be a sequence of admissible ε-continuous (ε > 0) functions, uniformly convergent on S. Then its limit f is admissible and 2ε-continuous. 10 ALDO J. LAZAR Proof. Clearly f is admissible; we have only to prove that it is 2ε-continuous. Let n be such that kf (s) − fn(s)k < ε/2 for every s ∈ S. For s0 ∈ S there are a neighbouhood V of s0 and a neighbourhood U ′ of fn(s0) such that fn(V ) ⊂ U ′. Now, fn is admissible so π(V ) ⊂ U ′. Put U := U ′ ∩ p−1(π(V )) and W := {x + y p(x) = p(y) ∈ π(V ), kxk < ε/2, y ∈ U }. Then Lemma 2.3 yields that W is an open neighbourhoo of f (s0) and it is easily seen that it is 2ε-thin. If s ∈ V then f (s) = fn(s) + (f (s) − fn(s)) ∈ W . We conclude that f is 2ε-continuous at s0. (cid:3) In all the applications but one of the theorem that follows the setup will be somewhat simpler; we shall have S = T and π will be the identity map of T . Theorem 2.9. Let ξ := (p, E, T ) be a Banach bundle, S a paracompact topological space, π a continuous open map of S onto T and Φ a lower semi-continuous map from S to the family of non-empty closed subsets of E such that Φ(s) is a convex subset of the fiber p−1(π(s)) for every s ∈ S. Then there exists a continuous admissible function f : S → E such that f (s) ∈ Φ(s) for every s ∈ S. Moreover, if s0 ∈ S and x0 ∈ Φ(s0) then f can be chosen so that f (s0) = x0. Proof. By Lemma 2.7 (i) there exists an 1/2-continuous admissible function f1 such that B(f1(s), 1/2) ∩ Φ(s) 6= ∅ for every s ∈ S. We proceed now by induction. Suppose that admissible functions {fi}n i=1 are given such that each fi is 1/2i- continuous, satisfies B(fi(s), 1/2i) ∩ Φ(s) 6= ∅ and kfi(s) − fi+1(s)k < 1/2i−1 for every s ∈ S and 1 ≤ i ≤ n − 1. By Lemma 2.7 (ii) there exists an i/2n+1- continuous admissible function fn+1 such that B(fn+1(s), 1/2n+1) ∩ Φ(s) 6= ∅ and kfn(s) − fn+1(s)k < 1/2n−1 for every s ∈ S. The sequence {fn} is uniformly convergent on S. The limit f is admissible and 1/2n-continuous for n ≥ 1 by Lemma 2.8. Thus f is continuous on S by Lemma 2.2. Obviously f (s) ∈ Φ(s) for each s ∈ S. The second assertion of the theorem follows from the first by considering the lower semi-continuous set valued map (7) Φ(s) :=(Φ(s), s 6= s0, x0, s = s0. A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 11 (cid:3) Theorem 2.9 is a generalization of [9, Proposition 15.13] where the case S = T , a compact Hausdorff space, π being the identity map of T , is discussed. By considering a trivial Banach bundle one can deduce from Theorem 2.9 the existence of the selection part of Michael's [17, Theorem 3.2"]. Another known result that can be deduced from Theorem 2.9 is the observation on p. 15 of [5] that partial section over closed subsets of the base space can be extended to sections over the entire base space. We state below the precise result. Corollary 2.10. Let ξ := (p, E, T ) be a Banach bundle with paracompact base space, A a closed subset of T and ϕ : A → E a section of ξ A. Then there exists a section ϕ : T → E of ξ that extends ϕ. Moreover, if ϕ is bounded then there exists a bounded extension ϕ. If ξ := (E, p, T ) is a Banach bundle with locally compact Hausdorff base space, A a compact subset of T and ϕ a section of ξ A then there is a section ϕ ∈ Γ0(ξ) that extends ϕ. Proof. The first claim is obtained by applying Theorem 2.9 to the set valued map (8) Φ(t) :=({ϕ(t)}, t ∈ A, p−1(t), t ∈ T \ A. To get a bounded extension in case ϕ is bounded one multiplies a section that extends ϕ by a bounded continuous scalar valued function that is constantly equal to 1 on A. Finally, when T is locally compact Hausdorff and A is compact one extends ϕ to a section ϕ′ of ξ V , V being an open neighbourhood of A with compact closure. With f : T → [0, 1] a continuous function such that f A ≡ 1 and f T \ V ≡ 0 one defines (9) ϕ(t) :=(f (t)ϕ′(t), t ∈ V 0t, t ∈ T \ V. (cid:3) Of course, one can deduce from Corollary 2.10 the well known theorems of A. Douady, L. dal Soglio-Herault and K. H. Hofmann ([8, pp. 640-641]) on the fullness of Banach bundles with paracompact or locally compact Hausdorff base spaces. 12 ALDO J. LAZAR 3. Banach bundle maps We shall now discuss Banach bundle maps and the way they affect various spaces of sections. We shall end this chapter with an extension of a well known result of Bartle and Graves [3] to Banach bundle maps. Let ξ1 := (E1, p1, T ) and ξ2 := (E2, p2, T ) be Banach bundles. Recall that a Banach bundle map from ξ1 to ξ2 is a continuous map ϕ : E1 → E2 such that p2 ◦ ϕ = p1 and which acts linearly on each fiber of ξ1. The restriction of ϕ to p−1 1 (t), t ∈ T , is a bounded linear operator into p−1 2 (t) which we shall denote by ϕt. The Banach bundle map ϕ induces a linear map ϕ : Γ(ξ1) → Γ(ξ2) as follows: ϕ(f )(t) := ϕ(f (t)), f ∈ Γ(ξ1), t ∈ T. This map ϕ need not be onto Γ(ξ2) even if ϕ is onto E2. To illustrate this we shall present an example inspired by one in [2]. In order to have ϕ(Γ(ξ1) = Γ(ξ2) one has to require ϕ to be open. Example 3.1. Let E1 := [0, 1] × (K × K), E2 := ({0} × ({0} × K)) ∪ ((0, 1] × (K × K))K)) and p be the projection of E1 onto [0, 1]. We give the linear space {t}×(K×K) ⊂ E1 the norm k(t, (x, y))k := max(x, y). Then ξ1 := (E1, p, [0, 1]) is a trivial Banach bundle and ξ2 := (E2, p E2, [0, 1]) a Banach subbundle of ξ1. Define the Banach bundle map ϕ : E1 → E2 by ϕ(t, (x, y)) := (t, (tx, y)). Then ϕ(E1) = E2. obvious that the section f (t) := (t, (t1/2, 1)) of Γ(ξ2) has no preimage in Γ(ξ1). It is Theorem 3.2. Let ξi := (Ei, p1, T ), i = 1, 2, be Banach bundles with T paracompact and ϕ := E1 → E2 an open Banach bundle map onto E2. Then ϕ(Γ(ξ1)) = Γ(ξ2). Proof. Let f ∈ Γ(ξ2) and define Φ(t) := ϕ−1(f (t)), t ∈ T . We are going to show that Φ is lower semi-continuous. To this end let U ⊂ E1 be open and suppose U ∩ Φ(t0) 6= ∅. Thus there exists x0 ∈ U with p(x0) = t0 and ϕ(x0) = f (t0). The set ϕ(U ) is open in E2 therefore there exists an open neighbourhood V of t0 such that f (V ) ⊂ ϕ(U ). It follows that t0 ∈ V ⊂ {t ∈ T U ∩ Φ(t) 6= ∅} and we conclude that the latter set is open. By Theorem 2.9 Φ admits a continuous selection g. This is a section of ξ1 that satisfies ϕ(g) = f . (cid:3) A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 13 Thus it seems to be of some interest to establish conditions for a Banach bundle map which is onto the bundle space of the image to be open. Proposition 3.3. Let ξ1 := (E1, p1, T ) be a full Banach bundle and ϕ a Banach bundle map of ξ1 onto ξ2 := (E2, p2, T ). Then ϕ is open if and only if there exists an open cover {Vα}α∈A of T and positive numbers {mα}α∈A such that for every y ∈ p−1 2 (Vα) there is x ∈ E1 that satisfies ϕ(x) = y and kxk ≤ mαkyk, α ∈ A, Proof. Suppose that ϕ satisfies the condition; we are going to show that ϕ is open on p−1 U ⊂ p−1 1 (Vα), α ∈ A, and this will prove the 'if' part of the statement. So let 1 (Vα) be an open set and y0 ∈ ϕ(U ). Denote t0 := p2(y0) ∈ Vα and let x0 ∈ U be such that ϕ(x0) = y0 and f ∈ Γ(ξ1) with f (t0) = x0. There exist an open neighbourhood V of t0 in Vα and ε > 0 such that x0 ∈ {x ∈ p−1 1 (Vα) p1(x) ∈ V, kx − f (p1(x))k < ε} ⊂ U. If not then there would exist a net {xι} in p−1 1 (Vα) \ U such that {p1(xι)} converges to t0 and kxι − f (p1(xι))k → 0. But that means xι → f (t0) = x0 ∈ U , a contradic- tion. Hence V and ε > 0 as claimed indeed exist. Put f := ϕ(f ). We claim now that the open neighbourhood U ′; = {y ∈ p−1 2 (Vα) p2(y) ∈ V, ky − f (p2(y))k < ε/mα} of y0 is contained in ϕ(U ). Indeed, let y ∈ U ′. By our assumption, there exists x ∈ p−1 1 (p2(y)) such that ϕ(x) = y − f (p2(y)) and kxk ≤ mαky − f (p2(y))k < ε. Since k(x + f (p2(y))) − f (p2(y))k = kxk < ε we get x + f (p2(y)) ∈ U and y = ϕ(x + f (p2(y))) ∈ ϕ(U ) as needed. Suppose now that ϕ is open. Let t ∈ T . The set ϕ({x ∈ E1 kxk < 1}) is an open neighbourhood of 0t. There exist an open neighbourhood Vt of t and a positive number kt such that {y ∈ E2 p2(y) ∈ Vt, kyk < kt} ⊂ ϕ({x ∈ E1 kxk < 1}). If not then there would exist a net {yι} in E2 \ ϕ({x ∈ E1 kxk}) such that {p2(yι)} converges to t and kyιk → 0. But then yι → 0t, a contradiction. We are going to 14 ALDO J. LAZAR show that Vt and kt have the property that for y ∈ p−1 2 (Vt) there exists x such that ϕ(x) = y and kxk ≤ 2 kt kyk. Now, for y ∈ p−1 2 (Vt), y 6= 0, there exists x′ ∈ p−1 1 (p2(y)) such that kx′k < 1 and We got ϕ(x′) = mt 2kyk y. ϕ( 2kyk kt x′) = y, k 2kyk kt x′k ≤ 2 kt kyk. (cid:3) The openness of a map between two Banach bundles intervenes also when looking at the preimage of a Banach subbundle. Returning to Example 3.1, one can see that the preimage of the null subbundle {0t t ∈ T } of ξ2 by the map ϕ is E ′ := {(0, (K × {0}))} ∪ {(0, 1] × {(0, 0)} which cannot be the space bundle of a subbundle of ξ1. The subset {(0, ({Re(x) > 0 x ∈ K} × {0})} is relatively open in E ′ but its image by the projection p1 is {0}. Proposition 3.4. Let ϕ be an open Banach bundle map of the Banach bundle ξ1 := (E1, p1, T ) 2 := (E ′ ξ′ 1, p1 E ′ (E ′ 2, p2 E ′ 1, T ) is a subbundle of ξ1. onto the Banach (E2, p2, T ) 2, T ) be a Banach subbundle of ξ2. Then, with E ′ bundle := ξ2 and let 1 := ϕ−1(E ′ 2), Proof. Let U ⊂ E1 be open and U ∩ E ′ relatively open in E ′ 1 6= ∅. Then ϕ(U ∩ E′ 1) = p2(ϕ(U ) ∩ E ′ 1) = ϕ(U ) ∩ E ′ 2) is open in T . We 2 is 2. It follows that p1(U ∩ E ′ 1 is an open map for the relative topology of E ′ 1. conclude that p1 E ′ (cid:3) We turn our attention now to the spaces of bounded sections. In order to have ϕ(Γb(ξ1)) ⊂ Γb(ξ2) for a Banach bundle map between two Banach bundles it is quite natural to require supt∈T kϕtk < ∞. However, even in the situation of Theorem 3.2 one does not necessarily have ϕ(Γb(ξ1)) = Γb(ξ2). Indeed, let E1 = E2 := N × K, p1 = p2 the projection on N, and ξi := (Ei, pi, N), i = 1, 2, ϕ(n, x) := (n, x/n2). The bounded section of ξ2 given by n → (n, 1/n) is the image of only one section of ξ1: the unbounded section n → (n, n). A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 15 Theorem 3.5. Let T be a paracompact space and ξi := (Ei, pi, T ), i = 1, 2, Banach bundles. Let ϕ be a Banach bundle map of ξ1 onto ξ2 such that supt∈T kϕtk < ∞. Then ϕ(Γb(ξ1)) = Γb(ξ2) if and only if there is a positive number m with the property that for each y ∈ E2 there exists x ∈ E1 such that ϕ(x) = y and kxk ≤ mkyk. Proof. Suppose a constant m > 0 as in the above statement exists. Then ϕ is an open map by Proposition 3.3. Let f ∈ Γb(ξ2) and define the set valued map Φ(t) := {x ∈ p−1 non-void convex subset of p−1 1 (t) ϕ(x) = f (t), kxk < (m + 1)kf k}, t ∈ T . Then Φ(t) is a 1 (t) for every t ∈ T . We are going to show that Φ is lower semi-continuous. Let U be an open subset of E1 and suppose U ∩ Φ(t0) 6= ∅. Thus there exists x0 ∈ U ∩ {x ∈ E1 kxk < (m + 1)kf k} ∩ Φ(t0). The subset ϕ(U ∩ {x ∈ E1 kxk < (m + 1)kf k} of E2 is an open neighbourhood of f (t0); hence there exists a neighbourhood V of t0 in T such that f (t) ∈ ϕ(U ∩ {x ∈ E1 kxk < (m + 1)kf k}). t ∈ V. Therefore, for each t ∈ T there exists xt ∈ U ∩ {x ∈ E1 kxk < (m + 1)kf k} such that ϕ(xt) = f (t). It follows that V is a neighbourhood of t0 such that U ∩ Φ(t) 6= ∅ for every t ∈ V and we obtained that Φ is lower semi-continuous. Now, it is easily seen that t → Φ(t) is lower semi-continuous too. Moreover, from Φ(t) ⊂ p−1 and the fact that p−1 1 (t) 1 (t) is closed in E1 we gather that kxk ≤ (m + 1)kf k for each x ∈ Φ(t). Thus, a continuous selection g of t → Φ(t) given by Theorem 2.9 is a bounded section of ξ1 that satisfies ϕ(g) = f and we are done with this half of the proof. Suppose now that ϕ(Γb(ξ1) = Γb(ξ2). It is a well known consequence of the Banach open mapping theorem that there exists a constant k > 0 with the property that for every f ∈ Γ(ξ2) there exists g ∈ Γb(ξ1) such that ϕ(g) = f and kgk ≤ kkf k. We claim that for every y ∈ E2 there exists x ∈ E1 such that ϕ(x) = y and kxk ≤ 3k/2kyk. Indeed, for y ∈ E2, y 6= 0, denote t := p2(y) and let f ′ ∈ Γb(ξ2) satisfy f ′(t) = y. There exists a neighbourhood V of t in T such that kf ′(t′)k < 3/2kf ′(t)k whenever t′ ∈ V . Let h : T → [0, 1] be a continuous function such that h(t) = 1 and h (T \ V ) ≡ 0. Then f ; = hf ′ satisfies f (t) = y and kf k ≤ 3/2kyk. Let now g ∈ Γb(ξ1) be such that ϕ(g) = f and kgk ≤ kkf k. Then ϕ(g(t)) = y and kg(t)k ≤ kgk ≤ kkf k ≤ 3k/2kyk 16 ALDO J. LAZAR and this establishes our claim. (cid:3) Remark 3.6. The 'if' direction of the above theorem in the case of a quotient map is part of [9, Theorem 9.14]. For a Banach bundle ξ whose base space is locally compact Hausdorff the space Γ0(ξ) is of interest and we shall look now at the behavior of such spaces under a Banach bundle map. Lemma 3.7. Let ξ := (E, p, T ) be a Banach bundle and h : T → (0, ∞) a contin- uous function. Then F := {x ∈ E p(x) = t, kxk < h(t)} is an open subset of E. Proof. Let x0 ∈ F and put t0 = p(x0). There exists an open neighbourhood V of t0 such that h(t) > h(t0) − 1/2(h(t0) − kx0k) if t ∈ V . Now, it is easily checked that p−1(V ) ∩ {x ∈ E kxk < 1/2(kx0k + h(t0))} ⊂ F and we are done. (cid:3) Theorem 3.8. Let ξ1 := (E1, p1, T ) be a Banach bundle and ξ2 := (E2, p2, T ) be a continuous Banach bundle with T locally compact Hausdorff. Suppose that ϕ is a Banach bundle map of E1 onto E2 such that supt∈T kϕtk < ∞. If there exists a positive constant m with the property that for each y ∈ E2 there exists x ∈ E1 satisfying ϕ(x) = y and kxk < mkyk then ϕ(Γ0(ξ1)) = Γ0(ξ2). Proof. Let g ∈ Γ0(ξ1) and ε > 0. From k ϕ(g)(t)k = kϕ(g(t))k ≤ kg(t)k sup t′∈T kϕt′ k we obtain that {t ∈ T k ϕ(g)(t)k ≥ ε} is a closed subset of the compact set {t ∈ T kg(t)k ≥ ε/ supt′ kϕt′k}. Thus ϕ(g) ∈ Γ0(ξ2) and we got ϕ(Γ0(ξ1)) ⊂ Γ0(ξ2). Let now f ∈ Γ0(ξ2). Then T ′ := {t ∈ T kf (t)k > 0} is an open σ-compact subset of T ; it is paracompact by [6, Corollary 2, p. 211]. Define Φ(t) := {x ∈ p−1 1 (t) ϕ(x) = f (t), kxk < (m + 1)kf (t)k} for t ∈ T ′. Then Φ(t) is a non-void convex subset of the Banach space p−1 1 (t). We are going to show that t → Φ(t) is lower semi-continuous on T ′. With U an open A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 17 subset of p−1 (m + 1)kf (t0)k and x0 ∈ U ∩ {x ∈ p−1 The set U ′ := U ∩ {x ∈ p−1 1 (T ′) suppose U ∩ Φ(t0) 6= ∅, t0 ∈ T ′. Let x0 ∈ U ∩ Φ(t0). Then kx0k < 1 (T ′) kxk < 1/2(kx0k + (m + 1)kf (t0)k)}. 1 (T ′) kxk < 1/2(kx0k + (m + 1)kf (t0)k)} is an open subset of E1 and ϕ is an open map by Proposition 3.3; hence ϕ(U ′) is an open neighbourhood of ϕ(x0) = f (t0). There exists an open neighbourhood V of t0 in T ′ such that f (t) ∈ ϕ(U ′) and kf (t)k > kf (to)k − 1/2(kf (t0)k − kx0k/(m + 1)) for t ∈ V . Thus, if t ∈ V there exists xt ∈ U ′ such that ϕ(xt) = f (t). Then kxtk < 1/2(kxok + (m + 1)kf (t0)k) < (m + 1)kf (t)k and we have xt ∈ U ∩ Φ(t). We got t0 ∈ V ⊂ U ∩ Φ(t) and we can conclude that t → Φ(t) is indeed lower semi-continuous. The set valued map t → Φ(t), t ∈ T ′, is lower semi-continuous too and each set Φ(t) is a non-void closed subset of {x ∈ p−1 1 (t) kxk ≤ (m+1)kf (t)k}. By Theorem 2.9 there exists a continuous selection g′ of Φ on T ′. We have ϕ(g′(t)) = f (t) and kg′(t)k ≤ (m + 1)kf (t)k for each t ∈ T ′. Define now g(t) =(g′(t), 0t, if t ∈ T ′, if t ∈ T \ T ′. The function g : T → E is continuous. Indeed, let {tα} be a net in T ′ that converges to t ∈ T \ T ′. then f (tα) → f (t) = 0t; from 0 ≤ kg(tα)k ≤ (m + 1)kf (tα)k we derive kg(tα)k → 0. Hence g(tα) → 0t = g(t). If ε > 0 then {t ∈ T kg(t)k ≥ ε} is compact since it is a closed subset of the compact set {t ∈ T kf (t)k ≥ ε/(m + 1)}. We got g ∈ Γ0(ξ1) with ϕ(g) = f and we conclude ϕ(Γ0(ξ1)) = Γ0(ξ2). (cid:3) We shall now consider results suggested by [3] and [17] on the existence of right inverses for some Banach bundle maps. Proposition 3.9. Let ϕ be an open Banach bundle map of ξ1 := (E1, p1, T ) onto ξ2 := (E2, p2, T ). Suppose that E2 is paracompact. Then there exists a continuous map ψ : E2 → E1 such that ϕ(ψ(y)) = y for every y ∈ E2. Proof. Define Φ(y) := ϕ−1(y), y ∈ E2. By using the fact that ϕ is open one easily sees that Φ is lower semi-continuous. An application of Theorem 2.9 in the obvious manner yields a continuous selection ψ : E2 → E1 of Φ that fulfills what is needed. (cid:3) 18 ALDO J. LAZAR To obtain a homogeneous right inverse we shall suppose that the range of the Banach bundle map is a continuous Banach bundle. But first we have to state two simple lemmas. We omit their elementary proofs; the first can be proven by the means of an obvious compactness argument and the second lemma is an easy consequence of the first. We let µ be the normalized Lebesgue measure on the unit circle in C and [λ1λ2 will denote the arc of the unit circle from λ1 to λ2 in the positive direction. Lemma 3.10. Let ξi := (Ei, pi, T ), i = 1, 2 be two Banach bundles over the complex field, f : E2 → E1 a continuous function and {yα}α∈A a net in E2 that converges to y ∈ E2. Given ε > 0 there exists δ > 0 such that f (λ1yα) − f (λ2yα) < ε and f (λ1y) − f (λ2y) < ε whenever λ1 = λ2 = 1, µ([λ1λ2) < δ and for every α ∈ A. Lemma 3.11. Let ξ1, ξ2, f , {yα}α∈A, and y be as in Lemma 3.10. Given ε > 0 there exists δ > 0 with the following property: if {λi}n i=1, λ1 = λn, is a division of the unit circle in the complex plane with mesh smaller than δ then f (λy)dµ(λ) − Zλ=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (λyα)dµ(λ) − Zλ=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) and for every α ∈ A. n−1 Xi=1 n−1 Xi=1 f (λiy)µ( \λiλi+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) f (λiyα)µ( \λiλi+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < ε < ε Proposition 3.12. Let ϕ be a Banach bundle map from the Banach bundle ξ1 := (E1, p1, T ) onto the continuous Banach bundle ξ2 := (E2, p2, T ). Suppose that E2 is paracompact and there exists m > 0 with the propriety that for every y ∈ E2 there exists x ∈ E1 such that ϕ(x) = y and kxk ≤ mkyk. Then there exists a continuous map ψ := E2 → E1 such that ϕ(ψ(y)) = y and ψ(λy) = λψ(y) for every y ∈ E2 and λ ∈ K. Proof. The set S := {y ∈ E2 kyk = 1} is closed in E2 thus paracompact. We define Φ(y) := {x ∈ E1 ϕ(x) = y, kxk < m + 1}, y ∈ S. Then Φ(y) is a non-empty convex subset of p−1 1 (p2(y)) and we want to show that Φ is lower semi- continuous. So let U be an open subset of E1 and suppose U ∩ Φ(y0) 6= ∅. Then A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 19 U ′ := ϕ(U ∩ {x ∈ E1 kxk < m + 1}) is an open subset of E2 since ϕ is an open map by Proposition 3.3 and y0 ∈ U ′ ∩ F . If y ∈ U ′ ∩ F then there exists x ∈ U such that kxk < m + 1 and ϕ(x) = y. Thus U ∩ Φ(y) 6= ∅ if y ∈ U ′ ∩ F and we have proved that Φ is lower semi-continuous. The map y → Φ(y) is also lower semi-continuous and Theorem 2.9 yields a continuous function g : F → E1 such that ϕ(g(y)) = y and kg(y)k ≤ m + 1 for y ∈ F . By defining h(y) :=(kykg(y/kyk), 0, if kyk 6= 0, otherwise for y ∈ E2 one gets a continuous function from E2 to E1 such that ϕ(h(y)) = y for y ∈ E2.. If the scalars are real then one easily checks that ψ(y) := 1/2(h(y)−h(−y)) has the required properties. Suppose now that the scalars are complex. Then we define ψ(y) := Zλ=1 ¯λh(λy)dµ(λ), y ∈ E2. The function ψ has all the needed properties. We shall prove only its continuity all its other attributes being straightforward to check. So let y ∈ E2 and suppose {yα}α∈A is a net that converges to y. We denote t := p2(y) and tα := p2(yα). Let U be an open neighbourhood of ψ(y). Now T is paracompact being homeomorphic to the closed subspace {0τ }τ ∈T of E2. So there exist a section σ of ξ1, a neighbourhood V of t contained in p1(U ) and ε > 0 such that σ(t) = ψ(y) and U ′ := {x ∈ E1 p1(x) ∈ V, kx − σ(p1(x))k < 2ε} ⊂ U. Let now δ be the positive number given by Lemma 3.11 for our number ε and the integrant appearing in the definition of ψ. Suppose {λi}n of the unit circle of mesh less than δ. Then kψ(y) −Pn−1 We denote i=1 i=1, λ1 = λn, is a division ¯λih(λiy)µ( \λ1λi+1)k < ε. U ′′ := {x ∈ E1 p1(x) ∈ V, kx − σ(p1(x))k < ε}; i=1 ¯λih(λiy)µ( \λiλi+1) belongs to the open set U ′′. There exists α0 ∈ A ¯λih(λiyα)µ( \λiλi+1) also belongs to U ′′. Thus we ¯λih(λiyα)µ( \λiλi+1) − σ(p2(yα))k < ε if α ≻ α0 and then Pn−1 such that if α ≻ α0 then Pn−1 have kPn−1 Xi=1 kψ(yα) − i=1 i=1 converges to ψ(y). n−1 ¯λih(λiyα)µ( \λiλi+1)k < ε. We obtained that ψ(yα) ∈ U ′ ⊂ U if α is large enough so we conclude that {ψ(yα)} 20 ALDO J. LAZAR (cid:3) Remark 3.13. It follows from the proof of Proposition 3.12 that the map ψ satisfies kψ(y)k ≤ (m + 1)kyk for every y ∈ E2. Thus one can easily obtain a proof of Theorem 3.8 by using Proposition 3.12 if in addition to the other hypotheses of that theorem one supposes that E2 is paracompact. 4. M-ideals of Γ0(ξ) For the present section ξ := (E, p, T ) will be a fixed Banach bundle. We shall mainly discuss the M-ideals of the Banach space Γ0(ξ) when the base space is locally compact Hausdorff; most of the results that follow are generalizations of results from [9] where only the case of a compact Hausdorff base space was considered. As in [9], for a closed subset A of T the notation NA will stand for {σ ∈ Γb σ(t) = 0t for all t ∈ A} or for {σ ∈ Γ0(ξ) σ(t) = 0t for all t ∈ A} if T is locally compact Hausdorff; it will be clear from the context which is the case. If A is a singleton, say {t}, then we shall use the notation Nt instead of N{t}. In [9, Proposition 13.6] the base space is assumed to be compact Hausdorff. However the proof given there is valid with obvious modifications in the more general situations detailed in the next Proposition. Of course, we omit its proof. Proposition 4.1. The closed subspace NA is an M-ideal of Γb(ξ) if A is a closed subset of the normal space T . For a locally compact Hausdorff base space T NA is an M-deal of Γ0(ξ) if A is a compact subset of T or if A is a closed subset of T such that the complement of its interior is compact. Proposition 4.2. Suppose T is paracompact (locally compact Hausdorff ) and A and B are closed (compact, respectively) subsets of T . Then NA∩B = NA + NB. Proof. We must prove only NA∩B ⊂ NA + NB. Let σ ∈ NA∩B and define ρ : A ∪ B → E by ρ(t) := σ(t) if t ∈ A and ρ(t) := 0t if t ∈ B. Then ρ is well defined and continuous on A ∪ B. Corollary 2.10 provides us with an extension of in Γb(ξ) (Γ0(ξ), respectively); we shall call this extension ρB. We have ρB ∈ NB and ρA := σ − ρB is in A. (cid:3) A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 21 Proposition 4.2 appeared as [9, Corollary 15.8] for T compact Hausdorff and was given there a different proof from the one above. An M-ideal L of a Banach space X is called primitive if there is an extreme point ω of the closed unit ball of the dual space X ′ such that L is maximal among the M-ideals contained in the null subspace of ω. If L is a primitive M-ideal and L1, L2 are M-ideals such that L1 ∩ L2 ⊂ L then L1 ⊂ L or L2 ⊂ L. Every M-ideal is the intersection of the primitive M-ideals containing it. For all these facts see [1, Section 3]. In the remainder of this section the base space T will always be a locally compact Hausdorff space. The case T compact Hausdorff of the following proposition is part of [9, Propo- sition 13.11]. Proposition 4.3. If L is a primitive M-ideal of Γ0(ξ) then there is a unique t0 ∈ T such that Nt0 ⊂ L. Proof. Let K be a non-void compact subset of T . Then NK and NT \K are M- ideals of Γ0(ξ) by Proposition 4.1 and NK ∩ NT \K = {0} ⊂ L. We infer that NK or NT ⊂K is included in L. If for every such K we have NT \K ⊂ L then L = Γ0(ξ) since the sections with compact support are dense in Γ0(ξ). Thus there exists a non-void compact set K such that NK ⊂ L. Let F be the non-void family of all non-void compact subsets of T such that NK ⊂ L. If K1 and K2 belong to F then NK1∩K2 = NK1 + NK2 ⊂ L and it follows that F has the finite intersection property. Hence K0 := ∩{K K ∈ F} 6= ∅. Denote Λ := ∪ K∈F NK. We claim that NK0 = Λ. Obviously Λ ⊂ NK0. Let now σ ∈ NK0 and ε > 0. Set V := {t ∈ T kσ(t)k < ε}; then V is open and K0 ⊂ V . If it were true that for each K ∈ F there would be tK ∈ K \ V then the net {tK}K∈F would have a convergent subnet whose limit would be in K0. On the other hand, kσ(tk)k ≥ ε for all K ∈ F , a contradiction. Thus there exists K ∈ F such that K ⊂ V . Let f : T → [0, 1] be a continuous function such that f K ≡ 0 and f (T \ V ) ≡ 1. Then with σ′ := f · σ we have σ′ ∈ NK and kσ − σ′k < ε. We infer that σ ∈ Λ and the claim follows from this. In particular, K0 ∈ F . 22 ALDO J. LAZAR It remains only to show that K0 consists of only one point. If there are in K0 at least two distinct points then there exist compact subsets K1, K2 of K0 such that K0 = K1 ∪ K2 and K1 6= K0 6= K2. These subsets will have to satisfy NK1 ∩ NK2 = NK0 ⊂ L. Hence NK1 ⊂ L or NK2 ⊂ L. But Ki ∈ F , i = 1 or 2, and Ki $ K0 is a contradiction. Therefore K0 = {t0} for some t0 ∈ T and we found Nt0 ⊂ L. The uniqueness of t0 with the property Nt0 ⊂ L follows from the definitions of F and K0. (cid:3) Now we can apply the same proof as in the compact case of [9, Proposition 15.20] to get Proposition 4.4. Every M-ideal L of Γ0(ξ) is a Cb(T )-module. Proof. Suppose first that L is a primitive M-ideal and let t0 ∈ T be the element given by Proposition 4.3. If f ∈ Cb(T ) and σ ∈ L then f · σ − f (t0)σ belongs to Nt0 ⊂ L. Hence f · σ ∈ L and the case of a primitive M-ideal is settled. The general case follows from this particular case and the fact mentioned above that each M-ideal is the intersection of the primitive M-ideals containing it. (cid:3) We proceed now to give a characterization of those closed subspaces of Γ0(ξ) that are Cb(T )-modules. The parallel result for paracompact base spaces and spaces of bounded sections is [9, Theorem 8.6]. First we state and prove a simple lemma. Lemma 4.5. Let L be an M-ideal of Γ0(ξ), σ ∈ L, and suppose that kσ(t0)k < θ for some t0 ∈ T and θ > 0. Then there is σ′ ∈ L such that σ′(t0) = σ(t0) and kσ′k < θ. Proof. Denote θ′ := 1/2(θ + kσ(t0)k). Then V := {t ∈ T kσ(t)k < θ′} is an open neighbourhood of t0 and T \ V is compact. Let f : T → [0, 1] be a continuous function such that f (t0) = 1 and f (T \ V ) ≡ 0. Define σ′ := f · σ. It is easily checked that σ′ has the required properties. (cid:3) A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 23 Theorem 4.6. A closed subspace of Γ0(ξ) is a Cb(T )-submodule if and only if there exists a subbundle η of ξ such that L = Γ0(η). Proof. Obviously Γ0(η) for η a subbundle of ξ is a Cb(T )-submodule of Γ0(ξ). Suppose now that the closed subspace L of Γ0(ξ) is a Cb(T )-submodule. Denote F (t) := {σ(t) σ ∈ L}, t ∈ T . Clearly F (t) is a linear subspace of E(t) and we are going to show that it is closed in E(t). Let {xn} be a sequence in F (t) that converges to x ∈ E(t). By passing to a subsequence if necessary we may suppose that kxn+1 − xnk < 2−n for each n ∈ N. There exist σn ∈ L such that σn(t) = xn, n ∈ N. Put 1 := σ1; Lemma 4.5 yields n ∈ L with n(t) = σn+1(t) − σn(t) and knk < 2−n, n ≥ 2. The section ; =P∞ we deduce that x ∈ F (t). n=1 n is in L and (t) = lim σn(t) = x and Set F := ∪t∈T F (t); we are going to show that p F is open. Let U be an open subset of E and t0 ∈ p(U ∩ F ). There exists σ ∈ L such that σ(t0) ∈ U ∩ F . there exists a neighbourhood V of t0 in T such that σ(t) ∈ U whenever t ∈ V . Then V ⊂ p(U ∩ F ) and we obtained that p(U ∩ F ) is an open subset of T . We have shown that η := (F , p F , T ) is a subbundle of ξ. We want to show now that L = Γ0(η). Clearly L ⊂ Γ0(η). Let σ ∈ Γ0(η) and ε > 0. Then K := {t ∈ T kσ(t)k ≥ ε} is a compact subset of T and, L being a Cb(T ) module, for each t ∈ K there exists t ∈ L such that t(t) = σ(t) and t (T \ K) ≡ 0. The set Vt := {t′ ∈ T kσ(t′) − t(t′)k < ε}, t ∈ K, is open; for each t ∈ K we choose on open set Wt such that Wt is compact and t ∈ Wt ⊂ Wt ⊂ Vt. There is a finite subcover {Wti }n i=1 of the cover {Wt}t∈K of K; let fi : T → [0, 1] be a continuous function such that fi Wti ≡ 1 and fi (T \ Vti ) ≡ 0, 1 ≤ i ≤ n. On ∪n and extend each gi to a continuous and bounded function on T . For simplicity, the extension will be denoted also by gi. Define now ρ :=Pn for t ∈ K we have i=1 giti . Then ρ ∈ L and n i=1Wti set fi j=1 fj gi := Pn kρ(t) − σ(t)k ≤ gikti − σ(t)k < ε. Xi=1 Define the open neighbourhood V := {kρ(t) − σ(t)k < ε of K and let h : T → [0, 1] be a continuous function such that h K ≡ 1 and h (T \ V ) ≡ 0. Then h · ρ ∈ L 24 ALDO J. LAZAR and kh(t)ρ(t) − σ(t)k < ε on K. If t ∈ T \ V then kh(t)ρ(t) − σ(t)k = kσ(t)k < ε since K ⊂ V . Now, if t ∈ V \ K we have (10) kh(t)ρ(t) − σ(t)k ≤ h(t) − 1kρ(t)k + kρ(t) − σ(t)k ≤ (1 − h(t))(kρ(t) − σ(t)k + kσ(t)k) + kρ(t) − σ(t)k < 3ε. We have found h · ρ ∈ L such that kh · ρ − σ ≤ 3ε hence σ ∈ L and we conclude that L = Γ0(η). (cid:3) The case of a compact base space of the following result is [9, Theorem 15.21] for which only a brief indication of the proof was given. Theorem 4.7. A closed subspace L of Γ0(ξ) is an M-ideal if and only if there exists a subbundle η := (F , p F , t) of ξ such that L = Γ0(η) and and each fiber F (t) is an M-deal of the Banach space E(t). Proof. Suppose first that L is an M-ideal of Γ(ξ). It follows from Proposition 4.4 and Theorem 4.6 that there exists a subbundle η := (F , p F , T ) of ξ such that L = Γ0(η). It remains to show that F (t) = {σ(t) σ ∈ L} is an M-ideal of E(t). For this purpose we proceed to show that F (t0) has the 3-ball-property in E(t0), t0 ∈ T . Let {xi}3 exists x ∈ E(t0) with kxi − xk < ri, 1 ≤ i ≤ 3, and there exist {yi}3 i=1 be positive numbers such that there i=1 in F (t0) such that kxi − yik < ri, 1 ≤ i ≤ 3. There are sections σ, σi ∈ Γ0(ξ), such that i=1 ⊂ E(t0) and let {ri}3 σ(t0) = x and σi(t0) = xi, 1 ≤ i ≤ 3, and sections i ∈ Γ0(η) such that i(t0) = yi, 1 ≤ i ≤ 3. The set V ; = {t ∈ T kσi(t) − σ(t)k < ri, kσi(t) − i(t)k < ri, 1 ≤ i ≤ 3} is an open neighbourhood of t0. Let f : T → [0, 1] be a continuous function such that f (t0) = 1 and f (T \ V ) ≡ 0. Then f · σ, f · σi ∈ Γ0(ξ) and f · i ∈ Γ0(η), 1 ≤ i ≤ 3. We have kf · σi − f · σk < ri and kf · σi − f · ik < ri, 1 ≤ i ≤ 3. Since Γ0(η) is an M-ideal there is a section ∈ Γ0(η) such that kf · σi − k < ri, 1 ≤ i ≤ 3. Then (t0) ∈ F (t0) and kxi − (t0)k = kf (t0)σi(t0) − i(t0)k < ri, 1 ≤ i ≤ 3. Thus F (t0) has indeed the 3-ball-property in E(t0). Now we suppose that η := (F , p F , t) is a subbundle of ξ such that each F (t) is an M-ideal of E(t). Since each subspace F (t) is closed in E(t) it is obvious that A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 25 Γ0(η) is a closed subspace of Γ0(ξ). We are going to show that Γ0(η) has the 3-ball- property in Γ0(ξ) and is thus an M-ideal of Γ0ξ). Let {ri}3 i=1 be positive numbers, σi, σ ∈ Γ0(ξ), and i ∈ Γ0(η), 1 ≤ i ≤ 3, so that kσi − σk < ri and kσi − ik < ri, 1 ≤ i ≤ 3. Let r′ i be a number satisfying max{kσi − σk, kσi − ik} < r′ i < ri, 1 ≤ i ≤ 3 and set C(t) := ∩3 i=1{y ∈ F (t) kσi(t) − yk < r′ i}. The subspace F (t) of E(t) has the 3-ball-property; hence C(t) is a non-void convex subset of F (t), t ∈ T . We are going to show that the map t → C(t) is lower semi-continuous. Let U be an open subset of E and suppose C(t0) ∩ U 6= ∅ for some t0 ∈ T . Then there exists an element y0 ∈ C(t0) ∩ U and 0 ∈ Γ0(η) such that 0(t0) = y0. The open neighbourhood V := (∩3 i=1{t ∈ T kσi(t0) − 0(t0)k < r′ i}) ∩ {t ∈ T (t) ∈ U } of t0 satisfies V ⊂ {t ∈ T C(t) ∩ U 6= ∅} and the claim about t → C(t) is proved. We infer that t → C(t), t ∈ T , is also lower semi-continuous. Remark that C(t) is a non-void closed convex subset of F (t). Moreover, if y ∈ overlineC(t) then kyk ≤ r′ i The set K1 := ∪3 i=1{t ∈ T kσi(t)k ≥ r′ i/2} is compact and there exists an open subset U ⊃ K1 of T such that K2 : +U is compact. Let ρ′ be a continuous selection of t → C(t) over K2 given by Theorem 2.9 and f : T → [0, 1] be a continuous function satisfying f K1 ≡ 1 and f (T \U ) ≡ 0. Define ρ(t) :=(f (t) · ρ′(t), 0t, if t ∈ K2, if t ∈ T \ U . Then ρ is well defined and ρıΓ0(η). We claim that kσi − ρk ≤ r′ i, 1 ≤ i ≤ 3 and we shall conclude from this that Γ0(η) has the 3-ball-property in Γo(ξ). If t ∈ K1 then kσi(t) − ρ(t)k = kσi(t) − ρ′(t)k ≤ r′ i since ρ′(t) ∈ C(t). If t ∈ T \ U then kσi(t) − ρ(t)k = kσi(t)k < r′ i/2. Finally, if t ∈ U \ K1 then (11) kσi(t) − ρ(t)k ≤ kσi(t) − f (t) · σi(t)k + kf (t) · σi(t) − f (t) · ρ′(t)k (1 − f (t))kσi(t)k + f (t)kσi(t) − ρ′(t)k ≤ (1 − f (t))r′ i/2 + f (t)r′ i ≤ r′ i, and with this the claim is proved. 26 ALDO J. LAZAR (cid:3) 5. Uniform and locally uniform Banach bundles In this section we shall describe a class of Banach bundles that is more general than the class of locally trivial Banach bundles but it is still quite manageable. This class was introduced in [15] under the guise of continuous fields of Banach spaces. We remind the definition of the Hausdorff distance between two bounded subsets A, B of a metric space (M, ρ): d(A, B) := max{sup a∈A ρ(a, B), sup b∈B ρ(b, A)}. It is a metric in the family of all bounded closed subsets of M . Recall that if X is a Banach space then X1 denotes its closed unit ball. Lemma 5.1. Let X be a Banach space, M the family of its bounded closed subsets endowed with the Hausdorff metric, T a topological space, and t → X(t) a map Φ from T to the collection of closed subspaces of X. If the map t → X(t)1 from T to M is continuous then Φ is lower semi-continuous. Proof. Let O be an open subset of X and suppose that O ∩ X(to) 6= ∅, t0 ∈ T . Pick x0 ∈ O ∩ X(t0) and let r > 0 be so that B(x0, r) ⊂ O. From the continuity of t → X(t)1 it follows that there exists a neighbouhood V of t0 in T such that for each t ∈ V there exists yt ∈ X(t) satisfying kytk ≤ kx0k and kx0 − ytk < r. Thus, if t ∈ V then O ∩ X(t) 6= ∅ and the proof is complete. (cid:3) We keep the setting and the notations of Lemma 5.1; to be consistent with our blanket requirement we shall suppose that T is a Hausdorff topological space. We make {t} × X(t) into a Banach space by transfering onto it in the obvious way the structure of X(t), t ∈ T . Put E; = ∪t∈T ({t} × X(t)) and define p : E → T by p((t, x)) := t. For an open subset V of T and an open subset O of X let U(V, O) := {(t, x) t ∈ V, x ∈ O ∩ X(t)}. Proposition 5.2. The family of all the sets U(V, O) when V runs through all the open subsets of T and O runs through all the open subsets of X is a base of a topology on E. When E is endowed with this topology ξ := (E, p, T ) is a continuous Banach bundle. A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 27 Proof. If (t, x) ∈ U(V1, O1) ∩ U(V2, O2) then (t, x) ∈ U(V1 ∩ V2, O1 ∩ O2) ⊂ U(V1, O1) ∩ U(V2, O2); hence the first statement of the proposition. If V ⊂ T is open then p−1(V ) = U(V, X) so p is continuous. If V ⊂ T and O ⊂ X are open then p(U(V, O)) = V ∩{t ∈ T O∩X(t) 6= ∅} so Lemma 5.1 implies that this set is open in E. Thus ξ is a Banach bundle. Let (t0, x0) ∈ E, r > 0 and set V := {t ∈ T B(x0, r)∩X(t) 6= ∅}. The continuity of the norm at (t0, x0) ∈ E follows from (t0, x0) ∈ U(V, B(x0, r)) = {(t, x) t ∈ V, x ∈ B(x0, r) ∩ X(t)} ⊂ {(t, x) ∈ E k(t0, x0)k − r < k(t, x)k < k(t0, x0)k + r}. (cid:3) A continuous Banach bundle isomorphic to a Banach bundle as described in Proposition 5.2 is called a uniform Banach bundle. A (continuous) Banach bundle ξ := (E, p, T ) is called locally uniform if there exists a family of closed subsets {Tα}α∈A of T such that {Int(Tα)}α∈A is an open cover of T and each restriction ξ Tα is a uniform Banach bundle. One should recall that if T is paracompact or locally compact Hausdorff then each subset Tα inherits the same property. If two subspaces of a Banach space have different dimensions then, by [10], the distance between their closed unit balls is at least 1/2. Thus if ξ := (E, p, T ) is a locally uniform Banach bundle and for some t ∈ T the fiber E(t) is n-dimensional, n < ∞, then there exists a neighbourhood V of t in T such that for each s ∈ V the fiber E(s) is n-dimensional. It follows that if all the fibers of ξ are finite dimensional and T is locally compact Hausdorff then ξ is locally trivial by [9, Theorem 18.5]. On the other hand there exist uniform Banach bundles with infinite dimensional fibers that are not locally trivial. Indeed, Kadets [14] constructed a Banach space X having subspaces {Xn}∞ n=1 and Y such that each subspace Xn is isometric to the space ℓpn for a sequence {pn} that increases to 2 and Y is isometric to ℓ2}. Moreover, the the sequence of the closed unit balls of the subspaces Xn converges to the closed unit ball of Y 1. Thus we have here an example of a uniform Banach bundle whose base space is N ∪ {∞} with the natural topology that is not isomorphic to a locally 1The author is grateful to Professor E. Gluskin for providing him this reference. 28 ALDO J. LAZAR trivial bundle since, as it is well known, the spaces ℓp′ and ℓp′′ are not isomorphic if p′ 6= p′′. We describe now a way to construct locally uniform Banach bundles. Let T be a regular topological space, {Vα}α∈A a cover of T with open non-void subsets, {Xα}α a family of Banach spaces such that whenever Vα ∩ Vβ 6= ∅, Xα ∩ Xβ is a non-trivial closed subspace of Xα and of Xβ on which the norms k · kα and k · kβ coincide. Denote by Mα the hyperspace of all bounded closed subsets of Xα endowed with the Hausdorff metric. Suppose that for each t ∈ Vα, α ∈ A, X(t) is a closed subspace of Xα and t → X(t)1 is continuous as a map from Vα to Mα. Of course, if t ∈ Vα ∩ Vβ then X(t) ⊂ Xα ∩ Xβ. Put Eα := {(t, x) t ∈ Vα, x ∈ X(t)} and give Eα the topology specified in 5.2. That is, each pair of open subsets V ⊂ Vα and O ⊂ Xα determines an open subset U(V, O) := {(t, x) ∈ Eα t ∈ V, x ∈ O ∩ X(t)}. Observe that if Vα ∩ Vβ 6= ∅ then on Eα ∩ Eβ the two relative topologies coincide. Set now E := ∪α∈AEα and define a topology in E as follows: U ⊂ E is open if U ∩ Eα is open in Eα for every α ∈ A. Let p : E → T be the natural map. Proposition 5.3. With the above notations ξ := (E, p, T ) is a locally uniform Banach bundle. Proof. For t ∈ Vα let At,α be a closed set such that t ∈ Int(At,α) ⊂ At,α ⊂ Vα. Clearly the restriction of ξ to At,α is a uniform Banach bundle. (cid:3) Continuing with the above notations, if Y (t) is a closed subspace of X(t), t ∈ T , and t → Y (t)1 is continuous on T then, with F := {(t, x) t ∈ T, x ∈ Y (t)}, ζ := (F , p F , T ) is a (locally uniform) Banach subbundle of ξ. Indeed, if V is an open subset of Vα and O is an open subset of Xα then {(t, x) t ∈ V, x ∈ O ∩ X(t)} ∩ {(t, x) t ∈ Vα, x ∈ Y (t)} = {(t, x) t ∈ V, x ∈ O ∩ Y (t)} is a typical open subset of F . Now it is clear that the Banach bundle constructed in Proposition 5.2 is a Banach subbundle of a trivial Banach bundle. Indeed, with A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 29 the notations used there, let E ′ := T × X and p′ the projection on T . Then ξ is by the preceding remarks a Banach subbundle of ξ′ := (E ′, p′, T ). Note that E is a closed subset of E ′. To see this, let {(tι, xι)}ι∈I be a net in E converging to (t, x) ∈ T × X. There is no loss of generality if we suppose that xι ∈ BX (0X , 1), ι ∈ I. For each ι there exists yι ∈ E(t) such that kxι − yιk ≤ d(E(tι)1, E(t)1), d being the Hausdorff metric. Now d(E(tι)1, E(t)1) → 0 so (t, x) ∈ E. On the other hand, not every Banach subbundle of a trivial Banach bundle which has a closed bundle space is a uniform Banach bundle. Example 5.4. Let T := N ∪ {∞} with the usual topology, c0 the Banach space of all null-convergent sequences of scalars, E := T × c0 and p the projection of E onto T . We shall construct a Banach subbundle of the trivial Banach bundle ξ := (E, p, T ). Denote by en ∈ c0 the sequence {δk k=1 and let F (n) be the linear subspace of {n} × c0 spanned by (n, e1) and (n, en). We let F (∞) be the one-dimensional space n}∞ spanned by (∞, e1). With F := ∪t∈T F (t) the Banach subbundle ζ := (F , p F , T ) of ξ has a closed bundle space. However, the bundle ζ is not locally uniform because of the drop in dimension at ∞. We have seen in Theorem 3.8 that for a bundle map it is of interest to have a continuous norm in the image bundle. We discuss now conditions that insure that the norm in a quotient bundle is continuous. It may happen that a quotient bundle of a continuous Banach bundle has a discontinuous norm as the following example shows. Example 5.5. Let T := N ∪ {∞} with the usual topology, ξ := (T × K, p, T ) the one-dimensional trivial bundle over T , and η its subbundle whose bundle space is F := (N×K)∪({∞}×{0}). Then the quotient bundle space is (N×{0})∪({∞}×K) with a non-Hausdorff topology. Hence the quotient bundle is not a continuous Banach bundle. Question 5.6. Suppose η is a quotient bundle of a continuous Banach bundle and the bundle space of η is Hausdorff. Must η be a continuous Banach bundle? Proposition 5.7. Let ξ := (E, p, T ) be a locally uniform Banach bundle and ξ′ := (E ′, p E ′, T ) a locally uniform subbundle. Then the norm on the quotient bundle η := (E/E ′, p′, T ) is continuous. 30 ALDO J. LAZAR Proof. A moment's reflection shows that it is enough to consider only the case when ξ and ξ′ are uniform. Thus we shall suppose that ξ is a bundle as described in Proposition 5.2 keeping the notation used there. We shall also suppose that Y (t) is a closed subspace of X(t) and that the map t → Y (t)1, t ∈ T , is continuous into the the space of all closed subsets of the ambient space X. Then E; + ∪t∈T ({t} × X(t)) and E ′; = ∪t∈T ({t} × Y (t)). Denote by q the quotient map from E to E/E ′ and let α ≥ 0; we are going to show that O := {y ∈ E/E ′ kyk > α} is open in E/E ′. Suppose then that y0 ∈ O and choose numbers β and ε satisfying α < β < ky0k and 0 < ε < min{1/2, 1/3(ky0k−β)}. Let (t0, x0) ∈ E with q((t0, x0)) = y0 and put r := 2kx0k+2. There exists a neighbourhood V of t0 in T such that d(Y (t)r, Y (t0)r) < ε if t ∈ V . The set U(V, B(x0, ε)) := {(t, x) t ∈ V, x ∈ B(x0, ε) ∩ X(t)} is an open neighbourhood of (t0, x0) and q(U(V, B(x0, ε))) is an open neighbourhood of y0, q being an open map. We shall show that q(U(V, B(x0, ε))) ⊂ O and this will establish our claim about O thereby completing the proof. Let (t, x) ∈ U(V, B(x0, ǫ)) and denote y := q(x). Our goal is to show that kyk > α. There exists x ∈ Y (t) such that kyk ≤ kx − xk < kyk + ε. We have kxk ≤ r. Indeed, if we assume by contradiction that kxk > r = 2kx0k + 2 then kxk > 2kx − (x − x0)k + 2 > 2(kxk − ε) + 2 > 2kxk + 1. But then kx − xk ≥ kxk − kxk > kxk + 1 > kyk + ε, a contradiction. Now, since t ∈ V and x ∈ Y (t)r, there exists x0 ∈ Y (t0)r such that kx − x0k < ε. We have (12) kyk > kx − xk − ε ≥ k(x − x0) + (x0 − x0)k − kx0 − xk − ε ≥ kx0 − x0k − kx − x0k − kx0 − xk − ε ≥ ky0k − 3ε > β > α and we are done. (cid:3) The quotient bundle of a uniform Banach bundle may have a continuous norm even when the subbundle that produces it is not a locally uniform subbundle as the following example shows. Example 5.8. As before we denote by c0 the Banach space of all null converging sequences of scalars and en := {δk k=1, n ≥ 1. Let T := N ∪ {∞} with the usual n}∞ A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 31 topology, E := T × c0, and p be the projection onto T . We let Xn be the subspace l=1, E ′(n) := {n} × Xn, and E ′(∞) := E(∞). We affirm of c0 spanned by {el}n that ξ′ := (E ′, p E ′, T ) is a subbundle of the trivial bundle ξ := (E, p, T ). Indeed, suppose U ⊂ E is open and (∞, x) ∈ U ∩ E ′(∞). There are r > 0 and m ∈ N with the property that if y ∈ c0, ky − xk < r, and n ≥ m then (n, y) ∈ U . Let k ∈ N be such that ¯x; = (x1, . . . , xk, 0 . . . ) satisfies k¯x − xk < r. Then if n > max{m, k} one has (n, ¯x) ∈ U ∩ E ′(n). Thus we found a neighbourhood of ∞ included in p(U ) and the claim is proved. Clearly ξ′ is not a locally uniform subbundle because of the disparity in the di- mension of the fiber at ∞ with the dimensions of the other fibers in any neighbour- hood of ∞. It remains to show that the norm in the quotient bundle is continuous. Only the continuity at the points of the fiber at ∞ must be proven; but since the fiber at ∞ of the quotient bundle is trivial this fact is obvious. Appendix A. Some topological properties of the bundle space We shall now record several topological properties of the bundle space of a Ba- nach bundle. In the following ξ := (E, p, T ) will denote a Banach bundle. The first result treats the complete regularity of the bundle space. A result with the same conclusion but under different hypotheses is [11, Lemma 2.1.6]. Proposition A.1. If ξ is a continuous Banach bundle and T is locally paracompact then E is completely regular. Proof. First remark that E is a Hausdorff space since the norm is continuous on it. Let A be a closed subset of E and x0 ∈ E with x0 /∈ A. There are an open neighbourhood V of t0 := p(x0) in T such that V is paracompact and a continuous function h : T → [0, 1] satisfying h(t0) = 1 and h T \ V ≡ 0. Corollary 2.10 yields a section ϕ of ξ V such that ϕ(t0) = x0; there exists α > 0 such that {x ∈ E p(x) ∈ V, kx − ϕ(p(x))k < α} ∩ A = ∅. Let µ : [0, ∞) → [0, 1] be given by µ(s) :=(1 − s/α, 0, if 0 ≤ s ≤ α, if α < s. The function f : E → [0, 1] defined by f (x) :=(µ(kx − ϕ(p(x))k)h(p(x)), 0, if p(x) ∈ V , if p(x) /∈ V 32 ALDO J. LAZAR satisfies f (x0) = 1 and f A ≡ 0 as needed. (cid:3) We proceed now to discuss a setting in which E is a Baire space. We begin with some definitions and results from [19]. A family B of non-void open subsets of a topological space is call ed a pseudo-base of the space if every open non-void subset contains an element of B. A topological space Z is called quasi-regular if each non- void open set of Z contains the closure of some non-void open set. A topological space Z is called pseudo-complete if it is quasi-regular and there exists a sequence {Bn} of pseudo-bases of Z with the property that if Un ∈ Bn and Un+1 ⊂ Un, n = 1, 2, . . . then ∩nUn 6= ∅. It is easily seen that complete metric spaces and locally compact Hausdorff spaces are pseudo-complete. The cartesian product of any family of pseudo-complete spaces is pseudo-complete, see [19, Theorem 6]. Any pseudo-complete-space is a Baire space, see [19, 5.1]. Proposition A.2. Suppose that the base space T is pseudo-complete and locally paracompact. Then E is a Baire space. Proof. Let {Bn} be sequence of pseudo-bases of T as in the definition of the pseudo- completeness. Let {Gn} be a sequence of open dense subsets of E and G an open subset of E. By the assumption of local paracompactness of T and [5, p. 10] there are an open subset V1 of T , a section ϕ1 of ξ over V1, and a number ε1, 0 < ε1 < 2−1, such that U1 := {x ∈ E p(x) ∈ V1, kx − ϕ1(p(x))k ≤ ε1} ⊂ G ∩ G1. From the assumption of pseudo-completeness of T we infer that there exists W1 ∈ B1 such that W1 ⊂ W1 ⊂ V1. Now there are an open subset V2 ⊂ W1 of T , a section ϕ2 of ξ over V2 ⊂ W1, and a number ε2, 0 < ε2 < 2−2, such that U2 := {x ∈ E p(x) ∈ V2, kx − ϕ2(p(x))k ≤ ε2} ⊂ U1 ∩ G2. There exists W2 ∈ B2 such that W2 ⊂ W2 ⊂ V2. An obvious induction process shows that for every natural number n ≥ 2 there exist an open subset Vn ⊂ Wn−1 of T , a section ϕn of ξ over Vn, and a number εn, 0 < εn < 2−n, such that Un := {x ∈ E p(x) ∈ Vn, kx − ϕn(p(x))k ≤ εn} ⊂ Un−1 ∩ Gn and there exists Wn ∈ Bn such that Wn ⊂ Wn ⊂ Vn. A SELECTION THEOREM FOR BANACH BUNDLES AND APPLICATIONS 33 We have ∩∞ n=1Vn = ∩∞ n=1Wn 6= ∅. Pick t0 ∈ ∩nVn. From {ϕn(t0) − ϕn−1(t0)k ≤ εn−1 < 2n−1 we infer that {ϕn(t0)} is a Cauchy sequence in the Banach space p−1(t0); it converges to a point x0 in this space. The inequality kx0 − ϕn(t0)k ≤ εn, n = 1, 2, . . ., shows that x0 ∈ ∩∞ n=1 Gn and the proof is complete. n=1Un ⊂ G ∩T∞ (cid:3) We have seen in Propositions 3.9 and 3.12 that it is of interest to know if a given Banach bundle has a paracompact bundle space. We turn now to examining some cases when E is paracompact or metrizable. Suppose ξ is a uniform Banach bundle.Then, as remarked above in the paragraph preceding Example 5.4, there exists a Banach space X such that E is a closed subset of T × X. If T is regular and σ − compact or paracompact and perfectly normal then T × X is paracompact by [16, Propositions 4 and 5]. Hence in these cases E is paracompact by [6, p. 218]. With some more work one can obtain the same conclusion when ξ is a locally uniform Banach bundle. Indeed, let {Tα} be a family of closed subsets of T such that {Int(Tα)} is an open cover of T with non-void subsets and the restriction of ξ to each Tα is a uniform Banach bundle. In both cases mentioned above T is paracompact so we may suppose that the family of sets {Tα} is locally finite. Each Tα inherits the properties of T we consider so each p−1(Tα) is paracompact. But {p−1(Tα)} is a locally finite closed cover of E so E is paracompact. Now if ξ is a uniform Banach bundle with a metrizable base space T then E is metrizable by being a subset of a metrizable cartesian product. Again this remains true for a locally uniform Banach bundle ξ. If {Tα} is a locally finite closed cover of T as in the previous paragraph then {p−1(Tα)} is a locally finite closed cover of E with metrizable subspaces. Then a result of Nagata, see [6, p. 204], implies that E is metrizable. Proposition A.3. Let ξ be a continuous Banach bundle with a regular base space T that has a countable base. Moreover, suppose that the Banach space Γb(ξ) is separable. Then E is regular and has a countable base. Proof. Remark first that T is a metrizable space therefore, by Proposition A.1, E is completely regular. Now, say that {Vm} is a countable base of T and {ϕn} is a 34 ALDO J. LAZAR sequence dense in Γb(ξ). For m, n, k ∈ N denote U(m, n, k) := {x ∈ E p(x) ∈ Vm, kx − ϕn(p(x))k < 1/k. We shall show that {U(m, n, k)} is a base of E. Let U ⊂ E be open and x0 ∈ U . The Banach bundle ξ is full so by [5, p. 10] there exist m ∈ N, ϕ ∈ Γb(ξ), and δ > 0 such that p(x0) ∈ Vm, ϕ(p(x0) = x0, and U ′ := {x ∈ E p(x) ∈ Vm, kx − ϕ(p(x))k < δ} ⊂ U. Choose k ∈ N such that 1/k < δ/2 and ϕn such that kϕ − ϕnk < 1/k. It is now routine to check that x0 ∈ U ′ ⊂ U(m, n, k). (cid:3) References 1. E. M. Alfsen and E. G. Effros, Structure in Banach spaces, Part II, Annals of Math. 96 (1972), 129 -- 173. 2. C. W. Baker, A closed graph theorem for Banach bundles, Rocky Mountain J. of Math. 12 (1982), 537 -- 543. 3. R. G. Bartle and L. M. Graves, Mappings between function spaces, Trans. Amer. Math. Soc. 72 (1952), 400 -- 413. 4. E. Behrends, M-Structure and the Banach-Stone Theorem, Springer-Verlag, Berlin, 1979. 5. M. J. Dupr´e and R. M. Gillette, Banach bundles, Banach modules and automorphisms of C ∗-algebras, Pitman Pub Co., New York, 1983. 6. R. Engelking, Outline of general topology, North-Holland, Amsterdam, 1968. 7. J. M. G. Fell, The structure of algebras of operator fields, Acta Math. 106 (1961), 233 -- 280. 8. J. M. G. Fell and R. S. Doran, Representations of ∗-Algebras, Locally Compact Groups, and Banach ∗-Algebraic Bundles, vol. 1, Academic Press, Boston, 1988. 9. G. Gierz, Bundles of Topological Vector Spaces and their Duality, Springer-Verlag, Berlin, 1982. 10. V. I. Gurarii, On openings and inclinations of subspaces of a Banach space (in Russian), Teor. Funktsii, Funktsional. Anal. i Prilozhen. 1 (1965), 194 -- 204. 11. A. E. Gutman, Banach bundles in the theory of lattice-normed spaces (in Russian), Trudy Inst. Mat.,Novosibirsk, v. 29 (1995), 63 -- 211. 12. A. E. Gutman, Banach bundles in the theory of lattice normed spaces I, Siberian Adv. Math. 3 (1993), no. 3, 1 -- 55. 13. A. E. Gutman and A. V. Koptev, On the notion of the dual of a Banach bundle, Siberian Adv. Math. 9 (1999), no. 1, 46 -- 98. 14. M. I. Kadets, Note on the gap between subspaces, Funct. Anal. Appl. 9 (1975), 156 -- 157. 15. A. J. Lazar, Continuous Fields of Postliminal C ∗-Algebras, to appear in Rocky Mountain J. Math. 16. E. Michael, A note of paracompact spaces, Proc. Amer. Math. Soc. 4 (1953), 831 -- 838. 17. E. Michael, Continuous Selections. I, Annals of Mathematics, 63 (1956), 361 -- 382. 18. J. R. Munkres, Topology, second edition, Prentice Hall, 2000. 19. J. C. Oxtoby, Cartesian products of Baire spaces, Fund. Math. 49 (1960/61), 157 -- 166. School of Mathematical Sciences, Tel Aviv University, Tel Aviv 69978, Israel E-mail address: [email protected]
1505.06235
1
1505
2015-05-22T21:57:43
Strengthening of weak convergence for Radon measures in separable Banach spaces
[ "math.FA" ]
We prove in this short report that for arbitrary weak converging sequence of sigma-finite Borelian measures in the separable Banach space there is a compact embedded separable subspace such that this measures not only are concentrated in this subspace but weak converge therein.
math.FA
math
Strengthening of weak convergence for Radon measures in separable Banach spaces. Ostrovsky E., Sirota L. Department of Mathematics and Statistics, Bar-Ilan University, 59200, Ramat Gan, Israel. e-mail: [email protected] Department of Mathematics and Statistics, Bar-Ilan University, 59200, Ramat Gan, Israel. e-mail: [email protected] Abstract. We prove in this short report that for arbitrary weak converging sequence of sigma-finite Borelian measures in the separable Banach space there is a compact embedded separable subspace such that this measures not only are concentrated in this subspace but weak converge therein. Key words and phrases: Separable Banach spaces, random variables (r.v.), distri- butions, measures, support, compact embedded subspace, Skorokhod's representa- tion, weak convergence, Prokhorov's theorem, distance, sharp quantitative estimate, modulus of continuity, Young-Orlicz function, Orlicz spaces. Mathematics Subject Classification (2000): primary 60G17; secondary 60E07; 60G70. 1 Introduction. Notations. Statement of prob- lem. Previous works. Let (X = {x}, · X) be a separable Banach space relative the norm function (Ω, B, P) be a non-trivial probability space, ξ, {ξn}, n = 1, 2, . . . be a · X; sequence of random variables with values in the space X having Borel distributions (1.0) Recall that the Banach subspace (Y = {y}, · Y ) of the space X is named µn(A) = P(ξn ∈ A), µ(A) = P(ξ ∈ A). compact embedded into the space X, write iff the space Y is linear subspace of the space X : Y ⊂ X and the closed unit ball BY = {y : yY ≤ 1} of the space Y is pre-compact set in the space X, i.e. the c.e. ⊂ X, Y 1 closure [BY ]X is compact set in the space X relative the source topology generated by the norm · X. It is known, see [13], [4], [14], that for one Borelian distribution, say µ = µξ, on the space X, probabilistic or at last sigma finite, there exists a separable compact embedded Banach subspace (Y = {y}, · Y ), such that µ(X \ Y ) = 0. Note that this proposition is false in the Linear Topological Spaces instead the Banach space X, see [14]. Obviously, for the enumerable, or more generally dominated family of sigma- finite measures µn there exists a single Banach separable compact embedded into the space X subspace (Y = {y}, · Y ), Y ⊂ X such that (1.1) Suppose now in addition that the sequence {µn} converges weakly as n → ∞ such that for ∀n ⇒ µn(X \ Y ) = 0. to the measure µ in the classical Prokhorov-Skorokhod sense, i.e. arbitrary continuous bounded functional F : X → R lim n→∞ZX F (x) µn(dx) = ZX F (x) µ(dx). (1.2) Write µn X,w → µ. Question 1.1. One can choose either the compact embedded subspace (Y = {y}, · Y ) in (1.1) such that the sequence {µn} is not only concentrated in the space Y, but convergent also in the space Y ? Our aim this short report is to ground the positive answer on this question. 2 Main result. Theorem 2.1. Let (X, · X) be a separable Banach space and µ, µn be an enu- merable set of Borelian probability measures (distributions) on X converging weakly to the measure µ. There exists a separable compact embedded into X Banach sub- ⊂ X, such that all the measures µ, µn are space (Y, · Y ) of the space X : Y concentrated on the Y and moreover weak converge also in the space Y : c.e. µn Y,w → µ. (2.1) Proof. 1. It is sufficient to consider by virtue of universality only the case when X = C[0, 1], see [13], [4], [14]. On the other words, we can and will suppose ξ = 2 ξ(t), ξn = ξn(t), t ∈ [0, 1] are continuous a.e. numerical values random processes. As before, µn C[0,1],w → µ. (2.2) 2. Further, we intend to use the famous Skorokhod's representation theorem, [20], see also [3]. Indeed, there exists a sufficiently rich new probability space (Ω1, B1, P1) and identically with ξ(t), ξn(t) distributed separable r.p. η(t), ηn(t) : ξ(t) dis= η(t), (2.3) such that with the P1 probability one the sequence {ηn(·)} converges uniformly to the random process η(·). Here the symbol dis= denotes the coincidence of distribution. Evidently, the r.p. ξn(t) dis= ηn(t), η, ηn are continuous a.e. The corresponding "accompanying" sequence {η(t), ηn(t)} is said to be Strength- ened Converging Copy of the sequence for the initial one {ξ(t), ξn(t)}, not necessary to be unique, write One can assume {η(t), ηn(t)} SCC= {ξ(t), ξn(t)}. lim n→∞ ζn = 0, ζn := sup t∈[0,1]ηn(t) − η(t) = 0. (2.4) 3. It follows from (2.4) that there exists a deterministic sequence ǫn tending to zero and a random variable τ defined on the new probability space such that ζn ≤ τ · ǫn, (2.5) see e.g. [9], chapter 2, section 3. Moreover, since the sequence of continuous functions η(t), ηn(t) converges uni- formly (a.e.), therefore it is compact set in the space C[0, 1]. It follows from the Arzela-Ascoli theorem that they are equicontinuous, i.e. there exists the (random) non-negative continuous increasing function δ → h(ω, δ), δ ∈ [0, 1] : such that lim δ→0+ h(ω, δ) = 0, ∆(ηn − η, δ) ≤ h(ω, δ) → 0, δ → 0 + . (2.6) (2.7) In what follows ∆(f, δ) will be denote the ordinary module (modulus) of conti- nuity of an uniform continuous function f ∈ C[0, 1]. 4. It follows from the main result of the preprint [13] that there exists a random variable θ defined on the new probability space and deterministic non-negative 3 continuous function δ → g(δ), which takes zero value at the origin: g(0) = g(0+) = 0 such that Then h(ω, δ) ≤ θ · g(δ). ∆(ηn − η, δ) ≤ θ · g(δ). (2.8) (2.9) 5. Let us introduce the following modification of the classical Holder's spaces H o(√g). By definition, H o(√g) consists on all the (continuous) functions f : [0, 1] → R for which lim δ→0+ ∆(f, δ) qg(δ) = 0, with (finite) norm fH o(√g) def= max t∈[0,1]f (t) + sup δ∈(0,1) (2.10) (2.11)   ∆(f, δ) qg(δ) .   These Banach spaces are separable and compact embedded into the space C[0, 1], see, e.g. the monograph [7], chapter 1, where these spaces are used in particular in the theory of non-linear singular integral equation; some another applications, for example, in the theory of CLT in Banach spaces, may be found in [15]. the introduced above space H o(√g). 6. We can suppose without loss of generality that the r.f. η(·), ηn(·) belong to Indeed, as long as the sequence ηn(·) converges uniformly to η(·), it is also com- pact. It remains to repeat the considerations of fourth item (2.9): ∆(ηn, δ) ≤ θ · g(δ), ∆(η, δ) ≤ θ · g(δ), and choose g(δ) := max(g(δ), g(δ)). 7. It follows immediately from the equality (2.9) that the sequence of the r.p. {ηn(·)} converge as n → ∞ to the r.p. η(·) also in the norm H o(√g) with P1 probability one: ηn(·) − η(·)H o(√d) → 0. (2.12) The last equality implies for the source sequence of r.p. ξ(·), ξn(·) weak its This completes the proof of theorem 2.1. distribution convergence in the space H o(√d). 4 Example 2.1. Suppose that the sequence of centered continuous random fields ξn(s), s ∈ S, somehow dependent, where (S = {s}, r) is compact relative certain distance r(·,·) metric space, converges weakly to the continuous Gaussian random field ξ = ξ(s) in the ordinary space of all continuous functions C(S); on the other words, CLT in C(S), [11], chapter 9, [12], chapter 4; uniform CLT [5]. We deduce based on the theorem 2.1 that there exists some modified Holder space H o(g) over (S, r) such that the sequence ξn(·) convergent weakly in the space H o(g), i.e. is subgaussian, as well. 3 Orlicz norm estimates for the tail of random coefficient. The case of the space of continuous functions. In this subsection X = C(S), where the set S = {s} is compact set relative certain distance ρ = ρ(s1, s2), s1, s2 ∈ S. It is interest by our opinion for the practical using, for instance, in the Monte - Carlo method, to estimate the tails of distribution for the r.v. θ in the estimation 2.9. For this purpose assume that the r.v. ν def= sup n ξnC(S) = sup n s∈S ξn(s) sup (3.1) belongs to certain Orlicz space L(Φ) with Luxemburg norm · L(Φ) = · Φ constructed over source probability space; here Φ(·) is some Young-Orlicz function. The detail investigation of Orlicz's spaces may be found in the classical books [10], [18], [19]. We can and will suppose without loss of generality EΦ(ν) = 1. (3.2) Proposition 3.1. It follows in particular from one of results of the recent preprint [16] that if the function Φ satisfies the so-called ∆2 condition: Φ ∈ ∆2, then the scaling function g = g(δ) in (2.9) may be picked such that also θ ∈ L(Φ). The converse predicate is trivially true. In the opposite case the situation is more complicated. Recall that the other Orlicz function Ψ(·) is called weaker than the function Φ, notation Ψ << Φ, if for all positive constant v; v = const > 0 lim u→∞ Ψ(uv) Φ(u) = 0. (3.3) It is alleged that if the function Ψ, Ψ << Φ is given, then the scaling function g = g(δ) = gΦ,Ψ(δ) may be picked such that also θ ∈ L(Ψ). 5 Example 3.1. For instance, if sup n E ξnC(S) p < ∞, ∃p = const ≥ 1, then the scaling function g = g(δ) may be picked such that also Eθp = 1. Example 3.2. Let us consider now the so-called Gaussian centered case, i.e. when the common distribution of the infinite-dimensional vector ~ξ = {ξ, ξ1, ξ2, ξ3, . . .} has a mean zero Gaussian distribution. We have to take Φ(u) = ΦG(u) := eu2/2 − 1. It follows from one of the main results of an article X.Fernique [6], see also [4], [16] that for any choice of the function g(δ) satisfying the relation (2.9) the r.v. θ belongs also to the Orlicz's space L(ΦG), i.e. is subgaussian. The last example show us that the second assertion of proposition 3.1 is in general case improvable. 4 Bernstein's moment convergence for compact embedded subspace. The general case of the arbitrary separable Banach space. The Bernstein's moment convergence for weakly convergent sequence of mea- sures {µn} imply by definition the following integral convergence ZX V (x) µn(dx) → ZX V (x) µ(dx) for certain continuous unbounded functional V : X → R. This problem goes back to S.N.Bernstein [2]; see also [8], [17]. As a rule, in the aforementioned articles the functional V (x) has a form V (x) = xp, x ∈ X, p = const ≥ 2. Let again (X, · X) be separable Banach space and let ~µ = (µ, {µn}) be the family of Borelian probability measures defined on all the Borelian subsets X. Let also V : X → R be continuous functional acting from X to the real axis R : V = V (x), x ∈ X. Definition 4.1. The functional V (·) is named uniform integrable relative the family ~µ, if lim N→∞ sup n Zx:V (x)>N V (x) µn(dx) = 0. (4.1) 6 Lemma 4.1. Suppose that µn w,X integrable relative the family ~µ, µ. We propose → µ and that the functional V (·) is uniform lim n→∞ZX V (x) µn(dx) = ZX V (x) µ(dx). (4.2) Proof is elementary. Let the positive number ǫ > 0 be a given. We introduce the truncated functional, also continuous, VN (x) = V (x), V (x) ≤ N; VN (x) = −N, V (x) < −N; VN (x) = +N, V (x) > N; N = 1, 2, . . . , and denote . ZX V (x) µn(dx) − ZX V (x) µ(dx) (cid:12)(cid:12)(cid:12)(cid:12) We get using the triangle inequality κ ≤ κ1 + κ2 + κ3, where VN (x) dµn(cid:12)(cid:12)(cid:12)(cid:12) VN (x) dµ(cid:12)(cid:12)(cid:12)(cid:12) VN (x) dµ(cid:12)(cid:12)(cid:12)(cid:12) κ := (cid:12)(cid:12)(cid:12)(cid:12) κ1 = (cid:12)(cid:12)(cid:12)(cid:12) ZX V (x) dµn − ZX κ2 = (cid:12)(cid:12)(cid:12)(cid:12) VN (x) dµn − ZX ZX κ3 = (cid:12)(cid:12)(cid:12)(cid:12) ZX V (x) dµ − ZX , , . Since the functional V = V (x) is uniform integrable relative the family ~µ, there exists the value N0 = N0(ǫ) such that for all the values N > N0(ǫ) Further, as long as µn values n > n0 X,w → µ, there is a value n0 = n0(ǫ, N0(ǫ)) so that for all the κ1 < ǫ/3, κ3 ≤ ǫ/3. Totally, n > n0 ⇒ κ < ǫ, Q.E.D. κ2 < ǫ/3. We deduce applying this assertion and the last section the following proposition. Theorem 4.1. Let all the conditions of Lemma 4.1 be satisfied. There exists common support for the all the measures µ, µn compact embedded Banach subspace (Y, ·Y ) such that the functional V (·) is uniform integrable also for these measures inside the new space Y and hence lim n→∞ZY V (y)µn(dy) = ZY V (y)µ(dy). (4.3) 7 5 Concluding remarks. A. Generalization on the non-normed measures. All the references about the random variables ξ, ξn generating the correspondent distributions µ, µn may be eliminated; it is sufficient to consider only the sequence of Borelian (Radon) sigma-finite measures {µn} converging weakly in at the same separable Banach space X to the measure µ, which also also Borelian. Wherein theorem 2.1 remains true, as long as each Borelian sigma-finite measure in equivalent in the Radon-Nikodym sense to the probability distribution. B. Generalization on the arbitrary family of measures. At the same result (theorem 2.1) remains true for an arbitrary dominated and convergent net µα, α ∈ A, α → α0 ∈ A of sigma-finite Borelian measures in the separable Banach space X. C. Open question. V.V.Buldygin in [4] proved that in the probabilistic case µ(X) = µn(X) = 1 the single subspace (Y,·) may be constructed to be reflexive and with continuous differentiable on the unit sphere in the Freshet sense norm. We do not know either or not possible to choose this "good" subspace (Y, · ) such that all the r.v. are concentrate in this space and in addition weakly converges therein. References [1] Bednorz W. Holder continuity of random processes. arXiv:math/0703545v1 [math.PR] 19 Mar 2007. [2] S.N. Bernstein. Quelques remarques sur le theoreme limit Liapounoff. CR. Dokl. Akad. Nauk SSSR, 24, (1939), 3-8. [3] Billingsley, Patrick. (1999). Convergence of Probability Measures. New York: John Wiley and Sons. [4] Buldygin V.V. Supports of probabilistic measures in separable Banach spaces. Theory Probab. Appl., 1984, 29 v.3, pp. 528-532, (in Russian). [5] Dudley R.M. Uniform Central Limit Theorem. Cambridge University Press, 1999. [6] Fernique X. (1975). Regularite des trajectoires des function aleatiores gaussi- ennes. Ecole de Probablite de Saint-Flour, IV 1974, Lecture Notes in Mathe- matic. 480, 1-96, Springer Verlag, Berlin. 8 [7] Gusejnov A.I., Muchtarov Ch.Sh. Introduction to the theory of non-linear singular integral equations. Moskow, Nauka, (1980), (in Russian). [8] P. Hall. The convergence of moments in the martingale central limit theorem. Z. Wahrsch. Verw. Gebiete, 4, (1978), 253-260. [9] Kantorovich L.V., Akilov G.P. Functional Analysis. Moskow, Nauka, (1984), Issue 3, (in Russian). [10] M.A.Krasnoselsky, Ya.B.Rutisky. Convex functions and Orlicz's Spaces. P. Noordhoff LTD, The Netherland, Groningen, 1961. [11] Ledoux M., Talagrand M. (1991) Probability in Banach Spaces. Springer, Berlin, MR 1102015. [12] Ostrovsky E.I. (1999). Exponential estimations for Random Fields and its applications, (in Russian). Russia, Moscow-Obninsk, OINPE. [13] Ostrovsky E.I. About supports of probability measures in separable Banach spaces. Doklady Akadeny Nauk of USSR, Band 255, (1980), N o 6, 836-838. (in Russian). [14] E. Ostrovsky. Support of Borelian measures in separable Banach spaces. arXiv:0808.3248v1 [math.FA] 24 Aug 2008 [15] E. Ostrovsky and L.Sirota. Central Limit Theorem in Holder spaces in the terms of majorizing measures. arXiv:1409.6054v1 [math.PR] 21 Sep 2014 [16] E. Ostrovsky and L.Sirota. Factorable continuity of random fields, with quantitative estimation. arXiv:1505.02839v1 [math.PR] 12 May 2015 [17] Magda Peligrad. The convergence of moments in the Central Limit Theorem for ρ − mixing sequences of random variables. Proceedings of the American Mathematical Society, Volume 101, Number 1, September 1987, 142-148. [18] Rao M.M., Ren Z.D. Theory of Orlicz Spaces. Basel-New York, Marcel Deck- er, (1991). [19] Rao M.M., Ren Z.D. Application of Orlicz Spaces. Basel-New York, Marcel Decker, (2002). [20] Skorokhod A.V. Investigations in the theory of random processes. (1965), Kiev, KSU, (in Russian). [21] Yimin Xiao. Uniform Modulus of Continuity of Random Fields. (2015), Inter- net publication, PDF, Michigan State University. 9
1806.10953
1
1806
2018-06-27T07:03:47
Generalized solutions of variational problems and applications
[ "math.FA", "math.AP" ]
Ultrafunctions are a particular class of generalized functions defined on a hyperreal field $\mathbb{R}^{*}\supset\mathbb{R}$ that allow to solve variational problems with no classical solutions. We recall the construction of ultrafunctions and we study the relationships between these generalized solutions and classical minimizing sequences. Finally, we study some examples to highlight the potential of this approach.
math.FA
math
GENERALIZED SOLUTIONS OF VARIATIONAL PROBLEMS AND APPLICATIONS VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Abstract. Ultrafunctions are a particular class of generalized functions defined on a hyperreal field R∗ ⊃ R that allow to solve variational problems with no classical solutions. We recall the construction of ultrafunctions and we study the relationships between these generalized solutions and classical minimizing sequences. Finally, we study some examples to highlight the potential of this approach. Contents Introduction 1. 1.1. Notations 2. Λ-theory 2.1. Non Archimedean Fields 2.2. The Λ-limit 2.3. Natural extension of sets and functions 2.4. Hyperfinite sets and hyperfinite sums 3. Ultrafunctions 3.1. Definition of ultrafunctions 3.2. The splitting of an ultrafunction 3.3. The Gauss divergence theorem 3.4. Ultrafunctions and distributions 4. Properties of ultrafunction solutions 5. Applications 5.1. Sign-perturbation of potentials 5.2. The singular variational problem References 1. Introduction 1 2 3 3 3 5 7 7 7 10 12 13 14 17 17 19 22 It is nowadays very well known that, in many circumstances, the needs of a theory require the intro- duction of generalized functions. Among people working in partial differential equations, the theory of distributions of L. Schwartz is the most commonly used, but other notions of generalized functions have been introduced, e.g. by J.F. Colombeau [15] and M. Sato [21, 22]. Many notions of generalized functions are based on non-Archimedean mathematics, namely mathematics handling infinite and/or infinitesimal quantities. Such an approach presents several positive features, the main probably being the possibility of treating distributions as non-Archimedean set-theoretical functions (under the limitations imposed by Schwartz' result). This allows to easily introduce nonlinear concepts, such as products, into distribution theory. Moreover, a theory which includes infinitesimals and infinite quantities makes it possible to easily construct new models, allowing in this way to study several problems which are difficult even to formalize in classical mathematics. This has led to applications in various field, including several topics in analysis, geometry and mathematical physic (see e.g. [16, 18] for an overview in the case of Colombeau functions and their recent extension, called generalized smooth functions). 2010 Mathematics Subject Classification. 03H05, 26E35, 28E05, 46S20. Key words and phrases. Ultrafunctions, Non Archimedean Mathematics, Non Standard Analysis, Delta function. M. Squassina is a member of the Gruppo Nazionale per l'Analisi Matematica, la Probabilit`a e le loro Applicazioni (GNAMPA) and of Istituto Nazionale di Alta Matematica (INdAM). L. Luperi Baglini has been supported by grants M1876- N35, P26859-N25 and P 30821-N35 of the Austrian Science Fund FWF. 1 2 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA In this paper we deal with ultrafunctions, which are a kind of generalized functions that have been introduced recently in [1] and developed in [2, 4–11]. Ultrafunctions are a particular case of non-Archimedean generalized functions that are based on the hyperreal field R∗, namely the numerical field on which nonstandard analysis1 is based. No prior knowledge of nonstandard analysis is requested to read this paper: we will introduce all the nonstandard notions that we need via a new notion of limit, called Λ-limit (see [10] for a complete introduction to this notion and its relationships with the usual nonstandard analysis). The main peculiarity of this notion of limit is that it allows us to make a very limited use of formal logic, in constrast with most usual nonstandard analysis introductions. Apart from being framed in a non-Archimedean setting, ultrafunctions have other peculiar properties that will be introduced and used in the following: • every ultrafunction can be splitted uniquely (in a sense that will be precised in Section 3.9) as the sum of a classical function and a purely non-Archimedean part; • ultrafunctions extend distributions, in the sense that every distribution can be identified with an ultrafunction; in particular, this allows to perform nonlinear operations with distribution; • although being generalized functions, ultrafunctions shares many properties of C1 functions, like e.g. Gauss' divergence theorem. Our goal is to introduce all the aforementioned properties of ultrafunctions so to be able to explain how they can be used to solve certain classical problems that do not have classical solutions; in particular, we will concentrate on singular problems arising in calculus of variations and in relevant applications (see, e.g., [18] and references therein for other approaches to these problems based on different notions of generalized functions). The paper is organized as follows: in Section 2 we introduce the notion of Λ-limit, and we explain how to use it to construct all the non-Archimedean tools that are needed in the rest of the paper, in partic- ular how to construct the non-Archimedean field extension R∗ of R and what the notion of "hyperfinite" means. In Section 3 we define ultrafunctions, and we explain how to extend derivatives and integrals to them. All the properties of ultrafunctions needed later on are introduced in this section: we show how to split an ultrafunction as the sum of a standard and a purely non-Archimedean part, how to extend Gauss' divergence theorem and how to identify distributions with certain ultrafunctions. In Section 4 we present the main results of the paper, namely we show that a very large class of classical problems admits a generalized ultrafunction solutions. We study the main properties of these generalized solutions, concentrating in particular on the relationships between ultrafunction solutions and classical minimizing sequences for variational problems. Finally, in Section 5 we present two examples of applications of our methods: the first is the study of a variational problem related with the sign-perturbation of potentials, the second is a singular variation problem related with sign-changing boundary conditions. The first part of this paper contains some overlap with other papers on ultrafunctions, but this fact is necessary to make it self-contained and to make the reader comfortable with it. 1.1. Notations. If X is a set and Ω is a subset of RN , then • P (X) denotes the power set of X and Pfin (X) denotes the family of finite subsets of X; • F (X, Y ) denotes the set of all functions from X to Y and F (Ω) = F (Ω, R) . • C (Ω) denotes the set of continuous functions defined on Ω ⊂ RN ; • Ck (Ω) denotes the set of functions defined on Ω ⊂ RN which have continuous derivatives up to the order k (sometimes we will use the notation C 0(Ω) instead of C (Ω) ); • H k,p (Ω) denotes the usual Sobolev space of functions defined on Ω ⊂ RN • if W (Ω) is any function space, then Wc (Ω) will denote de function space of functions in W (Ω) having compact support; • C0 (Ω ∪ Ξ) , Ξ ⊆ ∂Ω, denotes the set of continuous functions in C (Ω ∪ Ξ) which vanish for x ∈ Ξ • D (Ω) denotes the set of the infinitely differentiable functions with compact support defined on Ω ⊂ RN ; D′ (Ω) denotes the topological dual of D (Ω), namely the set of distributions on Ω; 1We refer to Keisler [17] for a very clear exposition of nonstandard analysis. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 3 • if A ⊂ X is a set, then χA denotes the characteristic function of A; • supp(f ) = supp∗(f ) where supp is the usual notion of support of a function or a distribution; : x ∼ y} where x ∼ y means that x − y is infinitesimal; • mon(x) = {y ∈(cid:0)RN(cid:1)∗ • ∀a.e. x ∈ X means "for almost every x ∈ X"; • if a, b ∈ R∗, then – [a, b]R∗ = {x ∈ R∗ : a ≤ x ≤ b}; – (a, b)R∗ = {x ∈ R∗ : a < x < b}; • if W is a generic function space, its topological dual will be denoted by W ′ and the pairing by h·, ·iW ; • if E is any set, then E will denote its cardinality. 2. Λ-theory In this section we present the basic notions of Non Archimedean Mathematics (sometimes abbreviated as NAM) and of Nonstandard Analysis (sometimes abbreviated as NSA) following a method inspired by [3] (see also [1] and [4]). 2.1. Non Archimedean Fields. Here, we recall the basic definitions and facts regarding non-Archimedean fields. In the following, K will denote a totally ordered infinite field. We recall that such a field contains (a copy of) the rational numbers. Its elements will be called numbers. Definition 2.1. Let K be an ordered field. Let ξ ∈ K. We say that: • ξ is infinitesimal if, for all positive n ∈ N, ξ < 1 n ; • ξ is finite if there exists n ∈ N such as ξ < n; • ξ is infinite if, for all n ∈ N, ξ > n (equivalently, if ξ is not finite). Definition 2.2. An ordered field K is called Non-Archimedean if it contains an infinitesimal ξ 6= 0. It is easily seen that infinitesimal numbers are actually finite, that the inverse of an infinite number is a nonzero infinitesimal number, and that the inverse of a nonzero infinitesimal number is infinite. Definition 2.3. A superreal field is an ordered field K that properly extends R. It is easy to show, due to the completeness of R, that there are nonzero infinitesimal numbers and infinite numbers in any superreal field. Infinitesimal numbers can be used to formalize a new notion of closeness, according to the following Definition 2.4. We say that two numbers ξ, ζ ∈ K are infinitely close if ξ − ζ is infinitesimal. In this case, we write ξ ∼ ζ. Clearly, the relation ∼ of infinite closeness is an equivalence relation and we have the following Theorem 2.5. If K is a totally ordered superreal field, every finite number ξ ∈ K is infinitely close to a unique real number r ∼ ξ, called the standard part of ξ. Given a finite number ξ, we denote its standard part by st(ξ), and we put st(ξ) = ±∞ if ξ ∈ K is a positive (negative) infinite number. In Definition 2.16 we will see how the notion of standard part can be generalized to any Hausdorff topological space. Definition 2.6. Let K be a superreal field, and ξ ∈ K a number. The monad of ξ is the set of all numbers that are infinitely close to it mon(ξ) := {ζ ∈ K : ξ ∼ ζ}. 2.2. The Λ-limit. Let U be an infinite set of cardinality bigger that the continuum and let L = Pfin(U ) be the family of finite subsets of U . Notice that (L, ⊆) is a directed set. We add to L a point at infinity Λ /∈ L, and we define the following family of neighbourhoods of Λ : where U is a fine ultrafilter on L, namely a filter such that {{Λ} ∪ Q Q ∈ U}, 4 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA • for every A, B ⊆ L, if A ∪ B = L then A ∈ U or B ∈ U; • for every λ ∈ L the set Q(λ) := {µ ∈ L λ ⊑ µ} ∈ U. We will refer to the elements of U as qualified sets. A function ϕ : L → E defined on a directed set E is called net (with values in E). If ϕ(λ) is a real net, we have that if and only if lim λ→Λ ϕ(λ) = L ∀ε > 0, ∃Q∈U : ∀λ∈Q, ϕ(λ) − L < ε. As usual, if a property P (λ) is satisfied by any λ in a neighbourhood of Λ we will say that it is eventually satisfied. Proposition 2.7. If the net ϕ(λ) takes values in a compact set K, then it is a converging net. Proof. Suppose that the net ϕ(λ) has a converging subnet to L ∈ R. We fix ε > 0 arbitrarily and we have to prove that Qε ∈ U where Qε = {λ ∈ L ϕ(λ) − L < ε} . We argue indirectly and we assume that Qε /∈ U. Then, by the definition of ultrafilter, N = L\Qε ∈ U and hence ∀λ ∈ N, ϕ(λ) − L ≥ ε. This contradict the fact that ϕλ has a subnet which converges to L. (cid:3) Proposition 2.8. Assume that ϕ : L → E where E is a first countable topological space; then if there exists a sequence {λn} in L such that lim λ→Λ ϕ(λ) = x0, lim n→∞ ϕ(λn) = x0 We refer to the sequence ϕn := ϕ(λn) as a subnet of ϕ(λ). Proof. Let {An n ∈ N} be a countable basis of open neighbourhoods of x0. For every n ∈ N the set In := {λ ∈ L ϕ(λ) ∈ An} is qualified. Hence Jn :=Tj≤n Ij 6= ∅. Let λn ∈ Jn. Then the sequence {λn}n∈N has trivially the desired property: for every n ∈ N, for every m ≥ n we have that ϕ (λm) ∈ An. (cid:3) Example 2.9. Let ϕ : L → V be a net with values in a bounded subset of a reflexive Banach space equipped with the weak topology; then v := lim λ→Λ ϕ(λ), is uniquely defined and there exists a sequence n 7→ ϕ(λn) which converges to v. Definition 2.10. The set of the hyperreal numbers R∗ ⊃ R is a set equipped with a topology τ such that • every net ϕ : L → R has a unique limit in R∗ if L and R∗ are equipped with the Λ and the τ topology respectively; • R∗ is the closure of R with respect to the topology τ ; • τ is the coarsest topology which satisfies the first property. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 5 The existence of such R∗ is a well known fact in NSA. The limit ξ ∈ R∗ of a net ϕ : L → R with respect to the τ topology, following [1], is called the Λ-limit of ϕ and the following notation will be used: (2.1) ξ = lim λ↑Λ ϕ(λ); namely, we shall use the up-arrow "↑" to remind that the target space is equipped with the topology τ . Given we set (2.2) and (2.3) ξ := lim λ↑Λ ϕ(λ), η := lim λ↑Λ ψ(λ), ξ + η := lim λ↑Λ (ϕ(λ) + ψ(λ)) , ξ · η := lim λ↑Λ (ϕ(λ) · ψ(λ)) . Then the following well known theorem holds: Theorem 2.11. The definitions (2.2) and (2.3) are well posed and R∗, equipped with these operations, is a non-Archimedean field. Remark 2.12. We observe that the field of hyperreal numbers is defined as a sort of completion of real numbers. In fact R∗ is isomorphic to the ultrapower where RL/I I = {ϕ : L → R ϕ(λ) = 0 eventually} The isomorphism resembles the classical one between real numbers and equivalence classes of Cauchy sequences: this method is surely known to the reader for the construction of the real numbers starting from the rationals. 2.3. Natural extension of sets and functions. To develop applications, we need to extend the notion of Λ-limit to sets and functions (but also to differential and integral operators). This will allow to enlarge the notions of variational problem and of variational solution. Λ-limits of bounded nets of mathematical objects in V∞(R) can be defined by induction (a net ϕ : L → V∞(R) is called bounded if there exists n ∈ N such that ∀λ ∈ L, ϕ(λ) ∈ Vn(R)). To do this, consider a net (2.4) ϕ : L → Vn(R). Definition 2.13. For n = 0, limλ↑Λ ϕ(λ) is defined by (2.1); so by induction we may assume that the limit is defined for n − 1 and we define it for the net (2.4) as follows: ϕ(λ) =(cid:26)lim λ↑Λ ψ(λ) ψ : L → Vn−1(R), ∀λ ∈ L, ψ(λ) ∈ ϕ(λ)(cid:27) . lim λ↑Λ A mathematical entity (number, set, function or relation) which is the Λ-limit of a net is called internal. Definition 2.14. If ∀λ ∈ L, Eλ = E ∈ V∞(R), we set limλ↑Λ Eλ = E∗, namely E∗ is called the natural extension of E. E∗ :=(cid:26)lim λ↑Λ ψ(λ) ψ(λ) ∈ E(cid:27) . Notice that, while the Λ-limit of a constant sequence of numbers gives this number itself, a constant sequence of sets gives a larger set, namely E∗. In general, the inclusion E ⊆ E∗ is proper. Given any set E, we can associate to it two sets: its natural extension E∗ and the set Eσ, where (2.5) Eσ := {X ∗ X ∈ E} . Clearly Eσ is a copy of E; however it might be different as set since, in general, X ∗ 6= X. 6 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Remark 2.15. If ϕ : L → X is a net with value in a topological space we have the usual limit lim λ→Λ ϕ(λ), which, by Proposition 2.7, always exists in the Alexandrov compactification X ∪ {∞}. Moreover we have the Λ-limit, that always exists and it is en element of X ∗. In addition, the Λ-limit of a net is in X σ if and only if ϕ is eventually constant. If X = R and both limits exist, then (2.6) ϕ(λ) = st(cid:18)lim λ↑Λ ϕ(λ)(cid:19) . lim λ→Λ The above equation suggests the following definition. Definition 2.16. If X is topological space equipped with a Hausdorff topology, and ξ ∈ X ∗ we set if there is a net ϕ : L → X converging in the topology of X and such that StX (ξ) = lim λ→Λ ϕ(λ), and StX (ξ) = ∞ otherwise. By the above definition we have that Definition 2.17. Let ξ = lim λ↑Λ ϕ(λ), ϕ(λ) = StX(cid:18)lim λ↑Λ ϕ(λ)(cid:19) . lim λ→Λ fλ : Eλ → R, λ ∈ L, be a net of functions. We define a function f :(cid:18)lim λ↑Λ Eλ(cid:19) → R∗ as follows: for every ξ ∈ (limλ↑Λ Eλ) we set f (ξ) := lim λ↑Λ fλ (ψ(λ)) , where ψ(λ) is a net of numbers such that A function which is a Λ-limit is called internal. In particular if, ∀λ ∈ L, ψ(λ) ∈ Eλ and lim λ↑Λ ψ(λ) = ξ. we set fλ = f, f : E → R, f ∗ = lim λ↑Λ fλ. f ∗ : E∗ → R∗ is called the natural extension of f. If we identify f with its graph, then f ∗ is the graph of its natural extension. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 7 2.4. Hyperfinite sets and hyperfinite sums. Definition 2.18. An internal set is called hyperfinite if it is the Λ-limit of a net ϕ : L → F where F is a family of finite sets. For example, if E ∈ V∞(R), the set is hyperfinite. Notice that (λ ∩ E) λ↑Λ eE = lim Eσ ⊂ eE ⊂ E∗ A := lim λ↑Λ Aλ, Xa∈A a = lim a. λ↑Λ Xa∈Aλ so, we can say that every standard set is contained in a hyperfinite set. It is possible to add the elements of a hyperfinite set of numbers (or vectors) as follows: let be a hyperfinite set of numbers (or vectors); then the hyperfinite sum of the elements of A is defined in the following way: In particular, if Aλ =(cid:8)a1(λ), ..., aβ(λ)(λ)(cid:9) with β(λ) ∈ N, then setting β(λ) ∈ N∗, β = lim λ↑Λ we use the notation βXj=1 aj = lim λ↑Λ β(λ)Xj=1 aj(λ). 3. Ultrafunctions 3.1. Definition of ultrafunctions. We start by introducing the notion of hyperfinite grid: is called hyperfinite grid. Definition 3.1. A hyperfinite set Γ such that RN ⊂ Γ ⊂(cid:0)RN(cid:1)∗ From now on we assume that Γ has been fixed once forever. Notice that, by definition, RN ⊆ Γ, and the following two simple (but useful) properties of Γ can be easily proven via Λ-limits: • for every x ∈ RN there exists y ∈ Γ ∩ mon(x) so that x 6= r; • there exists a hyperreal number ρ ∼ 0, ρ > 0 such that d(x, y) ≥ ρ for every x, y ∈ Γ, x 6= y. Definition 3.2. A space of grid functions is a family G(RN ) of internal functions u : Γ → R∗ defined on a hyperfinite grid Γ. If E ⊂ RN , then G(E) will denote the restriction of the grid functions to the set E∗ ∩ Γ. Let E be any set in RN . To every internal function u ∈ F(E)∗, it is possible to associate a grid function by the "restriction" map (3.1) defined as follows: ◦ : F(E)∗ → G(E) ∀x ∈ E∗ ∩ Γ, u◦(x) := u∗(x); moreover, if f ∈ F(E), for short, we use the notation (3.2) So every function f ∈ F(E), can be uniquely extended to a grid function f ◦ ∈ G(E). f ◦(x) := (f ∗)◦ (x). In many problem we have to deal with functions defined almost everywhere in Ω, such as 1/x. Thus it is useful to give a "rule" which allows to define a grid function for every x ∈ Γ: 8 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Definition 3.3. If a function f is defined on a set E ⊂ RN , we put where ∀a ∈ Γ the grid fuction σa is defined as follows: σa(x) := δax. If E ⊂ RN is a measurable set, we define the "density function" of E as follows: (3.3) f ◦(x) = Xa∈Γ∩E ∗ f ∗(a)σa(x), m(Bη(x)) (cid:19) , θE(x) = st(cid:18) m(Bη(x) ∩ E∗) where η is a fixed infinitesimal and m is the Lebesgue measure. Clearly θE(x) is a function whose value is 1 in int(E) and 0 in RN \ E; moreover, it is easy to prove that θE(x) is a measurable function and we have that θE(x)dx = m(E) whenever m(E) < ∞; also, if E is a bounded open set with smooth boundary, we have that θE(x) = 1/2 for every x ∈ ∂E. Now let V (RN ) be a vector space such that D(RN ) ⊂ V (RN ) ⊂ L 1(RN ). Definition 3.4. A space of ultrafunctions V ◦(RN ) modelled over the space V (RN ) is a space of grid functions such that there exists a vector space VΛ(RN ) ⊂ V ∗(RN ) such that the map2 ◦ : VΛ(RN ) → V ∗(RN ) is a R∗-linear isomophism. From now on, we assume that V (RN ) satisfies the following assumption: if Ω is a bounded open set such that mN −1(∂Ω) < ∞ and f ∈ C0(RN ), then f θΩ ∈ V (RN ). Next we want to equip V ◦(RN ) with the two main operations of calculus: the integral and the derivative. Definition 3.5. The pointwise integral " : V ◦(RN ) → R∗ is a linear functional which satisfies the following properties: (1) ∀u ∈ VΛ(RN ) (3.4) " u◦(x) dx = u(x)dx; (2) there exists a ultrafunction d : Γ → R∗ such that ∀x ∈ Γ, d(x) > 0 and ∀u ∈ V ◦(RN ), If E ⊂ RN is any set, we use the obvious notation " u(x) dx =Xa∈Γ u(x) dx := Xa∈Γ∩E ∗ "E u(a)d(a). u(a)d(a) " σa(x) dx = d(a). Few words to discuss the above definition. Point 2 says that the poinwise integral is nothing else but a hyperfinite sum. Since d(x) > 0 every non-null positive ultrafunction has a strictly positive integral. In particular if we denote by σa(x) the ultrafunctions whose value is 1 for x = a and 0 otherwise, we have that 2V ∗(E) is a shorthand notation for [V (E)]∗ GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 9 The pointwise integral allows us to define the following scalar product: (3.5) " u(x)v(x)dx =Xa∈Γ u(a)v(a)d(a). From now on, the norm of an ultrafunction will be defined as kuk =(cid:18)" u(x)2 dx(cid:19) 1 2 . Now let us examine the point 1 of the above definition. If we take f ∈ C0 VΛ(RN ) and hence comp(RN ), we have that f ∗ ∈ Thus the pointwise integral is an extension of the Riemann Integral defined on C0 we take a bounded open set Ω such that m(∂Ω) = 0, then we have that comp(RN ). However, if " f ◦(x) dx = f (x)dx. Ω f (x)dx = Ω f (x)dx; however, the pointwise integral cannot have this property; in fact "Ω f ◦(x)dx −"Ω f ◦(x)dx = "∂Ω f ◦(x)dx > 0 since ∂Ω 6= ∅. In particular, if Ω is a bounded open set with smooth boundary and f ∈ C0(RN ), then "Ω f ◦(x) dx = " f ◦(x)χΩ(x) dx Ω(x) dx − = " f ◦(x)θ◦ = Ω f (x) dx − 1 2 "∂Ω ∂Ω(x) dx 1 2 " f ◦(x)χ◦ f ◦(x) dx, and similarly "Ω f ◦(x) dx = Ω f (x) dx + 1 2 "∂Ω f ◦(x) dx; of course, the term 1 problems. 2 › f ◦(x)χ∂E(x) dx is an infinitesimal number and it is relevant only in some particular Definition 3.6. The ultrafunction derivative Di : V ◦(RN ) → V ◦(RN ) is a linear operator which satisfies the following properties: (1) ∀f ∈ C1(RN ), and ∀x ∈ (RN )∗, x finite, (3.6) (2) ∀u, v ∈ V ◦(RN ), Dif ◦(x) = ∂if ∗(x); " Diuv dx = −" uDiv dx; (3) if Ω is a bounded open set with smooth boundary, then ∀v ∈ V ◦, " DiθΩv dx = − ∗ ∂Ω v (ei · nE) dS where nE is the unit outer normal, dS is the (n − 1)-dimensional measure and (e1, ...., eN ) is the canonical basis of RN . (4) the support3 of Diσa is contained in mon(a) ∩ Γ. 3If u is an ultrafunction the support of u is the set {x ∈ Γ u(x) 6= 0}. 10 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Let us comment the above definition. Point (1) implies that ∀f ∈ C1(RN ), and ∀x ∈ RN , (3.7) Dif ◦(x) = ∂if (x); namely the ultrafunction derivative concides with the usual partial derivative whenever f ∈ C1(RN ). The meaning of point (2) is clear; we remark that this point is very important in comparing ultrafunctions with distributions. Point (3) says that DiθΩ is an ultrafunction whose support is contained in ∂Ω ∩ Γ; it can be considered as a signed measure concentrated on ∂Ω. Point (4) says that the ultrafunction derivative, as well as the usual derivative or the distributional derivative, is a local operator, namely if u is an ultrafunction whose support is contained in a compact set K with K ⊂ Ω, then the support of Diu is contained in Ω∗. Moreover, property (4) implies that the ultrafunction derivative is well defined in V ◦(Ω) for any open set Ω by the following formula: Diu(x) = Xa∈Γ∩Ω∗ u(a)Diσa(x). Remark 3.7. If u ∈ V ◦(Ω) and ¯u is an ultrafunction in V ◦(Ω) such that ∀x ∈ Ω, u(x) = ¯u(x), then, by point (3), ∀x ∈ Ω∗ such that mon(x) ⊂ Ω∗ we have that Diu(x) = Di ¯u(x); however, this property fails for some x ∼ ∂Ω∗. In fact the support of Diσa is contained in mon(a) ∩ Γ, but not in {a}. Theorem 3.8. There exists a ultrafuctions space V ◦(RN ) which admits a pointwise integral and a ultra- function derivative as in Def. 3.5 and 3.6. Proof. In [2] there is a construction of a space VΛ(RN ) which satisfies the desired properties. The conclusion follows taking V ◦(RN ) = {u◦ : u ∈ VΛ} . (cid:3) 3.2. The splitting of an ultrafunction. In many applications, it is useful to split an ultrafunction u in a part w◦ which is the canonical extension of a standard function w and a part ψ which is not directly related to any classical object. If u ∈ V ◦(Ω), we set and S = {x ∈ Ω u(x) is infinite} w(x) =(st(u(x)) 0 if x ∈ Ω \ S; if x ∈ S. We will refer to S as to the singular set of the ultrafunction u. Definition 3.9. For every ultrafunction u consider the splitting u = w◦ + ψ where • w = wΩ\S and w◦, which is defined by Def. 3.3, is called the functional part of u; • ψ := u − w◦ is called the singular part of u. Notice that w◦, the functional part of u, may assume infinite values for some x ∈ Ω∗ \ S∗, but they are determined by the values of w which is a standard function defined on Ω. Example 3.10. Take ε ∼ 0, and In this case • w(x) =(log(x2) = 2 log(x) 0 u(x) = log(x2 + ε2). if x 6= 0 if x = 0 GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 11 • ψ(x) =(log(x2 + ε2) − log(x2) = log(cid:16)1 + ε2 log(ε2) x2(cid:17) if x 6= 0 if x = 0 • S := {0} . We conclude this section with the following trivial proposition which, nevertheless, is very useful in applications: Proposition 3.11. Take a Banach space W such that D(Ω) ⊂ W ⊂ L1(Ω). Assume that {un} ⊆ V (Ω) is a sequence which converges weakly in W and pointwise to a function w; then, if we set u := (cid:18)lim λ↑Λ uλ(cid:19)◦! , we have that where moreover, if u = w◦ + ψ, ∀v ∈ W, " ψv dx ∼ 0; lim n→∞ kun − wkW = 0 then kψkW ∼ 0. Proof. As a consequence of the pointwise convergence of {un} to w we have that ∀a ∈ Γ u(a) ∼ w◦(a). In particular, ∀a ∈ Γ ψ(a) ∼ 0. As Γ is hyperfinite, the set {ψ(a) a ∈ Γ} has a max η ∼ 0. Hence for every v ∈ W we have as η ∼ 0 and ´ vdx ∈ R. For the second statement let us notice that (cid:12)(cid:12)(cid:12)(cid:12)" ψv dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ " ψ v dx ≤ η" v dx ∼ η vdx ∼ 0, kψkW = ku − w◦kW =(cid:18)lim λ↑Λ(cid:13)(cid:13)uλ − w(cid:13)(cid:13)W(cid:19)◦ ∼ 0 (cid:3) as limn→∞ kun − wkW = 0. An immediate consequence of Prop. 3.11 is the following: Corollary 3.12. If w ∈ L1(Ω) then " w◦(x)dx ∼ w(x)dx. Proof. Since V (Ω) is dense in L1(Ω) there is a sequence un ∈ V (Ω) which converges strongly to w in L1(Ω). Now set By Prop. 3.11, we have that with kψkL1 ∼ 0. Then u :=(cid:18)lim λ↑Λ uλ(cid:19)◦ . u = w◦ + ψ " u(x)dx ∼ " w◦(x)dx. On the other hand, since u ∈ V ◦(Ω), by Def. 3.5.(1), " u(x)dx = ∗ u◦(x)dx = lim λ↑Λ uλdx undx = w(x)dx. ∼ lim n→Λ Ω (cid:3) 12 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA 3.3. The Gauss divergence theorem. First of all, we fix the notation for the main differential operators: • ∇ = (∂1, ..., ∂N ) will denote the usual gradient of standard functions; • ∇∗ = (∂∗ • D = (D1, ..., DN ) will denote the canonical extension of the gradient in the sense of ultrafunctions. N ) will denote the natural extension of internal functions; 1 , ..., ∂∗ Next let us consider the divergence: • ∇∗ · φ = ∂∗ • ∇·φ = ∂1φ1 + ...+ ∂N φN will denote the usual divergence of standard vector fields φ ∈(cid:2)C1(RN )(cid:3)N N φN will denote the divergence of internal vector fields φ ∈(cid:2)C1(RN )∗(cid:3)N • D·φ = D1φ1 +...+DN φN will denote the divergence of vector valued ultrafuction φ ∈(cid:2)V ◦(RN )∗(cid:3)N . And finally, we can define the Laplace operator of an ultrafunction u ∈ V ◦(Ω) as the only ultrafunction △◦u ∈ V ◦(Ω) such that 1 φ1 + ... + ∂∗ ; ; where ∀v ∈ V ◦ 0 (Ω), " △◦uvdx = −" Du · Dv dx, 0 (Ω) :=(cid:8)v ∈ V ◦(Ω) ∀x ∈ ∂Ω ∩ Γ, v(x) = 0(cid:9) . V ◦ By definition 3.6.(3), for any bounded open set Ω with smooth boundary, and by definition 3.6.(2) so that " DiθΩv dx = − ∗ ∂Ω v (ei · nE) dS, " DiθΩv dx = −" Div θΩdx " Div θΩdx = ∗ ∂Ω v (ei · nΩ) dS. Now, if we take a vector field φ = (v1, ..., vN ) ∈(cid:2)V ◦(RN )(cid:3)N " D · φ θΩ dx = ∗ (3.8) ∂Ω φ · nΩ dS. , by the above identity, we get Now, if φ ∈ C1, by definition 3.6.(1), we get the divergence Gauss theorem: Ω ∇ · φ dx = ∂Ω φ · nE dS. Then, (3.8) is a generalization of the Gauss' theorem which makes sense for any bounded open set Ω with smooth boundary and every vectorial ultrafunction φ. Next, we want to generalize Gauss' theorem to any subset of A ⊂ RN . It is well known that, for any bounded open set Ω with smooth boundary, the distributional derivative ∇θΩ is a vector valued Radon measure and we have that h∇θΩ, 1i = mN −1(∂Ω) Then, the following definition is a natural generalization: Definition 3.13. If A is a measurable subset of RN , we set mN −1(∂Ω) := " Dθ◦ A dx and ∀v ∈ V ◦(RN ), (3.9) "∂A v(x) dS := " v(x) Dθ◦ A dx. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 13 Remark 3.14. Notice that In fact the left hand term has been defined as follows: while the right hand term is v(x) dx. v(x) Dθ◦ A(x) d(x), "∂A v(x) dS 6= "∂A "∂A v(x) dS =Xx∈Γ v(x) dx = Xx∈Γ∩∂A∗ "∂A v(x) d(x); in particular if ∂A is smooth and v(x) is bounded, ›∂A v(x) dx is an infinitesimal number. Theorem 3.15. If A is an arbitrary measurable subset of RN , we have that (3.10) where Proof. By Definition 3.6.(3) A dx = "∂A " D · φ θ◦ A(x) =(− Dθ◦ Dθ◦ 0 n◦ A(x) A(x) φ · n◦ A(x) dS, if Dθ◦ if Dθ◦ A(x) 6= 0; A(x) = 0. then, using the definition of n◦ A(x) and (3.9), the above formula can be written as follows: " D · φ θ◦ Adx = −" φ · Dθ◦ Adx, " D · φ θ◦ Adx = " φ · n◦ A Dθ◦ A dx = "∂A φ · n◦ A dS. (cid:3) 3.4. Ultrafunctions and distributions. One of the most important properties of the ultrafunctions is that they can be seen (in some sense that we will make precise in this section) as generalizations of the distributions. Definition 3.16. The space of generalized distribution on Ω is defined as follows: D ′ G(Ω) = V ◦(Ω)/N, where N =(cid:26)τ ∈ V ◦(Ω) ∀ϕ ∈ D(Ω), τ ϕ dx ∼ 0(cid:27) . The equivalence class of u in V ◦(Ω) will be denoted by Definition 3.17. Let [u]D be a generalized distribution. We say that [u]D is a bounded generalized dis- GB(Ω) the set of bounded generalized tribution if ∀ϕ ∈ D(Ω), ´ uϕ∗ dx is finite. We will denote by D ′ distributions. [u]D . We have the following result. Theorem 3.18. There is a linear isomorphism Φ : D ′ GB(Ω) → D ′(Ω) such that, for every [u] ∈ D ′ GB(Ω) and for every ϕ ∈ D(Ω) hΦ ([u]D ) , ϕiD(Ω) = st(cid:18)" u ϕ∗ dx(cid:19) . 14 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Proof. For the proof see e.g. [7]. (cid:3) From now on we will identify the spaces D ′ GB(Ω) and D ′(Ω); so, we will identify [u]D with Φ ([u]D ) and we will write [u]D ∈ D ′(Ω) and h[u]D , ϕiD(Ω) := hΦ[u]D , ϕi = st(cid:18)" u ϕ∗ dx(cid:19) . If f ∈ C0 comp(Ω) and f ∗ ∈ [u]D , then ∀ϕ ∈ D(Ω), h[u]D , ϕiD(Ω) = st(cid:18) ∗ u ϕ∗ dx(cid:19) = st(cid:18) ∗ f ∗ϕ∗dx(cid:19) = f ϕ dx. Remark 3.19. The set V ◦(Ω) is an algebra which extends the algebra of continuous functions C0(RN ). If we identify a tempered distribution4 T = ∂mf with the ultrafunction Dmf ◦, we have that the set of tempered distributions S ′ is contained in V ◦(Ω). However the Schwartz impossibility theorem is not violated as (V ◦(Ω), +, · , D) is not a differential algebra since the Leibnitz rule does not hold for some pairs of ultrafunctions. See also [7]. 4. Properties of ultrafunction solutions The problems that we want to study with ultrafunctions have the following form: minimize a given functional J on V (Ω) subjected to certain restrictions (e.g., some boundary constrictions, or a minimization on a proper vector subspace of V (Ω)). This kind of problems can be studied in ultrafunctions theory by means of a modification of the Faedo-Galerkin method, based on standard approximations by finite dimensional spaces. The following is a (maybe even too) general formulation of this idea. Theorem 4.1. Let W (Ω) 6= ∅ be a vector subspace of V (Ω). Let F = {f : V (Ω) → R ∀E finite dimensional vector subspace of W (Ω) ∃u ∈ E f (u) = min v∈E f (v)}. Then every F ∈ F ∗ has a minimizer in WΛ(Ω). Proof. Let F = limλ↑Λ fλ, with fλ ∈ F for every λ ∈ L. By hypothesis, for every λ ∈ L there exists uλ ∈ Wλ := Span(W ∩ λ) that minimizes fλ on Wλ. Then u = limλ↑Λ uλ minimizes F on limλ↑Λ Wλ = WΛ as, if v = limλ↑Λ vλ ∈ Wλ(Ω), then for every λ ∈ L we have that fλ (vλ) ≤ fλ (uλ), hence F (v) = lim λ↑Λ fλ (vλ) ≤ lim λ↑Λ fλ (uλ) = F (u). (cid:3) For applications, the following particular case of Theorem 4.1 is particularly relevant: Corollary 4.2. Let f (ξ, u, x) be cohercive in ξ on every finite dimensional subspace of V (Ω) and for every x ∈ Ω. Let F (u) := › f (∇u, u, x)dx. Then F ◦ has a min on VΛ. Proof. Just notice that F ∈ F , in the notations of Theorem 4.1. (cid:3) Theorem 4.1 provides a general existence results. However, such a general result poses two questions: the first is how wild such generalized solutions can be; the second is if this method produces new generalized solutions for problems that already have classical ones. The answer to these questions depends on the problem that is studied. However, regarding the second question we have the following result, which strengthens Theorem 4.1: Theorem 4.3. Let F : VΛ(Ω) → R∗, F = limλ↑Λ Fλ. For every λ ∈ L let Mλ :=(cid:26)u ∈ Vλ(Ω) Fλ(u) = min v∈Vλ Fλ(v)(cid:27) . 4We recall that, by a well known theorem of Schwartz, any tempered distribution can be represented as ∂mf where m is a multi-index and f is a continuous function. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 15 Assume that limλ↑Λ Mλ 6= ∅. Then MΛ :=(cid:26)u ∈ VΛ(Ω) F (u) = min v∈VΛ F (v)(cid:27) = lim λ↑Λ Mλ 6= ∅. Proof. MΛ ⊆ limλ↑Λ Mλ: Let v = limλ↑Λ vλ ∈ MΛ, and let u = limλ↑Λ uλ ∈ limλ↑Λ Mλ. As F (v) ≤ F (u), there is a qualified set Q such that for every λ ∈ Q Fλ (vλ) ≤ F (uλ). But then vλ ∈ Mλ for every λ ∈ Q, hence v = limλ↑Λ vλ ∈ limλ↑Λ Mλ. MΛ ⊇ limλ↑Λ Mλ: Let u = limλ↑Λ uλ ∈ limλ↑Λ Mλ. Let v = limλ↑Λ vλ ∈ VΛ(Ω). Let Q = {λ ∈ L uλ ∈ Mλ}. Then Q is qualified and, for every λ ∈ Q, Fλ (uλ) ≤ Fλ (vλ). Therefore F (u) ≤ F (v), and so u ∈ MΛ. (cid:3) The following easy consequences of Theorem 4.3 hold: Corollary 4.4. In the same notations of Theorem 4.3, let us now assume that there exists k ∈ N such that Mλ ≤ k for every λ ∈ L. Then MΛ ≤ k. Proof. This holds as the hypothesis on Mλ trivially entails that limλ↑Λ Mλ ≤ k. (cid:3) Corollary 4.5. In the same notations of Theorem 4.3, let us now assume that F = J ∗, where J : V (Ω) → R. Let M :=(cid:26)v ∈ V (Ω) v = min w∈V (Ω) J(w)(cid:27) . Assume that M 6= ∅. Then the following facts are equivalent: (1) u is a minimizer of F : VΛ(Ω) → R∗; (2) u ∈ M ∗ ∩ VΛ(Ω). In particular, if u ∈ M then u∗ minimizes F . Proof. (1) ⇒ (2) Let u ∈ M . Let Q(u) := {λ ∈ L u ∈ λ}. Then for every λ ∈ Q(u) v ∈ Mλ ⇔ J(v) = J(u) ⇒ v ∈ M , hence Mλ ⊆ M for every λ ∈ Q(u), which is qualified, and so limλ↑Λ Mλ ⊆ M ∗ ∩ VΛ, and we conclude by Theorem 4.3. (2) ⇒ (1) By definition, v∈[V (Ω)]∗ hence if u ∈ M ∗ ∩ Vλ(Ω) it trivially holds that u minimizes F . u ∈ M ∗ ⇔ F (u) = min F (v), (cid:3) Corollary 4.6. In the same hypotheses and notations of Corollary 4.3, let us assume that M = {u1, . . . , un} is finite. Then v minimizes F in VΛ(Ω) if and only if there exists u ∈ M such that u∗ = v. Proof. Just remember that S◦ = {s∗s ∈ S} for every finite set S, and that M σ = {u∗ u ∈ M } ⊆ V (Ω) ⊆ VΛ(Ω). (cid:3) In general, one might not have minima, but minimization sequences could still exist. In this case, we have the following result (in which for every ρ ∈ R∗ we set stR(ρ) = −∞ if and only if ρ is a negative infinite number). Notice that in the following result we are not assuming the continuity of J with respect to any topology on V (Ω), in general. Theorem 4.7. Let V (Ω) be a Banach space, let J : V (Ω) → R and let inf u∈V (Ω) J(u) = m ∈ R ∪ {−∞}. The following facts hold: 1. J ∗(v) ≥ m for every v ∈ VΛ(Ω). 2. There exists v ∈ VΛ(Ω) such that stR(J ∗(v)) = m. 3. If v ∈ VΛ(Ω) is a minimum of J ∗ : VΛ(Ω) → R∗ then J ∗(v) ≥ stR (J ∗(v)) = m. 4. Let {un}n∈N be a minimizing sequence that converges to u ∈ V (Ω) in some topology τ . Then there exists v ∈ VΛ(Ω) such that stτ (v) = u and J ∗(v) ≥ stR (J ∗(v)) = m. Moreover, if w◦ + ψ is the canonical splitting of v, then: • if τ is the topology of pointwise convergence, then w = v and w(x) = u(x) for every x ∈ Ω; • if τ is the topology of pointwise convergence a.e., then w = v and w(x) = u(x) a.e. in x ∈ Ω; 16 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA • if τ is the topology of weak convergence, then w(x) = u(x) for every x ∈ Ω and hψ, ϕ∗i∗ ∼ 0 for every ϕ in the dual of V (Ω); • if τ is the topology associated with a norm · and, moreover, {un}n converges pointwise to u, then w = u and ψ∗ ∼ 0. 5. If all minimizing sequences of J converge to u ∈ V (Ω) in some topology τ and v is a minimum of J ∗ : VΛ(Ω) → R∗ then stτ (v) = u and J ∗(v) ≥ stR(J ∗(v)) = m. Proof. (1) Let v = limλ↑Λ vλ. Since m = inf u∈V (Ω) J(u) we have that for every λ ∈ Λ J (vλ) ≥ m, hence J ∗ (v) ≥ m. (2) By (1) it sufficies to show that stR(J ∗(v)) = m. Let {un}n∈N be a minimizing sequence for J. For every λ ∈ L let vλ := uλ. Let v := limλ↑Λ vλ. We claim that v is the desired ultrafunction. To prove that stR(J ∗(v)) = limn→+∞ J (un) = m we just have to observe that, by our definition of the net {vλ}λ, it follows that5 lim n→+∞ J (un) = stR(cid:18)lim λ↑Λ J (vλ)(cid:19) , and we conclude as limλ↑Λ J (vλ) = J ∗(v) by definition. (3) Let v = limλ↑Λ vλ, and let w ∈ VΛ(Ω) be such that stR (J ∗(w)) = m. Then m ≤ J ∗(v) by (1), whilst stR (J ∗(v)) ≤ stR (J ∗(w)) = m. Hence st (J ∗(v)) = m, as desired. (4) Let v be given as in point (2). Let us show that stV (Ω)(v) = u: let A ∈ τ be an open neighborhood of u. As {un}n converges to u, there exists N > 0 such that for every m > N un ∈ A. Let µ ∈ L be such that µ > N . Then ∀λ ∈ Qµ := {λ ∈ L µ ⊆ λ} vλ ∈ A, and as Qµ is qualified, this entails that v ∈ A∗. Since this holds for every A neighborhood of u, we deduce that stτ (v) = u, as desired. Now let u = w◦ + ψ be the splitting of u. If τ is the pointwise convergence, stτ (v)(x) = u(x) for every x ∈ Ω, hence by Definition 3.9 we have that the singular set of u is empty and that w(x) = u(x) for every x ∈ Ω, as desired. A similar argument works in the case of the pointwise convergence a.e. If τ is the weak convergence topology, then stτ (v) = u means that hv, ϕ∗i∗ ∼ hu, ϕi for every ϕ in the dual of V (Ω). Now let S be the singular set of u. We claim that S = ∅. If not, let x ∈ S and let ϕ = δx. Then hv, ϕ∗i∗ = v(x) is infinite, whilst hu, δxi = u(x) is finite, which is absurd. Henceforth for every x ∈ Ω we have that ψ(x) = 0. But hu, ϕi ∼ hv, ϕ∗i∗ = hw◦ + ψ, ϕ∗i∗ = hw◦, ϕ∗i∗ + hψ, ϕ∗i∗ = hw, ϕi + hψ, ϕ∗i∗, hence stτ (ψ) = u − w. As ψ(x) = 0 for all x ∈ Ω, this means that u(x) = w(x) for every x ∈ Ω. Then hu, ϕi + hψ, ϕ∗i∗ = hw, ϕi + hψ, ϕ∗i∗ = hv, ϕi ∼ hu, ϕi, and so hψ, ϕ∗i∗ ∼ 0. Finally, if τ is the strong convergence with respect to a norm · and {un}n converges pointwise to u, then by what we proved above we have that v(x) ∼ u(x) ∀x ∈ Ω, hence u(x) ∼ w(x) for every x ∈ Ω, which means u = w as both u, w ∈ V (Ω). Then ψ = u − w◦ = u − v◦ + v◦ − w◦ ∼ 0. (5) Let v = limλ↑Λ vλ. By point (2), the only claim to prove is that stτ (v) = u. We distinguish two cases: Case 1: J ∗(v) ∼ r ∈ R. As we noticed in point (2), it must be r = m. By contrast, let us assume that stτ (v) 6= u. In this case, there exists an open neighborhood A of u such that the set is qualified. For every n ∈ N, let 5A proof of this simple claim is given in Lemma 28 in [11]. Q := {λ ∈ L vλ /∈ A} GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 17 Qn :=(cid:26)λ ∈ L J(vλ) − r < 1 n(cid:27) ∩ Q. Every Qn is qualified, hence nonempty. For every n ∈ N, let λn ∈ Qn. Finally, set un := vλn . By construction, limn∈N J (un) = m. This means that {un}n∈N is a minimizing sequence, hence it converges to u in the topology τ , and this is absurd as, for every n ∈ N, by construction un /∈ A. Henceforth stτ (v) = u. Case 2: J ∗(v) ∼ −∞. As we noticed in the proof of point (2), in this case m = −∞. Let us assume that stV (Ω)(v) 6= u. Then there exists an open neighborhood A of u such that the set is qualified. For every n ∈ N, let Q := {λ ∈ L vλ /∈ A} and let λn ∈ Qn. Finally, let un := vλn . Then J (un) < −n for every n ∈ N, hence {un}n∈N is a minimizing sequence, and so it must converge to u. However, by construction un /∈ A for every n ∈ N, which is absurd. (cid:3) Qn = {λ ∈ L J(vλ) < −n} ∩ Q Example 4.8. Let Ω = (0, 1), let V (Ω) = {u : Ω → R u is the restriction to Ω of a piecewise C1([0, 1]) function} and let J : V (Ω) → R be the functional It is easily seen that inf u∈V (Ω) J(u) = 0, and that the minimizing sequences of J converge pointwise and J(u) := Ω u2(x)dx +Ω(cid:16)(u′)2 − 1(cid:17)2 dx. strongly in the L2 norm to 0, but J(0) = 1. Let v ∈ VΛ(Ω) be the minimum of J ∗ : VΛ(Ω). From points (4) and (5) of Theorem 4.7 we deduce that 0 < J ∗(v) ∼ 0, that stV (Ω)(v) = 0 and that the canonical decomposition of v is v = 0◦ + ψ, with ψ = 0 for every x ∈ Ω and ´ ∗ Ω∗ ψ2dx ∼ 0. Moreover, as J ∗(ψ) = 0, we also have that ´ ∗ 5. Applications Ω∗(cid:16)(ψ′)2 − 1(cid:17)2 dx ∼ 0. 5.1. Sign-perturbation of potentials. The first problem that we would like to tackle by means of ultrafunctions regards the sign-perturbation of potentials. Let us start by recalling some results recently proved by L. Brasco and M. Squassina in [13] as a refinement and extension of some classical result by Brezis and Nirenberg [14]. Let Ω be a bounded domain of RN with6 N > 2. Consider the minimization problem (5.1) S(a) := where a ∈ LN/2(Ω) is given, 2∗ = 2N/(N − 2), inf u∈D1,2 0 (Ω)(cid:26)k∇uk2 0 (Ω) :=(cid:8)u ∈ L2∗ a u2 dx : kukL2∗ (Ω) = 1(cid:27) , L2(Ω) +Ω (Ω) ∇u ∈ L2(Ω), u = 0 on ∂Ω(cid:9). D1,2 By Lagrange multipliers rule, minimizers of the previous problem (provided they exist) are constant sign weak solutions of (5.2) with µ = S(a), namely Ω u = 0, (−∆u + a u = µ u2∗−2 u, a u ϕ dx = µ Ω ∇u · ∇ϕ dx +Ω in Ω, on ∂Ω, u2∗−2 u ϕ dx, 6In [13] the authors work more in general with a p ∈ (1, N ), and consider also a fractional version of Problem 5.2; however, in this paper we prefer to consider only the local case p = 2. 18 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA for every ϕ ∈ D1,2 0 (Ω). The main result in [13] is the following Theorem, where the standard notations a+ = max{a, 0}, a− = max{−a, 0}, BR(x0) = {x ∈ RN : x − x0 < R} are used: Theorem 5.1 (Brasco, Squassina). Let Ω ⊂ RN be an open bounded set. Then the following facts hold: 1. If a ≥ 0, then S(a) does not admit a solution. 2. Let N > 4. Assume that there exist σ > 0, R > 0 and x0 ∈ Ω such that a− ≥ σ, a. e. on BR(x0) ⊂ Ω. Then S(a) admits a solution. 3. Let 2 < N ≤ 4. For any x0 ∈ Ω, for any R > 0 s.t. BR(x0) ⊂ Ω there exists σ = σ(R, N ) > 0 such that if then S(a) admits a solution. a− ≥ σ, a. e. on BR(x0), In [1], V. Benci studied in the ultrafunctions setting the following similar (simpler) problem: minimize min u∈Mp J(u), J(u) = Ω ∇u2 dx Mp =(cid:26)u ∈ C2 0(Ω) : Ω up dx = 1(cid:27) . Here Ω is a bounded set in RN with smooth boundary, N ≥ 3 and p > 2. In the ultrafunctions setting introduced in [1] (and with the notations of [1]), the problem takes the following form: where and (5.3) where and min u∈fMp J(u), ∇u2 dx Ω J(u) = ∗ fMp =(cid:26)u ∈ V 2,0 B (Ω) ∗ Ω up dx = 1(cid:27) with V 2,0 show that 0 (Ω); up ∈ C2 0(Ω)(cid:3) . B (Ω) = B(cid:2)C2 For every p > 2, problem (5.3) has a ultrafunction solution up and, by setting emp = J(up), one can • (i) if 2 < p < 2∗, then emp = mp ∈ R+ and there is at least one standard minimizer up, namely • (ii) if p = 2∗ (and Ω 6= RN ), then em2∗ = m2∗ + ε where ε is a positive infinitesimal; • (iii) if p > 2∗, then emp = εp where εp is a positive infinitesimal. au2 dx : kuk[L2∗ (RN )]∗ = 1) , Our goal is to show that a similar result can be obtained for Problem 5.2. In the present ultrafunctions setting, Problem 5.2 takes the following form: find u∈VΛ(Ω)( ∗ (5.4) ∇u2 dx + ∗ 0 (Ω)iΛ where a ∈ ∗(cid:2)LN/2(Ω)(cid:3) is given, and VΛ(Ω) =hD1,2 eS(a) := inf following: (RN )∗ (RN )∗ . With the above notations, we can prove the GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 19 Theorem 5.2. Let Ω ⊂ RN be an open bounded set. Then the following facts hold: (1) For every a ∈(cid:2)LN/2(Ω)(cid:3)∗ (2) Let a ∈(cid:2)LN/2(Ω)(cid:3). If u ∈ C1 (Ω) ∩ C0(cid:0)Ω(cid:1) is a minimizer of Problem 5.1 then u∗ is a minimizer of eS(a∗). (3) If a = 0, then eS(0) = S + ε, where there exists u ∈ VΛ(Ω) that minimizes eS(a). S := inf u∈D0(RN )\{0} k∇uk2 L2 u2 L2∗ and ε =(0, a strictly positive infinitesimal, if Ω = RN , if Ω 6= RN ; moreover, if u is the minimizer in VΛ(Ω), then the functional part w of u is 0; (4) Let a ≥ 0 have an isolated minimum xm, and let u ∈ VΛ(Ω) be the minimum of Problem 5.4. If u = w◦ + ψ is the canonical splitting of u, then w = 0 and ψ concentrates in xm, in the sense that for every x /∈ mon(xm) ψ(x) ∼ 0. Moreover, hψ, ϕ∗i∗ ∼ 0 for every ϕ in the dual of V (Ω). Proof. (1) This follows from Theorem 4.3, as the functional k∇uk2 on every finite dimensional subspace of VΛ. L2(Ω) + ´Ω a u2 dx admits a minimum (2) This follows from Corollary 4.5. (3) In [13, Lemma 3.1] it was proved that, if we consider problem 5.1, we have that S(0) = S and S(0) is attained in D0(Ω) if and only if Ω = RN . Therefore if Ω = RN the results follows from point (2). If Ω 6= RN , the fact that eS(0) = S + ε follows from Theorem 4.7.(3). Moreover, all minimizing sequences {un}n converge weakly to 0 in H 1, therefore they converges strongly in L2(Ω) and so they converge pointwise a.e., hence by Theorem 4.7.(5) we deduce that, in the splitting u = w◦ + ψ, we have that w = 0, namely the ultrafunction solution coincides with its singular part. (4) We start by following the approach of [13]: we let U be a minimizer of inf u∈D1,2 0 (Ω) D1,2 [u]2 kuk2 L2∗ and, for every ε > 0, let Uε(r) := ε as follows ε(cid:1). Let δ > 0 be such that Bδ(xm) ⊆ Ω, and let uδ,ε be defined where Θ is a constant given in Lemma 2.4 of [13]. Moreover, if F (u) := k∇uk2 small enough we have that F (uδ1,ε) ≤ F (uδ2,ε). Then (uε,ε) is a minimizing net (for ε → 0), so we can use Theorem 4.7.(4): as (uε,ε) converges pointwise to 0 we obtain that w = 0, whilst the definition of the net ensure the concentration of ψ in xm. The last statement is again a direct consequence of Theorem 4.7.(4). (cid:3) L2(Ω) +´Ω a u2 dx, for δ1, δ2 Let us notice that the above Theorem shows a strong difference between the ultrafunctions and the classical case: the existence of solutions in VΛ(Ω) is ensured independently of the sign of a whilst, as discussed in [13, Section 4], the conditions on a for the existence of solutions in the approach of L. Brasco and M. Squassina are essentially optimal. Of course, ultrafunction solutions might be very wild in general; their particular structure can be described in some cases, depending on a. 5.2. The singular variational problem. 2−N 2 U(cid:0) r uδ,ε = Uε(r) Uε(δ) Uε(r)−Uε(δΘ) Uε(δ)−Uε(δΘ) 0 if r ≤ δ, if δ < r ≤ δΘ; if r > δΘ; 20 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA 5.2.1. Statement of the problem. Let W be a C1-function defined in R \ {0} such that W (t) = +∞ lim t→0 and lim t→±∞ We are interested in the singular problem (SP): W (t) t2 = 0. Naive formulation of Problem SP: find a continuous function which satisfies the equation: (5.5) with the following boundary condition: u : Ω → R, − ∆u + W ′(u) = 0 in Ω (5.6) where Ω is an open set such that ∂Ω 6= ∅ and g ∈ L1(∂Ω) is a function different from 0 for every x which change sign; e.g. g(x) = ±1. Clearly, this problem does not have any solution in C1. This problem could be riformulated as a kind of free boundary problem in the following way: u(x) = g(x) for x ∈ ∂Ω, Classical formulation of Problem SP: find two open sets Ω1 and Ω2 and two functions such that all the following conditions are fulfilled: ui : Ωi → R, i = 1, 2 (5.7) (5.8) Ω = Ω1 ∪ Ω2 ∪ Ξ where Ξ = Ω1 ∩ Ω2 ∩ Ω; − ∆ui + W ′(ui) = 0 in Ωi, i = 1, 2; ui(x) = g(x) for x ∈ ∂Ω ∩ ∂Ωi, i = 1, 2; lim x→Ξ ui(x) = 0; Ξ is locally a minimal surface. Condition (5.8) is natural, since formally equation (5.5) is the Euler-Lagrange equation relative to the energy (5.9) E(u) = 1 2 Ω(cid:16)∇u2 + W (u)(cid:17) dx and the density of this energy diverges as x → Ξ. In general this problem is quite involved since the set Ξ cannot be a smooth surface and hence it is difficult to be characterized. However this problem becomes relatively easy if formulated in the framework of ultrafuctions. Let us recall that we the Laplace operator of a ultrafuction u ∈ V ◦(Ω) is defined as the only ultrafunction ∆◦u ∈ V ◦(Ω) such that where ∀v ∈ V ◦ 0 (Ω), "Ω ∆◦uvdx = −"Ω Du · Dv dx Notice that, we can assert that ∆◦u(x) = D · D(x) only in x ≁ ∂Ω∗. V ◦ 0 (Ω) :=(cid:8)v ∈ V ◦(Ω) ∀x ∈ ∂Ω ∩ Γ, v(x) = 0(cid:9) . Ultrafunction formulation of Problem SP 7: find u ∈ V ◦(Ω) such that (5.10) 7If u is a ultrafunction and W, W ′, etc. W ∗(u), (W ′)∗ (u), etc. u(x) 6= 0, ∀x ∈(cid:0)Ω(cid:1)∗ are functions, ∩ Γ, for short, we shall write W (u), W ′(u), etc instead of GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 21 (5.11) (5.12) − ∆◦u + W ′(u) = 0, for x ∈ Ω∗ ∩ Γ u(x) = g◦(x), for x ∈ (∂Ω)∗ ∩ Γ. As we will see in the next section, the existence of this problem can be easily proven using variational methods. 5.2.2. The existence result. The easiest way to prove the existence of a ultrafunction solution of problem SP is achieved exploiting the variational structure of equation (5.11). Let us consider the extension (5.13) of the functional (5.9) to the space E◦(u) = "Ω(cid:18) 1 2 Du2 + W (u)(cid:19) dx V ◦ g (Ω) :=(cid:8)u ∈ V ◦(Ω) ∀x ∈ (∂Ω)∗ ∩ Γ, u(x) = g(x)(cid:9) . ∗ Remark 5.3. We remark that the integration is taken over Ω∗, but u is defined in Ω . This is important, in fact for some x ∈ Ω∗, x ∼ ∂Ω∗, the value of Du(x) depends on the value of u in some point y ∈ ∂Ω∗, y ∼ x. This is a remarkable difference between the usual derivative and the ultrafunction derivative. Lemma 5.4. Equation (5.11) is the Euler-Lagrange equation of the functional (5.13). Proof. We use the expression of › as given in Definition 3.6. As E◦(u) = "Ω(cid:18) 1 2 Du2 + W (u)(cid:19) dx let us compute separately the variations given by 1 2 Du2 and W (u). As Du(a)2 da 1 2 is a quadratic form, for v ∈ V ◦ 0 (Ω) we have that "Ω∗ 1 2 Du2 dx = Xa∈Γ∩Ω∗ Du2 dx(cid:19) [v] = "Ω∗ du(cid:19)∗ Xa∈Γ∩Ω∗ 0 (Ω) is 1 2 (cid:18) d du(cid:19)∗(cid:18)"Ω∗ du(cid:19)∗(cid:18)"Ω∗ (W (u)dx)(cid:19) [v] =(cid:18) d (cid:18) d dE◦(u) [v] = "Ω∗ The variation given by W (u) for v ∈ V ◦ Therefore the total variation of E◦ is DuDvdx = "Ω∗ (−∆◦u · v) dx. W (u(a))v(a)da = "Ω∗ W ′(u(x))v(x)dx. (−∆◦u + W ′(u))) v dx, which proves our thesis. (cid:3) The existence of a ultrafunction solution of problem SP follows from the following Theorem: Lemma 5.5. The functional (5.13) has a minimizer. Proof. The functional E◦(u) is coercive in the sense that for any c ∈ R∗ Ec :=(cid:8)u ∈ Vg(Ω) E◦(u) ≤ c(cid:9) is hypercompact (in the sense of NSA) since Vg(Ω) is a hypefinite dimensional affine manifold. Then, since E◦ is hypercontinuous (in the sense of NSA), the result follows. (cid:3) Regarding Ξ being a minimal surface, we can prove the following: 22 VIERI BENCI, LORENZO LUPERI BAGLINI, AND MARCO SQUASSINA Proposition 5.6. Let u be the ultrafunction minimizer of Problem 5.13 as given by Lemma 5.5. Then the sets Ω1 = { x ∈ Ω ∀y ∈ mon(x) ∩ Γ, u(y) > 0} , Ω2 = { x ∈ Ω ∀y ∈ mon(x) ∩ Γ, u(y) < 0} are open, hence is closed. Ξ :=(cid:8) x ∈ Ω ∃y1, y2 ∈ mon(x) ∩ Γ, u(y1) < 0, u(y2) > 0(cid:9) Proof. This follows from overspill8: let us prove it for Ω1. Let x ∈ Ω1. By definition of Ω1, for every ε ∼ 0 we have that u(y) < 0 for every y ∈ Bε(x) ∩ Γ. Hence by overspill there exists a real number r > 0 such that u(y) < 0 for every y ∈ B∗ r (x) ∩ Γ, we deduce that the open ball Br(x) ⊆ Ω1. (cid:3) r (x) ∩ Γ. As Bσ r (x) ⊂ B∗ Notice that Property (1) in Proposition 5.6 is a first step towards Property 5.8 in the classical formulation of Problem SP. It is our conjecture, in fact, that Ξ is a minimal surface, at least under some rather general hypothesis. We have not been able to prove this yet, however. Let us conclude with a remark. When studying problems like Problem 5.13 with ultrafunctions, one would like to be able to generalize certain properties of elliptic equations based on the maximum principle: for example, one would expect to have the following properties: (1) Let Ω be a bounded connected open set with smooth boundary and let g be a bounded function. Then if u = w◦ + ψ is the canonical splitting of u as given in Definition 3.9, we have that w ∈ L∞ e ψ(x) ∼ 0 for every x ∈ Ω; (2) let Ω1, Ω2 be the sets defined in Proposition 5.6. Then in Ω1 ∪ Ω2 we have −∆w + W ′(w) = 0, ∆ψ(x) ∼ 0; (3) if a = inf (g) e b = sup(g) and W ′(t) ≥ 0 for all t ∈ R \ (a, b), we have that However, in the spaces of ultrafunctions constructed in this paper the maximum principle does not hold directly: this is due to the fact that the kernel of the derivative is, in principle, larger than the space of constants. This problem could be avoided by modifying the space of ultrafunctions: as this leads to some technical difficulties, we prefer to postpone this study to a future paper. a ≤ u(x) ≤ b. References [1] V. Benci, Ultrafunctions and generalized solutions, Adv. Nonlinear Stud. 13, (2013), 461–486. 2, 3, 5, 18 [2] V. Benci, L. Berselli, C. Grisanti, The Caccioppoli Ultrafunctions, Adv. Nonlinear Anal., DOI: https://doi.org/10.1515/anona-2017-0225. 2, 10 [3] V. Benci, M. Di Nasso, Alpha-theory: an elementary axiomatic for nonstandard analysis, Expo. Math. 21, (2003), 355-386. 3 [4] V. Benci, L. Luperi Baglini, A model problem for ultrafunctions, in: Variational and Topological Methods: Theory, Applications, Numerical Simulations, and Open Problems, Electron. J. Diff. Eqns., Conference 21 (2014), 11–21. 2, 3 [5] V. Benci, L. Luperi Baglini, Basic Properties of ultrafunctions, in: Analysis and Topology in Nonlinear Dierential Equa- tions (D.G. Figuereido, J.M. do O, C. Tomei eds.), Progress in Nonlinear Differential Equations and their Applications, 85 (2014), 61-86. 2 [6] V. Benci, L. Luperi Baglini, Ultrafunctions and applications, DCDS-S 7, 4, (2014), 593–616. 2 [7] V. Benci, L. Luperi Baglini, A non archimedean algebra and the Schwartz impossibility theorem,, Monatsh. Math. (2014), 503–520. 2, 14 8Overspill is a well known and very useful property in nonstandard analysis. The idea behind the version that we use here is the following: if a certain property P (x) holds for every x ∼ 0 then there must be a real number r > 0 such that P (x) holds for every x < r. For a proper formulation of overspill we refer to [17]. GENERALIZED SOLUTIONS OF PDES AND APPLICATIONS 23 [8] V. Benci, L. Luperi Baglini, Generalized functions beyond distributions, AJOM 4, (2014). 2 [9] V. Benci, L. Luperi Baglini, A generalization of Gauss' divergence theorem, in: Recent Advances in Partial Dierential Equations and Applications, Proceedings of the International Conference on Recent Advances in PDEs and Applications (V. D. Radulescu, A. Sequeira, V. A. Solonnikov eds.), Contemporary Mathematics (2016), 69–84. 2 [10] V. Benci, L. Luperi Baglini, A topological approach to non-Archimedean mathematics, in: "Geometric Properties for Parabolic and Elliptic PDE"' (F. Gazzola, K. Ishige, C. Nitsch, P. Salani eds.), Springer Proceedings in Mathematics & Statistics, Vol. 176 (2016), 17–40, doi 10.1007/978-3-319-41538-3 2. 2 [11] V. Benci, L. Luperi Baglini, Generalized solutions in PDE's and the Burgers' equation, J. Differential Equations 263, Issue 10, 15 November 2017, 6916–6952. 2, 16 [12] L. Boccardo, G. Croce, Elliptic partial differential equations, De Gruyter, (2013). [13] L. Brasco, M. Squassina, Optimal solvability for a nonlinear problem at critical growth, J. Differential Equations 264 (2018), 2242–2269. 17, 18, 19 [14] H. Brezis, L. Nirenberg, Positive solutions of nonlinear elliptic equations involving critical Sobolev exponents, Comm. Pure Appl. Math. 36 (1983), 437–477. 17 [15] J.F. Colombeau, Elementary introduction to new generalized functions. North-Holland Mathematics Studies, 113. Notes on Pure Mathematics, 103. North-Holland Publishing Co., Amsterdam, 1985 1 [16] Grosser, M., Kunzinger, M., Oberguggenberger, M., Steinbauer, R., Geometric theory of generalized functions, Kluwer, Dordrecht (2001). 1 [17] H.J. Keisler, Foundations of Infinitesimal Calculus, Prindle, Weber & Schmidt, Boston, (1976). 2, 22 [18] A. Lecke, L. Luperi Baglini, P. Giordano, The classical theory of calculus of variations for generalized smooth functions, Adv. Nonlinear Anal., doi.org/10.1515/anona-2017-0150. 1, 2 [19] E. Nelson, Internal Set Theory: A new approach to nonstandard analysis, Bull. Amer. Math. Soc., 83 (1977), 1165–1198. [20] A. Robinson, Non-standard Analysis,Proceedings of the Royal Academy of Sciences, Amsterdam (Series A) 64, (1961), 432–440. [21] M. Sato, , Theory of hyperfunctions. II. J. Fac. Sci. Univ. Tokyo Sect. I 8 (1959) 139–193. 1 [22] M. Sato, Theory of hyperfunctions. II. J. Fac. Sci. Univ. Tokyo Sect. I 8 (1960) 387–437. 1 [23] M. Squassina, Exstence, multiplicity, perturbation and concentration results for a class of quasi linear elliptic problems, Electronic Journal of Differential Equations, Monograph 07, 2006, (213 pages). ISSN: 1072-6691. (M. Squassina) Dipartimento di Matematica e Fisica Universit`a Cattolica del Sacro Cuore Via dei Musei 41, I-25121 Brescia, Italy E-mail address: [email protected] (V. Benci) Dipartimento di Matematica Universit`a degli Studi di Pisa Via F. Buonarroti 1/c, 56127 Pisa, Italy and Centro Linceo interdisciplinare Beniamino Segre Palazzo Corsini - Via della Lungara 10, 00165 Roma, Italy E-mail address: [email protected] (L. Luperi Baglini) Faculty of Mathematics University of Vienna Oskar-Morgenstern-Platz 1, 1090 Wien, Austria E-mail address: [email protected]
1706.00821
2
1706
2017-06-17T00:10:35
Anisotropic Regularity Principle in sequence spaces and applications
[ "math.FA" ]
We refine a recent technique introduced by Pellegrino, Santos, Serrano and Teixeira and prove a quite general anisotropic regularity principle in sequence spaces. As applications we generalize previous results of several authors regarding Hardy--Littlewood inequalities for multilinear forms.
math.FA
math
ANISOTROPIC REGULARITY PRINCIPLE IN SEQUENCE SPACES AND APPLICATIONS NACIB ALBUQUERQUE AND LISIANE REZENDE Abstract. We refine a recent technique introduced by Pellegrino, Santos, Serrano and Teixeira and prove a quite general anisotropic regularity principle in sequence spaces. As applications we generalize previous results of several authors regarding Hardy -- Littlewood inequalities for multilinear forms. Contents Introduction Inclusion Theorems 1. 1.1. Summability of multilinear operators 1.2. 2. The new Inclusion Theorem 2.1. The proof of Theorem 3 3. A new Regularity Principle for sequence spaces 4. Applications: Hardy -- Littlewood's inequalities References 1 1 2 3 4 6 7 8 7 1 0 2 n u J 7 1 ] . A F h t a m [ 2 v 1 2 8 0 0 . 6 0 7 1 : v i X r a 1. Introduction Regularity techniques are crucial in many fields of pure and applied sciences. Recently, Pellegrino, Teixeira, Santos and Serrano addressed a regularity problem in sequence spaces with deep connections with the Hardy -- Littlewood inequalities. In this paper we follow this vein, exploring the ideas from the Regularity Principle proven by Pellegrino et al. [21] and providing a couple of applications. The paper is organized as follows. In Section 2, borrowing ideas from [21] we prove an anisotropic inclusion theorem for summing operators that will be useful along the paper. The techniques and arguments explored in Section 2 paves the way to the statement of a kind of anisotropic regularity principle for sequence spaces/series, in Section 3, completing results from [21]. In the final section the bulk of results are combined to prove new Hardy -- Littlewood inequalities for multilinear operators. 1.1. Summability of multilinear operators. The theory of multiple summing multilinear mappings was introduced in [18, 23]; this class is certainly one of the most useful and fruitful multilinear generalizations of the concept of absolutely summing linear operators, with important connections with the Bohnenblust -- Hille and Hardy-Littlewood inequalities and its applications in applied sciences. For recent results on absolutely summing operators and these classical inequalities we refer to [7, 11, 17] and the references therein. The Hardy -- Littlewood inequalities have its starting point in 1930 with Littlewood's 4/3 inequality [16]. In 1934 Hardy and Littlewood extended Littlewood's 4/3 inequality [15] to more general sequence spaces. Both results are for bilinear forms. In 1981 Praciano-Pereira [24] extended these results to the m-linear setting and recently various authors have revisited this subject. In 2016 Dimant and Sevilla-Peris [13] proved the following inequality: for all positive integers m and all m < p ≤ 2m we have ∞ Xj1,··· ,jm=1 T (ej1,··· , ejm)   p−m p ≤(cid:16)√2(cid:17)m−1 kTk p p−m  (1.1) p−m cannot be for all continuous m-linear forms T : ℓp × ··· × ℓp → K. It is also proved that the exponent improved. However, this optimality seems to be just apparent, as remarked in some previous works (see [4, 11]). Following these lines, the exponent can be potentially improved in the anisotropic viewpoint. In order to do that, the theory of summing operators shall play a fundamental role. Along the years, somewhat puzzling inclusion results for multilinear summing operators were obtained [8, 9, 23]. In this note we prove an inclusion result for multiple summing operators generalizing recent approaches of 2010 Mathematics Subject Classification. 46G25, 47H60. Key words and phrases. Inclusion Theorem, Multilinear operators, Regularity Principle. 1 p 2 ALBUQUERQUE AND REZENDE Pellegrino, Santos, Serrano and Teixeira [21] and Bayart [5]. It is interesting to note that, en passant, a sharper version of the Hardy -- Littlewood inequalities for m-linear forms, not encompassed by several recent attempts (for instance, [4]), is provided as application. In fact, we show that under the same hypothesis of (1.1) we have (1.2) with   ∞ Xj1=1   ∞ . . . Xjm=1  kT (ej1, . . . , ejm)ksm  sm−1 sm 1 s1 s1 s2  ≤(cid:16)√2(cid:17)m−1 kTk . . .  2mp s1 = , . . . , sm = p p − m , mp + p − 2m which is quite better than (1.1). Despite the huge advance recently obtained on this direction in [4], our results and techniques were not encompassed by its techniques. Throughout this paper X, Y shall stand for Banach spaces over the scalar field K of real or complex numbers. The topological dual of X and its unit closed ball are denoted by X ′ and BX ′ , respectively. For r, p ≥ 1, a linear operator T : X → Y is said (r; p)-summing if there exists a constant C > 0 such that kT (xj)kr  for any weakly p-summable vector sequence (xj )j∈N ∈ ℓw Xj=1   ∞ 1 r ≤ C k(xj)j∈Nkw,p p (X), where ℓw p (X) :=  ϕ∈BX′ (xj )j∈N ∈ X N : k(xj )j∈Nkw,p := sup  ∞ Xj=1 ϕ(xj )p  1 p . < ∞  For p := (p1, . . . , pm) ∈ [1,∞]m, using the definitions of mixed norm Lp spaces from [6], the mixed norm sequence space ℓp(X) := ℓp1 (ℓp2 (··· (ℓpm (X))··· )) gathers all multi-index vector valued matrices x := (xj)j∈Nm with finite p-norm; here j := (j1, . . . , jm) stands for a multi-index as usual. Notice that each norm k · kpk is taken over the index jk and that each index jk is related to the k · kpk norm. For instance, when p ∈ [1,∞)m a vector matrix x belongs to ℓp(X) if, and only if, kxkp := ∞ Xj1=1   ∞   Xj2=1   pm−2 pm−1 ···  ∞ Xjm−1=1 ∞  Xjm=1  kxjkpm  pm−1 pm   . . .  p2 p3  1 p1 p1 p2  < ∞. Over the last years, many different generalizations of the theory of absolutely and multiple summing operators were obtained. A natural anisotropic approach to multiple summing operators is the following: For r, p ∈ [1, +∞)m, a multilinear operator T : X1 × ··· × Xm −→ Y is said to be multiple (r, p)-summing if there exists a constant C > 0 such that for all sequences xk := (xk pk (Xk), k = 1, . . . , m, ∞ ∞ rm−1 (cid:13)(cid:13)(cid:13) (T (xj))j∈Nm(cid:13)(cid:13)(cid:13)r := Xj1=1  where T (xj) := T (cid:0)x1 jm(cid:1). The class of all multiple (r, p)-summing operators is a Banach space with the norm defined by the infimum of all previous constants C > 0. This norm is denoted by π(r,p)(·) and the space that gathers all such operators by Πm (r;p)(X1, . . . , Xm; Y ). When r1 = ··· = rm = r, s1 = ··· = sm = s we simply write (r; p), (r; s), respectively. . . . Xjm=1  . . .  r2  kxkkw,pk, j1 , . . . , xm   Yk=1 ≤ C rm m r1 1 r1 j )j∈N ∈ ℓw kT (xj)krm  1.2. Inclusion Theorems. Basic results from the theory of summing operators are inclusion theorems. For linear operators, it is folklore that p-summability implies q-summability whenever 1 ≤ p ≤ q. More generally, although basic, the following is quite useful (see [12]). Linear Inclusion Theorem. If s ≥ r, q ≥ p and 1 operator is absolutely (s; q)-summing s , then every absolutely (r; p)-summing linear r ≤ 1 p − 1 q − 1 Throughout the development of the theory, inclusion theorems reveals as challenging problems (see, for instance, [23]). In [21, Proposition 3.4], the authors proved the followimg multilinear inclusion result: ANISOTROPIC REGULARITY PRINCIPLE IN SEQUENCE SPACES AND APPLICATIONS 3 Theorem 1 (Pellegrino, Santos, Serrano and Teixeira). Let m be a positive integer, r, p, q ∈ [1, +∞) be such that q ≥ p and 1 r − m p m q + > 0. Then for any Banach spaces X1, . . . , Xm, with (r;p) (X1, . . . , Xm; Y ) ⊂ Πm Πm (s;q) (X1, . . . , Xm; Y ) , 1 s − m q = 1 r − m p , and the inclusion operator has norm 1. Independently, F. Bayart in [5, Theorem 1.2] obtained a more general version. For p ∈ [1, +∞]m and each k ∈ {1, . . . , m} , we define (cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≥k := 1 pk + ··· + 1 1 p(cid:12)(cid:12)(cid:12) instead of (cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≥1 Theorem 2 (Bayart). Let m be a positive integer, r, s ∈ [1, +∞), p, q ∈ [1, +∞)m are such that qk ≥ pk, k = 1, . . . , m and pm . When k = 1 we write (cid:12)(cid:12)(cid:12) r −(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12)(cid:12) q(cid:12)(cid:12)(cid:12)(cid:12) > 0. 1 1 1 Then (r;p) (X1, . . . , Xm; Y ) ⊂ Πm Πm (s;q) (X1, . . . , Xm; Y ) , for any Banach spaces X1, . . . , Xm, with In the next section we prove the following inclusion theorem; the techniques are inspired by [21] and contained in the proof of the forthcoming Regularity Principle, in Section 3: = 1 s −(cid:12)(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12)(cid:12) . 1 r −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12) Theorem 3. Let m be a positive integer, r ≥ 1, s, p, q ∈ [1, +∞)m are such that qk ≥ pk, for k = 1, . . . , m and Then > 0. 1 r −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12)(cid:12) for any Banach spaces X1, . . . , Xm, with (r;p) (X1, . . . , Xm; Y ) ⊂ Πm Πm (s;q) (X1, . . . , Xm; Y ) , for each k ∈ {1, . . . , m}, and the inclusion operator has norm 1. = 1 sk −(cid:12)(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12)(cid:12)≥k , 1 r −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k 2. The new Inclusion Theorem Despite of the general status of the result, only basic facts are used along its proof. The first one is the classical linear inclusion. We need other standard inclusion type result that we write for future reference. Inclusion on ℓp spaces. For q ≥ p > 0, k · kq ≤ k · kp. The last ingredient is a corollary of one of the many versions of Minkowski's inequality (see [14, Corollary 5.4.2]): Minkowski's inequality. For any 0 < p ≤ q < ∞ and for any scalar matrix (aij)i,j∈N, 1/q ∞  Xi=1    ∞ Xj=1 aijp  q/p  ≤  ∞ Xj=1 ∞ Xi=1 1/p aijq!p/q  . 4 ALBUQUERQUE AND REZENDE 2.1. The proof of Theorem 3. The argument is inspired on the Regularity Principle of [21, Theorem 2.1]. We will proceed by induction on m. The initial case bilinear is a straightforward application of classical inclusion of linear operators and ℓp spaces. The ideas used are revealed in the case m = 3, thus we it discuss in details. Let T ∈ Π3 (r;p) (X1, X2, X3; Y ). Then there exists a constant C > 0 such that ∞ Xj3=1   kT (xj)kr  ∞  Xj1,j2=1  j )j∈N ∈ ℓw 1 r 1 r ·r  (2.1) for all sequences xk = (xk v3 : X3 −→ ℓ(r,r)(Y ) by for all x3 ∈ X3. By (2.1) we obtain, for all x3 ∈ ℓw p3(X3), =(cid:13)(cid:13)(cid:13) (T (xj))j∈N3(cid:13)(cid:13)(cid:13)r ≤ C 3 Yk=1 kxkkw,pk , pk (Xk) with k = 1, 2 fixed. Defining pk (Xk), k = 1, 2, 3. Let xk ∈ ℓw v3(x3) =(cid:0)T (cid:0)x1 j2 , x3(cid:1)(cid:1)j1,j2∈N , j1 , x2 1 r ∞ r  Xj3=1(cid:13)(cid:13)v3(cid:0)x3 j3(cid:1)(cid:13)(cid:13)   , ≤ C3(cid:13)(cid:13)x3(cid:13)(cid:13)w,p3 k=1(cid:13)(cid:13)xk(cid:13)(cid:13)w,pk with C3 = CQ2 to v3 ∈ Π(s3;q3)(cid:0)X3; ℓ(r,r)(Y )(cid:1) with q3 ≥ p3, s3 ≥ r such that 1 q3 − 1 p3 − 1 r ≤ 1 s3 . , i.e., v3 ∈ Π(r;p3)(cid:0)X3; ℓ(r,r)(Y )(cid:1). The linear inclusion [12, Theorem 10.4] lead us Let us take 1 s3 = 1 + 1 q3 > 0. Applying norm inclusion on ℓp we obtain (2.2) Fixing x1 ∈ ℓw p1(X1), x3 ∈ ℓw q3(X3) and defining v2 : X2 −→ ℓ(s3,s3)(Y ) by we observe that (2.2) leads us to v2(x2) =(cid:0)T (cid:0)x1 j1 , x2, x3 j3(cid:1)(cid:1)j1,j3∈N , 1 s3 , ∞ s3  Xj2=1(cid:13)(cid:13)v2(cid:0)x2 j2(cid:1)(cid:13)(cid:13)   p2(X2) with C2 = C(cid:13)(cid:13)x3(cid:13)(cid:13)w,q3(cid:13)(cid:13)x1(cid:13)(cid:13)w,p1 , i.e., v2 ∈ Π(s3;p2)(cid:0)X2; ℓ(s3,s3)(Y )(cid:1). By the linear inclusion  s3 s3 Xj2=1    for all x2 ∈ ℓw theorem, v2 ∈ Π(s2;q2)(cid:0)X2; ℓ(s3,s3)(Y )(cid:1) with q2 ≥ p2, s2 ≥ s3 and 1 ≤ C2(cid:13)(cid:13)x2(cid:13)(cid:13)w,q2 ≤ C2(cid:13)(cid:13)x2(cid:13)(cid:13)w,p2 s3 ≤ 1 p2 − 1 kT (xj)ks3  = C(cid:13)(cid:13)x1(cid:13)(cid:13)w,p1 Yk=2(cid:13)(cid:13)xk(cid:13)(cid:13)w,qk q2 − 1 Xj1=1 Xj3=1     , we get 1 s2 ∞ ∞ ∞ s2 s3 s2 3 . Taking 1 s2 = 1 s3 − 1 p2 + 1 q2 = 1 r − 1 p2 − 1 p3 + 1 q2 + 1 q3 , p3 r − 1   Xj1=1 ∞ ∞  Xj2=1  ∞  Xj3=1  kT (xj)ks3  1 s3 s3 s3  s3 s3  1 s3 s3 s3  s3 s3  1 s3 ∞ ∞ ∞ ∞   Xj2=1 =  Xj3=1   ≤  Xj1,j2=1 Xj3=1   Yk=1(cid:13)(cid:13)xk(cid:13)(cid:13)w,pk ≤ C(cid:13)(cid:13)x3(cid:13)(cid:13)w,q3 Xj1=1 kT (xj)kr  kT (xj)ks3  r   ∞ s3 2 . ANISOTROPIC REGULARITY PRINCIPLE IN SEQUENCE SPACES AND APPLICATIONS 5 since s3 ≤ s2, by using norm inclusion on ℓp, we have s2  kT (xj)ks3   Xj1=1   Xj3=1  s3  Xj2=1   1 s2 ∞ ∞ ∞ s2 s2 (2.3) ∞ ∞ ∞ ∞ kT (xj)ks3  kT (xj)ks3  Xj3=1 Xj1=1 =   Xj2=1    ≤   Xj1=1 Xj3=1 Xj2=1    Yk=2(cid:13)(cid:13)xk(cid:13)(cid:13)w,qk ≤ C(cid:13)(cid:13)x1(cid:13)(cid:13)w,p1 ∞ ∞ 3 . s2 s3  s3  s3 1 s2 1 s2 s2 s2 s2  s3  Now let us fix xk ∈ ℓw qk (Xk) with k = 2, 3 and let us define, for all x1 ∈ X1, . p1 s1 s2 ∞ ∞ ∞ 1 s1 k=1 k=1 1 qk 1 pk = 1 = 1 1 s1 + 1 q1 j2 , x3 Xj1=1 > 0, we have By choosing 1 s1 j3(cid:1)(cid:1)j2,j3∈N . s2 − 1   r −P3   Xj3=1 Xj2=1   v1(x1) =(cid:0)T (cid:0)x1, x2 Thus v1 ∈ Π(s2;p1)(cid:0)X1; ℓ(s2,s3)(Y )(cid:1). By combining (2.3) and the linear inclusion theorem, we get that v1 ∈ Π(s1;q1)(cid:0)X1; ℓ(s2,s3)(Y )(cid:1) with q1 ≥ p1 and s1 ≥ s2 such that 1 q1 − 1 1 p1 − s2 ≤ +P3 s3 kT (xj)ks3   once that v1 ∈ Π(s1;q1)(cid:0)X1; ℓ(s2,s3)(Y )(cid:1). Therefore, T ∈ Π3 Now we shall conclude the proof by an induction argument. Let us suppose the result is true for m − 1 and let T ∈ Πm   . . .  Xj1=1 Xjm=1   pk (Xk). For a fixed x1 ∈ ℓw for all sequences xk ∈ ℓw r ·r . . .   v(x2, . . . , xm) :=(cid:0)T (x1 j1 , x2, . . . , xm)(cid:1)j1∈N , (r;p2,...,pm) (X2, . . . , Xm; ℓr(Y )). Consequently, by induction hypothesis, norm inclusion and the p1(X1), v : X2 × ··· × Xm → ℓr(Y ) given by kT (xj)kr  kT (xj)kr  Yk=1(cid:13)(cid:13)xk(cid:13)(cid:13)w,pk Yk=1(cid:13)(cid:13)xk(cid:13)(cid:13)w,qk   Xj2=1 (r;p) (X1, . . . , Xm; Y ), i.e., Xj1,...,jm=1 (s;q) (X1, X2, X3; Y ). r ·r  s2  =  ≤ C ≤ C 1 r ·r ∞ ∞ ∞ ∞ 1 1 r m 1 r 3 , , 1 belongs to Πm−1 Minkowski inequality, ∞ X j2 =1   . . .  ∞ X jm =1 (cid:13)(cid:13)T (cid:0)xj(cid:1)(cid:13)(cid:13) sm  sm−1 sm . . .     ∞ X j1 =1   (2.4) s2 s3   s2 s2      ∞ . . . ≤ j2 =1 X  ≤ C (cid:13)(cid:13)x1(cid:13)(cid:13)w,p1   ∞ X jm=1   ∞ X j1 =1 (cid:13)(cid:13)T (cid:0)xj(cid:1)(cid:13)(cid:13) r  sm r   sm−1 sm . . .   m Y k=2 (cid:13)(cid:13)(cid:13) xk(cid:13)(cid:13)(cid:13)w,qk . 1 s2 s2 s3   1 s2 m with r ≤ sm ≤ ··· ≤ s2 and Fixing xk ∈ ℓw 1 s2 1 r − = Xk=2 qk (Xk), k = 2, . . . , m and defining, for all x1 ∈ X1, j2 . . . , xm Xk=2 u(x1) =(cid:0)T (cid:0)x1, x2 + 1 pk m 1 qk . jm(cid:1)(cid:1)j2,...,jm∈N we have that u ∈ Π(s2;p1)(cid:0)X1; ℓ(s2,...,sm)(Y )(cid:1). Applying the classical linear inclusion on (2.4), with q1 ≥ p1 and s1 ≥ s2 such that 1 p1 − 1 s2 ≤ 1 q1 − 1 s1 , 6 ALBUQUERQUE AND REZENDE s1 we gain u ∈ Π(s1;q1)(cid:0)X1; ℓ(s2,...,sm)(Y )(cid:1). Taking 1 Π(s1;q1)(cid:0)X1; ℓ(s2,...,sm)(Y )(cid:1), we have kT (xj)ksm . . . Xjm=1    Xj1=1    ∞ ∞ sm−1 sm = 1 s2 − 1 p1 . . .  s1 s2  + 1 q1 1 s1 ≤ C 1 = 1 p(cid:12)(cid:12)(cid:12) r − (cid:12)(cid:12)(cid:12) + (cid:12)(cid:12)(cid:12) Yk=1(cid:13)(cid:13)xk(cid:13)(cid:13)w,qk m . Therefore, T ∈ Πm is preserved. This concludes the proof. (s;q) (X1, . . . , Xm; Y ). Also note that the inclusion operator has norm 1, since the constant C (cid:3) It is important to highlight the difference between Theorems 1, 2 and 3. Under the hypothesis of Theorem 3, by using the usual inclusion of ℓp spaces and Theorem 2 with 1 s1 (r;p) ⊂ Πm Πm (s1;q). := 1 r −(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12) one may get that > 0, since u ∈ 1 q(cid:12)(cid:12)(cid:12) But if r ≤ sm ≤ ··· ≤ s1, Nevertheless we may not conclude by Theorem 2 that (s;q) ⊂ Πm Πm (s1;q). (r;p) ⊂ Πm Πm (s;q). However this is provided by Theorem 3. For instance, let us illustrate with a numerical example: let r := 3, s := (5, 3), p := (3, 2) and q := (5, 2). From Theorem 2 we have (3;p) ⊂ Π2 Π2 (5;q), while Theorem 3 provides The same can also be done when m = 3: let r := 2, s := (6, 3, 2), p := (2, 2, 1) and q := (3, 3, 1). Then (3;p) ⊂ Π2 Π2 (5,3;q). where the first inclusion is assured by Theorem 3. (2;p) ⊂ Π3 Π3 (6,3,2;q) ⊂ Π3 (6;q), 3. A new Regularity Principle for sequence spaces The investigation of regularity-type results in this setting was initiated in [20] and expanded in [21]. In this short section we present a stronger version of these results. Let m ≥ 2 and Z1, V and w1, . . . , Wm be arbitrary non-empty sets and Z2, . . . , Zm be vector spaces. Let also be arbitrary maps, with k = 1, . . . , m, satisfying Rk : Zk × Wk → [0,∞) and S : Z1 × ··· × Zm × V → [0,∞) Rk(λz, w) = λRk(z, w) and S(z1, . . . , zj−1, λzj, zj+1, . . . , zm, ν) = λS(z1, . . . , zj−1, zj, zj+1, . . . , zm, ν) for all scalars λ ≥ 0 and j, k ∈ {2, . . . , m}. We shall work with each pk ≥ 1 and also assuming that 1 pk nk sup w∈Wk Xj=1  Rk(cid:0)zk j , w(cid:1)pk  < ∞, k = 1, . . . , m. Despite the abstract context, the proof is similar to the the proof of Theorem 3, and we omit the details. Theorem 4 (Anisotropic Regularity Principle). Let m be a positive integer, r ≥ 1, s, p, q ∈ [1, +∞)m be such that qk ≥ pk, for k = 1, . . . , m and Assume that there exists a constant C > 0 such that > 0. 1 q(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 1 r 1 1 n1 nm sup ··· j1 , . . . , zm Xjm=1 r −(cid:12)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) jm, ν(cid:1)r  j ∈ Zk and nk ∈ N with k = 1, . . . , m. Then   ν∈V  Xj1=1   ···  jm, ν(cid:1)sm  S(cid:0)z1 S(cid:0)z1 Xjm=1 j1, . . . , zm Xj1=1 sm−1 nm n1 sm for all z(k) sup ν∈V m Yk=1 sup w∈Wk  nk Xj=1 Rk(cid:0)zk j , w(cid:1)pk  1 pk , ···  1 s1 s1 s2  ≤ C · m Yk=1 nk sup w∈Wk Xj=1  1 qk , Rk(cid:0)zk j , w(cid:1)qk  ANISOTROPIC REGULARITY PRINCIPLE IN SEQUENCE SPACES AND APPLICATIONS 7 for all z(k) j ∈ Zk and nk ∈ N, k = 1, . . . , m, with = 1 1 sk −(cid:12)(cid:12)(cid:12)(cid:12) q(cid:12)(cid:12)(cid:12)(cid:12)≥k 1 r −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k , k ∈ {1, . . . , m}. 4. Applications: Hardy -- Littlewood's inequalities The Hardy -- Littlewood inequalities have been investigated in depth in the recent years (see, for instance, . There is a constant C K [1, 2, 3, 4, 10, 11, 13, 17, 19, 21, 22]). Here Xp stands for ℓp if p < ∞ and X∞ := c0. Theorem 5 (Albuquerque, Bayart, Pellegrino, Seoane [2], 2014). Let p, s ∈ [1, +∞]m such that 1/p ≤ 1 s ∈h(1 − 1/p)−1 , 2im  Xj1=1   ···  m,p,s ≥ 1 such that ···  s2  m,p,s kAk for every continuous m-linear form A : Xp1 × ··· × Xpm → K if, and only if, Xjm=1 ≤ C K sm−1 1 s1 ∞ ∞ s1 2 and Theorem 6 (Dimant, Sevilla-Peris [13], 2016). Let p ∈ [1, +∞]m such that 1/2 ≤ 1/p < 1. There exists a (optimal) constant DK (4.1) m,p ≥ 1 such that  Xj1,...,jm=1  ∞ 1−1/p ≤ DK m,pkAk 1 1 sm ≤ m + 1 A (ej1 , . . . , ejm )sm  p(cid:12)(cid:12)(cid:12)(cid:12) 2 −(cid:12)(cid:12)(cid:12)(cid:12) 1−1/p  A(ej1 , . . . , ejm) s(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) 1 . For every continuous m-linear form A : Xp1 × ··· × Xpm → K. Moreover, the exponent is optimal. The above exponent 1−1/p is optimal, but not in the anisotropic sense. In [4, Theorem 3.2] the authors improved Theorem 6, under some restriction over p. The following recent result is a kind of anisotropic extension of it: 1 Theorem 7 (Aron, N´unez, Pellegrino, Serrano [4], 2017). Let m ≥ 2, q1, . . . , qm > 0, and 1 < pm ≤ 2 < p1, . . . , pm−1, with 1/p < 1. There is a (optimal) constant Cp ≥ 1 such that s2  A (ej1 , . . . , ejm )sm   Xj1=1  ···  ≤ Cp kAk for every continuous m-linear form A : Xp1 × ··· × Xpm → K if, and only if, sm−1 1 s1 sm ∞ s1 , for k = 1, . . . , m. ∞ Xjm=1  ···  sk ≥"1 −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k#−1 Recently, W.V. Cavalcante has shown that (4.1) is a consequence of the inclusion result for multiple summing operators due to Pellegrino et al. combined with Theorem 5 (see [10]). The standard isometries between L(Xp, X) and ℓw p∗(X), for 1 < p ≤ ∞, allow us to read the previous Theorems 5, 6, 7 as coincidence results (see [12]). The key point is to begin with the coincidence below, obtained by revisiting Theorem 5 as a coincidence result with s1 = ··· = sm = 2 and p1 = ··· = pm = 2m, Πm (2;(2m)∗) (X1, . . . , Xm; K) = L (X1, . . . , Xm; K) , for all Banach spaces X1, . . . , Xm, and use an inclusion-type result. We shall combine these isometries with the inclusion result Theorem 3 to gain refined inclusions and coincidences. Theorem 8. Let m be a positive integer and s, p ∈ [1, +∞]m. If 1/p < 1 and p1, . . . , pm ≤ 2m, then Πm (2;(2m)∗) (X1, . . . , Xm; K) ⊂ Πm (s;p∗ 1,...,p∗ m) (X1, . . . , Xm; K) , for any Banach spaces X1, . . . , Xm, with sk =" 1 2 + m − k + 1 2m 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k#−1 −(cid:12)(cid:12)(cid:12)(cid:12) , for k = 1, . . . , m. 8 ALBUQUERQUE AND REZENDE Proof. Since each pk ≤ 2m and applying Inclusion Theorem 3, it is obtained the stated inclusion with for each k = 1, . . . , m. Corollary 1. If 1/p < 1 and p1, . . . , pm ≤ 2m, then for any Banach spaces X1, . . . , Xm and 1 1 2m(cid:19) + m −(cid:12)(cid:12)(cid:12)(cid:12) 2 − m ·(cid:18)1 − 2 − (m − k + 1) ·(cid:18)1 − 1 1 1 1 > 0, = 1 −(cid:12)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12)(cid:12) 2m(cid:19) + (m − k + 1) −(cid:12)(cid:12)(cid:12)(cid:12) 1 sk = Πm (s;p∗ 1,...,p∗ m) (X1, . . . , Xm; K) = L (X1, . . . , Xm; K) , = 1 sk −(cid:12)(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12)(cid:12)≥k 1 r −(cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k , k ∈ {1, . . . , m}. , 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k (cid:3) Bringing Theorem 8 to the context of sequence spaces, the announced anisotropic result will be achieved. s1 ∞ ∞ sm−1 (4.2) Xj1=1   The main result of this section reads as follows. Corollary 2. Let m be a positive integer and p ∈ [1, +∞)m such that 1/p < 1 and p1, . . . , pm ≤ 2m. Then, for all continuous m-linear forms A : ℓp1 × ··· × ℓpm → K kA (ej1 , . . . , ejm)ksm  p(cid:12)(cid:12)(cid:12)(cid:12)≥k#−1 2m(cid:19)(cid:21)−1 −(cid:12)(cid:12)(cid:12)(cid:12) 2 − (m − k + 1) ·(cid:18) 1 p − In order to clarify the new result, we illustrate the simpler case: when dealing with m < p1 = ··· = pm =  . . . Xjm=1   sk =" 1 p ≤ 2m, we get (4.2) with exponents . . .  s2  ≤ DK m,p,skAk sk =(cid:20) 1 m − k + 1 for k = 1, . . . , m. for k = 1, . . . , m, with 2m 1 s1 + 2 sm 1 1 , , that is, s1 = , . . . , sm = p p − m 2mp . mp + p − 2m It is obvious that the above exponents are better than the estimates of Theorem 6 that provides The following example is illustrative: s1 = ··· = sm = p p − m . Example 1. Suppose m = 3 and p = 4. By Theorem 6 we know that (4.2) holds with whereas by Corollary 2 we have sk ≥ 4 for k = 1, 2, 3, s1 ≥ 4, and s2 ≥ 3 References s3 ≥ 12/5. [1] N. Albuquerque, F. Bayart, D. Pellegrino and J.B. Seoane-Sep´ulveda, Sharp generalizations of the multilinear Bohnenblust -- Hille inequality, J. Funct. Anal. 266 (2014), 3726 -- 3740. [2] N. Albuquerque, F. Bayart, D. Pellegrino and J.B. Seoane-Sep´ulveda, Optimal Hardy -- Littlewood type inequalities for poly- nomials and multilinear operators, Israel J. Math. 211 (2016), no. 1, 197 -- 220. [3] G. Ara´ujo, D. Pellegrino and D. D. P. da S. e Silva, On the upper bounds for the constants of the Hardy -- Littlewood inequality, J. Funct. Anal. 267 (2014), no. 6, 1878 -- 1888. [4] R. Aron, D. Pellegrino, D. N´unez-Alarc´on and D. Serrano-Rodr´ıguez, Optimal exponents for Hardy -- Littlewood inequalities for m-linear operators, Linear Algebra and its App., in press. [5] F. Bayart, Multiple summing maps; coordinatewise summability, inclusion theorems and p-Sidon sets, arXiv:1704.04437v1. [6] A. Benedek, R. Panzone,T he space Lp, with mixed norm, Duke Math. J. 28 (1961) 301 -- 324. [7] O. Blasco, G. Botelho, D. Pellegrino and P. Rueda, Coincidence results for summing multilinear mappings. Proc. Edinb. Math. Soc. (2) [8] G. Botelho, H.-A. Braunss, H. Junek and D. Pellegrino, Inclusions and coincidences for multiple summing multilinear map- pings, Proc. Amer. Math. Soc. 137 (2009), no. 3, 991 -- 1000. ANISOTROPIC REGULARITY PRINCIPLE IN SEQUENCE SPACES AND APPLICATIONS 9 [9] G. Botelho, C. Michels, H. Junek and D. Pellegrino, Complex interpolation and summability properties of multilinear operators. Rev. Mat. Complut. 23 (2010), no. 1, 139 -- 161. [10] W.V. Cavalcante, Some applications of the Regularity Principle in sequence spaces, Positivity, in press. [11] W.V. Cavalcante, D.N´unez-Alarc´on, Remarks on an inequality of Hardy and Littlewood. Quaest. Math. 39 (2016), no. 8, 1101 -- 1113. [12] J. Diestel, H. Jarchow, and A. Tonge, Absolutely summing operators, Cambridge Studies in Advanced Mathematics 43, Cambridge University Press, Cambridge, 1995. [13] V. Dimant and P. Sevilla-Peris, Summation of coefficients of polynomials on ℓp spaces, Publ. Mat. 60 (2016), no. 2, 289 -- 310. [14] D. J.H. Garling, Inequalities: a journey into linear analysis, Cambridge University Press, Cambridge, 2007. [15] G. Hardy and J.E. Littlewood, Bilinear forms bounded in space [p, q], Quart. J. Math. 5 (1934), 241 -- 254. [16] J.E. Littlewood, On bounded bilinear forms in an infinite number of variables, Quart. J. Math. Oxford, 1 (1930), 164 -- 174. [17] M. Maia, T. Nogueira, D. Pellegrino, The Bohnenblust -- Hille inequality for polynomials whose monomials have a uniformly bounded number of variables, Integr. Equ. Oper. Theory 88 (2017), 143 -- 149. [18] M.C. Matos, Fully absolutely summing and Hilbert-Schmidt multilinear mappings, Collectanea Math. 54 (2003), 111 -- 136. [19] A. Nunes, A new estimate for the constants of an inequality due to Hardy and Littlewood. Linear Algebra Appl. 526 (2017), 27 -- 34. [20] D. Pellegrino, J. Santos. J.B. Seoane-Sepulveda, Some techniques on nonlinear analysis and applications. Adv. Math. 229 (2012), no. 2, 1235 -- 1265. [21] D. Pellegrino, J. Santos, D. Serrano-Rodr´ıguez, E. Teixeira, Regularity principle in sequence spaces and applications, arXiv:1608.03423v2 [22] D. Pellegrino, E. Teixeira, Towards sharp BohnenblustHille constants, Comm. Contemp. Math, in press. [23] D. P´erez-Garc´ıa, Operadores multilineales absolutamente sumantes, Ph.D. Thesis, Universidad Complutense de Madrid, 2004. [24] T. Praciano -- Pereira, On bounded multilinear forms on a class of ℓp spaces, J. Math. Anal. Appl. 81 (1981), 561 -- 568. (N. Albuquerque) Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900, Joao Pessoa -- PB, Brazil. E-mail address: [email protected] and [email protected] (L. Rezende) Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900, Joao Pessoa -- PB, Brazil. E-mail address: [email protected]
1805.08553
2
1805
2018-09-05T08:17:19
On the spectrum of multiplication operators
[ "math.FA" ]
We study relations between spectra of two operators that are connected to each other through some intertwining conditions. As application we obtain new results on the spectra of multiplication operators on $B(\cl H)$ relating it to the spectra of the restriction of the operators to the ideal $\mathcal C_2$ of Hilbert-Schmidt operators. We also solve one of the problems, posed in [B.Magajna, Proc. Amer. Math. Soc, 141 2013, 1349-1360] about the positivity of the spectrum of multiplication operators with positive operator coefficients when the coefficients on one side commute. Using the Wiener-Pitt phenomena we show that the spectrum of a multiplication operator with normal coefficients satisfying the Haagerup condition might be strictly larger than the spectrum of its restriction to $\mathcal C_2$.
math.FA
math
ON THE SPECTRUM OF MULTIPLICATION OPERATORS V. S. SHULMAN AND L. TUROWSKA To Yuri Stefanovich Samoilenko with love and gratitude for theorems and songs Abstract. We study relations between spectra of two operators that are connected to each other through some intertwining conditions. As application we obtain new results on the spectra of multiplication opera- tors on B(H) relating it to the spectra of the restriction of the operators to the ideal C2 of Hilbert-Schmidt operators. We also solve one of the problems, posed in [6], about the positivity of the spectrum of multipli- cation operators with positive operator coefficients when the coefficients on one side commute. Using the Wiener-Pitt phenomena we show that the spectrum of a multiplication operator with normal coefficients satis- fying the Haagerup condition might be strictly larger than the spectrum of its restriction to C2. 1. Introduction In operator theory often one has to deal with the "same" operators acting on different spaces, as for instance, the restrictions of an operator to (not necessarily closed) invariant subspaces supplied with different norms. The study of the structure, in particular spectral properties, of such operators may be difficult in one space and rather simple in another one. So it is natural to identify links between the operators that allow to connect their structural properties and in this way reduce difficult problems to simple ones. The present paper is devoted to establishing such links. The paper was influenced by the work of B. Magajna [6], where the author studied the spectrum of Luders operators, playing a significant role in quantum information theory. Luders operators or Luders operations are symmetric multiplication operators with positive operator coefficients on B(H) (the precise definition is given below); in [6] B. Magajna answered in negative the question, raised in [7], whether the spectra of such operators is always non-negative. The spectral theory of multiplication operators fits very well in the frame- work of the subject of the paper as such operators are naturally defined not only on the algebra B(H) of all bounded linear operators on a Hilbert space H but also on its symmetrically normed ideals; the most suitable such ideal is the ideal C2 of Hilbert-Schmidt operators, where many questions on multiplication operators have straightforward solutions. For example, the 1 2 V. S. SHULMAN AND L. TUROWSKA restriction of a Luders operator to C2 is a positive operator on the Hilbert space C2 and hence its spectrum is contained in R+ := [0, +∞). In this paper we obtain a new information about the spectra of multipli- cation operators. In particular, we solve one of the problems, posed in [6], about the spectra of multiplication operators with positive coefficients when the coefficients on one side build a commutative family. For this we study first general questions about relations between spectra in case when the op- erators (or representations of Banach algebras) are intertwined by map with trivial kernel or dense range. Through the paper B(X) denotes the algebra of all bounded linear oper- ators on a Banach space X. For T ∈ B(X) we write σ(T ) and ρ(T ) for the spectrum and the resolvent set of T respectively. 2. Intertwining Recall that a bounded linear operator S on a Banach space X is gen- eralized scalar if it admits a C ∞(R2)-calculation, that is there is a unital continuous homomorphism f 7→ f (S) from C ∞(R2) to B(X) satisfying con- dition z(S) = S, where z(x, y) = x + iy. In this case there are numbers k ∈ N such that kf (S)k ≤ Ckf kk,σ(S) where for each compact F ⊂ C, we denote kf kk,F = P0≤i,j≤k supz=x+iy∈F ∂i+jf ∂xi∂yj . The minimal such k is the order of S. Given arbitrary bounded operator T on X, the local resolvent set ρT (x) of T at the point x ∈ X is defined as the union of all open subsets U of C for which there is an analytic function f : U → X which satisfies (T − λ)f (λ) = x for all λ ∈ U. The local spectrum σT (x) of T at x ∈ X is defined as σT (x) := C \ ρT (x). The resolvent set ρ(T ) is always a subset of ρT (x) and hence σT (x) ⊂ σ(T ). For all subsets F ⊂ C we define the local spectral subspace XT (F ) of T to be XT (F ) = {x ∈ X : σT (x) ⊂ F }. The following result was proved in [5, chapter 3.5] for more general oper- ators but under the condition that the intertwining operator Φ is bounded. Theorem 2.1. Let X, Y be Banach spaces and let Φ : X → Y be a linear map with dense range. Suppose that (1) ΦT = SΦ, where T is a linear operator on X and S is a generalized scalar operator on Y . Then σ(S) ⊂ σ(T ). Proof. For a closed subset F ⊂ C, let YS(F ) denote the corresponding local spectral subspace. It is well known (see e.g. [2]) that YS(F ) is closed. Let k be the order of S. By [5, Theorem 1.5.4], for each p ≥ k + 3, (2) ∩λ /∈F (S − λ1)pY = YS(F ). ON THE SPECTRUM OF MULTIPLICATION OPERATORS 3 Now let F = σ(T ). Then, for each λ /∈ F , (T − λ1)pX = X. As (1) clearly implies Φ(T − λ1)p = (S − λ1)pΦ for all λ ∈ C, we obtain Φ(X) = Φ(∩λ /∈F (T − λ1)p(X)) ⊂ ∩λ /∈F Φ((T − λ1)p(X)) = ∩λ /∈F (S − λ1)p(Φ(X)) ⊂ ∩λ /∈F (S − λ1)p(Y ) = YF (S), whence Y = Φ(X) ⊂ YS(F ). σ(S) ⊂ F . It follows by [5, Proposition 1.2.20] that (cid:3) Corollary 2.2. Let H be a Hilbert space and X be a Banach space. (i) Let Φ : X → H be a linear map with dense range, T be a linear operator on X and N be a bounded normal operator on H such that ΦT = N Φ. Then σ(N ) ⊂ σ(T ). (ii) If Ψ : H → X is a linear injective bounded map, T is a bounded linear operator on X and N is a bounded normal operator on H such that ΨN = T Ψ then σ(N ) ⊂ σ(T ). Proof. The first statement follows directly from Theorem 2.1. To see the second one let Φ = Ψ∗ : X ∗ → H∗, then ΦX ∗ = H∗. Clearly ΦT ∗ = N ∗Φ and N ∗ is a normal operator on H∗ supplied with the natural scalar product. Since σ(N ∗) = σ(N ), σ(T ∗) = σ(T ) and σ(T ) is closed, the result follows from Theorem 2.1. (cid:3) Problem 2.3. Can one remove the conditions that Ψ and T are bounded? (i) The proof of Theorem 2.1 needs only one property of S: Remark 2.4. if, for some closed subset F ⊂ C, the subspace ∩λ /∈F (S − λ1)H is dense in H then F contains σ(S). It would be interesting to clear the class of such operators. (ii) The results of Theorem 2.1 and Corollary 2.2 was proved by Krein [4] for the case when N is selfadjoint. There are extensions of Krein's results to operators with spectra on a Jordan curve and satisfying restrictions on the norms of resolvent (see [11] and references therein). Let A be a unital regular Banach ∗-algebra of functions on a compact set K. Let π : A → B(H) be an (unital) injective ∗-representation of A on a Hilbert space H, and τ : A → B(X) be a (unital) representation on a Banach space X. We say that an operator Φ : X → H intertwines τ with π if Φτ (a) = π(a)Φ for all a ∈ A. Corollary 2.5. Suppose that π is injective. If there is a bounded linear operator Φ : X → H with dense image, which intertwines π with τ , then σ(τ (a)) = σ(a) for each a ∈ A. Proof. Clearly, τ (a) is invertible if a is invertible, whence σ(τ (a)) ⊂ σ(a). Similarly σ(π(a)) ⊂ σ(a). Conversely suppose that π(a) is invertible. Since the closure M of π(A) is a C*-algebra, π(a) is invertible in M . This means that χ(π(a)) 6= 0 for all χ ∈ Ω(M ), where Ω(M ) is the set of all characters of M . Let π∗ : Ω(M ) → K be the map adjoint to π: π∗(χ)(x) = χ(π(x)) for 4 V. S. SHULMAN AND L. TUROWSKA all x ∈ A. Then K0 := π∗(Ω(M )) is a compact set of K. If K0 6= K then by regularity of A there is 0 6= c ∈ A with c(t) = 0 for all t ∈ K0. Then χ(π(c)) = 0 for all χ ∈ Ω(M ), whence π(c) = 0. This contradicts to the injectivity of π. It follows that K0 = K and therefore a(t) 6= 0 for all t ∈ K. Thus a is invertible. We proved that σ(π(a)) = σ(a) for all a ∈ A. Clearly, the operator π(a) is normal whence σ(π(a)) ⊂ σ(τ (a)) by Corol- lary 2.2. Therefore σ(a) = σ(π(a)) ⊂ σ(τ (a)) ⊂ σ(a), giving the statement. (cid:3) Problem 2.6. For which non-commutative Banach ∗-algebras this is true? The answer is positive for C*-algebras. We will deduce it from a much more general result. Theorem 2.7. Any representation τ of a C*-algebra A on a Banach space X is a topological isomorphism of A/ ker(τ ) onto τ (A). Proof. Since A/ ker(τ ) is a C*-algebra we may assume that τ is injective. Let us firstly show that r(τ (a)) = r(a) for each positive element a ∈ A (where by r we denote the spectral radius). The inequality r(τ (a)) ≤ r(a) follows from the evident inclusion σ(τ (a)) ⊂ σ(a). Note that the opera- tor T = τ (a) admits calculus of continuous functions (f (T ) := τ (f (a)). If r(τ (a)) < r(a), choose a number t0 between r(τ (a)) and r(a) and a continu- ous function f on R with f (t) = 0 for t ∈ (−∞, t0), f (t) = 1 for t ≥ r(τ (a)). Then τ (f (a)) = f (τ (a)) = 0 while f (a) 6= 0. This contradicts to the injec- tivity of τ . Now for arbitrary a ∈ A we have kak2 = ka∗ak = r(a∗a) = r(τ (a∗a)) ≤ kτ (a∗a)k = kτ (a∗)τ (a)k ≤ kτ (a∗)kkτ (a)k ≤ kτ kkakkτ (a)k, whence kak ≤ kτ kkτ (a)k. We proved that the map τ (a) 7→ a is continuous. (cid:3) For the case when X is a Hilbert space, this result was obtained by Pitts [9], whose proof was based on the Dixmier-Day Theorem. Corollary 2.8. Let π be a ∗-representation of a C*-algebra on a Hilbert space, then for any Banach space representation τ with ker τ = ker π, one has σ(π(a)) = σ(τ (a)) for all a ∈ A. Proof. The equality immediately follows from Theorem 2.7 if σ(τ (a)) is the spectrum of τ (a) in τ (A). To see that the latter coincides with the spectrum of τ (a) in B(X) it suffices to show that if τ (a) is invertible in B(X), then a is invertible in A/ ker π. Replacing A by A/ ker π, suppose that a is not invertible in A. Then either a∗a or aa∗ is not invertible. In the first case a = (a∗a)1/2 is not invertible whence there is a sequence xn ∈ A with kxnk = 1 ON THE SPECTRUM OF MULTIPLICATION OPERATORS 5 and kaxnk → 0 whence kaxnk → 0. It follows that kτ (a)τ (xn)k → 0. Since infn kτ (xn)k > 0, the operator τ (a) is not invertible. Similarly one treats the case when aa∗ is not invertible. (cid:3) It remains to note that if representations π and τ are intertwined by a (non necessarily continuous) injective operator W with dense range, then their kernels coincide. If W is not injective then σ(τ (a)) = σ(π1(a)), where π1 is the restriction of π to (ker W )⊥. 3. Approximate inverse intertwinings Let us say that a Banach space X is supplied with a weak topology ω if in its adjoint X ∗ a ∗-weakly dense closed subspace M is chosen and ω = σ(X, M ). We denote by Bω(X) the space of all ω-continuous operators T ∈ B(X) (clearly T ∈ Bω(X) if and only if T ∗M ⊂ M ). Similarly Bω(X, Y ) is the subspace of B(X, Y ) consisting of operators continuous with respect to the weak topologies in X and Y . Definition 3.1. Let X, Y be Banach spaces with fixed weak topologies, and let T ∈ Bω(X), S ∈ Bω(Y ) be operators intertwined by some operator Ψ ∈ Bω(X, Y ) : ΨT = SΨ. A sequence of operators Fn ∈ Bω(Y, X) is called an approximate inverse intertwining (AII) for the triple (T, S, Ψ) with respect to these topologies if FnΨ → 1X , ΨFn → 1Y and T Fn − FnS → 0 in the point-weak topologies. Theorem 3.2. If an intertwining triple (T, S, Ψ) has an approximate in- verse intertwining then all eigenvalues of S belong to σBω(X)(T ). Proof. Let Sy = λy for some non-zero y ∈ Y and λ ∈ C. Then (T − λ)Fny = Fn(S − λ)y + (T Fn − FnS)y = (T Fn − FnS)y → 0. Assume λ /∈ σBω(X)(T ). Then (T − λ)−1 is ω-continuous, giving Fny → 0 and therefore ΨFny → 0. As ΨFny → y, we get y = 0. A contradiction. (cid:3) Note that if X is reflexive then Bω(X) = B(X) and σBω(X)(T ) = σ(T ) which simplifies the statement of Theorem 3.2. Problem 3.3. Is it true that σ(S) ⊂ σ(T ), under the assumptions of The- orem 3.2? For a Banach space X, let ℓ∞(X) be the space of all bounded sequences with entries in X and c0(X) be the subspace of all sequences convergent (in norm) to 0. Let eX = ℓ∞(X)/c0(X). Each T ∈ B(X) defines, by componentwise action, an operator on ℓ∞(X) which leaves c0(X) invariant, and hence induces a bounded operator T on the quotient space X. Definition 3.4. We say that an intertwining triple (T, S, Ψ) is strongly ap- proximate invertible if the corresponding triple (eT , eS, eΨ) has an approximate inverse intertwining. (3) (4) X Aj2 < ∞ and X j∈J j∈J Bj2 < ∞. ∆(X) = X AjXBj. 6 V. S. SHULMAN AND L. TUROWSKA Corollary 3.5. If an intertwining triple (T, S, Ψ) is strongly approximate invertible then σap(S) ⊂ σ(T ), where σap(S) is the approximate spectrum of S. If, in addition, Ψ has dense range then σ(S) ⊂ σ(T ). Proof. It is easy to see that if λ ∈ σap(S) then λ is an eigenvalue for eS. By Theorem 3.2, λ ∈ σ(eT ) and hence λ ∈ σ(T ). If Ψ has dense range then Ψ∗ ∈ B(Y ∗, X ∗) is injective. As Ψ∗S∗ = T ∗Ψ∗, any eigenvalue λ of S∗ is an eigenvalue of T ∗ and hence in σ(T ∗) = σ(T ). Since the residual spectrum σr(S) of S is a subset of the point spectrum of S∗ and σ(S) = σap(S) ∪ σr(S), we obtain σ(S) ⊂ σ(T ). (cid:3) Let A = {Aj }j∈J , B = {Bj}j∈J (J is finite or countable) be families of operators on a Hilbert spaces H such that Then one can define a multiplication operator ∆ : B(H) → B(H) by j∈J Clearly, ∆ is bounded and preserves all symmetrically normed ideals of B(H). Let ∆C2 be the restriction of ∆ to the ideal of all Hilbert-Schmidt operators C2 on H. We are interested in the relations between the spectra of the operators ∆ and ∆C2. Let Ψ2 : C2 → B(H) be the identity inclusion. Then (∆C2, ∆, Ψ2) is an intertwining triple. The following condition is close in spirit to Voiculescu's notion of quasidi- agonality with respect to a symmetrically normed ideal [14]. Definition 3.6. We say that a family A = {Aj }j∈J of operators is 2-semi- diagonal if there exists a sequence of projections Pn of finite rank such that Pn → 1 in the strong operator topology, and [Aj, Pn]2 sup C2 < ∞. X j∈J n Theorem 3.7. Suppose that (3) holds and the family A = {Aj}j∈J is 2- semidiagonal. Then all eigenvalues of ∆ are contained in σ(∆C2 ). Proof. It was established in [13] that under these assumptions there exists an approximate inverse intertwining for the triple (∆C2, ∆, Ψ2), with respect to the usual weak topology in C2 and ∗-weak topology σ(B(H), C1(H)). It remains to apply Theorem 3.2 taking into account that C2 is reflexive. (cid:3) The conditions that imply 2-semidiagonality of A were studied in [13]. We mention only two of them: 1) the matrices of all Aj, with respect to some basis in H, are supported by a finite number of diagonals; ON THE SPECTRUM OF MULTIPLICATION OPERATORS 7 2) A is a family of commuting normal operators such that A has finite muliplicity and ess-dim A ≤ 2, the latter is the essential Hausdorff dimension of the joint spectrum of A, defined in [13]. In particular, if all Aj are Lipschitz functions of a Hermitian operator then ess-dim A ≤ 2. 4. Luders operators The formally adjoint operator e∆ of ∆ is defined by the formula A∗ j XB∗ j . e∆(X) = X j∈J The restrictions ∆C2, e∆C2 of operators ∆ and e∆ to the ideal C2 are adjoint to each other as operators on the Hilbert space C2. We say that ∆ is formally selfadjoint if e∆ = ∆, and formally normal if e∆∆ = ∆e∆. Clearly, ∆ is formally selfadjoint) if and only if ∆C2 is normal formally normal (resp. (resp. selfadjoint). One has many ways to distinct "positive" operators among formally self- adjoint ones. Let us say that ∆ is C2-positive if ∆C2 ≥ 0 as an operator on the Hilbert space C2. Furthermore, ∆ is formally positive if ∆ = eΛΛ for some multiplication operator Λ. Clearly, any formally positive operator is C2-positive. Another important subclass of the class of C2-positive operators consists of operators with positive coefficients (Aj ≥ 0 and Bj ≥ 0 for all j ∈ J). If J is finite and Aj = Bj for all j ∈ J then such operators are called Luders operators. We are interested in relations between the spectra of the operators ∆ and ∆C2. In particular, under which conditions the spectra of a formally selfadjoint operators is real, and the spectra of a C2-positive operator is non-negative? The fact that this is not always true was established by B. Magajna [6], who constructed examples of Luders operators with non-real spectra. A similar construction can be used to show that formally positive oper- ators can have non-real eigenvalues. In fact, given λ ∈ C, by [6, Corollary 2.5] there exist positive operators Aj, Bj, j = 1, 2, 3, on a Hilbert space H, such that λI = P3 j=1 AjXBj we obtain ΛΛ(I) = λ2I. j=1 AjBj. By letting Λ(X) = P3 Note that if the spectrum of a formally selfadjoint operator ∆ is not positive then the same is true for its restriction ∆C1 to the ideal C1 (the trace class), because ∆C1 is the predual of ∆ with respect to the duality B(H) = C∗ 1. In its turn the restriction of ∆ to the ideal K(H) of all compact operators is the predual of ∆C1, so the spectrum of ∆K(H) also needs not be positive or even real. On the other hand, as C1 ⊂ C2, all eigenvalues of ∆C1 are positive whenever ∆ is C2 positive. Problem 4.1. Let ∆ be a Luders operator. a) Can ∆K(H) have non-positive eigenvalues? b) Is it true that σ(∆Cp) ⊂ R+, for each p ∈ (1, ∞)? 8 V. S. SHULMAN AND L. TUROWSKA In search of conditions that provide the positivity of the spectrum for a Luders operator ∆ : X 7→ Pn j=1 AjXBj, B. Magajna considered the case when the left coefficients A1, ..., An are commuting. He proved that this condition of "one-sided commutativity" is sufficient if n = 2, and asked if the same is true for all n. Now we will show that the answer to this question is affirmative even for operators of infinite length. Theorem 4.2. Let W = Ab⊗B, where A = C(K), (K is compact), B is a unital Banach algebra and ⊗ denotes the projective tensor product. Then for each w = Pi fi ⊗ bi ∈ W one has σ(w) = ∪t∈K σ(X fi(t)bi). i Proof. Let Prim(W ) be the set of all primitive ideals of W . For each I ∈ Prim(W ), let qI be the quotient map W → W/I. By theorem of Zemanek [15], σ(w) = ∪I∈Prim(W )σ(qI (w)). If πI is an irreducible representation of W with kernel I then πI(A ⊗ 1) is in the center of πI(W ). Since πI(A ⊗ 1) is isomorphic to A/J where J = {f ∈ A : f ⊗ 1 ∈ I}, A/J is either one-dimensional or there are non-zero operators T1, T2 ∈ πI (A ⊗ 1) with T1T2 = 0. Setting Y = ker T1 we see that Y is a non-trivial subspace invariant for πI(W ). So A/J is one-dimensional, J = {f ∈ A : f (t) = 0}, for some t ∈ K. It follows that πI (f ⊗ 1) = f (t)1 and πI (f ⊗ b) = f (t)τ (b), for all f ∈ C(K), b ∈ B , where τ is an irreducible representation of B. Therefore πI(w) = Pi fi(t)τ (bi) = τ (Pi fi(t)b) whence σ(qI (w)) = σ(πI (w)) = σ(τ (X fi(t)bi)) ⊂ σ(X fi(t)bi). It follows that i i σ(w) ⊂ ∪t∈K σ(X fi(t)bi). i The converse inclusion is evident. (cid:3) Corollary 4.3. Let ∆(X) = P∞ P∞ i=1 kAik2 < ∞, P∞ σ(∆) ⊂ R+. i=1 kBik2 < ∞. i=1 AiXBi, with Ai ≥ 0, Bi ≥ 0 such that If AiAj = AjAi for all i, j, then Proof. Let A be the unital C*-algebra generated by all Ai. Then there is an isometric isomorphism φ of A onto C(K), where K is a compact. B(H) that sends f ⊗ T to the operator X 7→ φ(f )XT . Then ∆ = π(w), It follows that σ(∆) ⊂ σ(w) = Let W = C(K)b⊗B(H) and π = φ ⊗ id be the representation of W on where w = P∞ ∪t∈Kσ(Pi fi(t)Bi). Since fi(t) ≥ 0 for all i, the operators fi(t)Bi are posi- tive for all i and σ(P∞ i=1 fi(t)Bi) ⊂ R+. Hence σ(∆) ⊂ R+. i=1 fi ⊗ Bi, fi = φ−1(Ai). (cid:3) ON THE SPECTRUM OF MULTIPLICATION OPERATORS 9 The most well studied class of formally normal operators are operators with commutative normal coefficients. If J is finite their spectra are de- scribed in even more general situation [3]. The answer for infinite J is the same. Proposition 4.4. Let A and B be commutative families of normal operators satisfying (3). Then σ(∆) = {λ · µ : λ ∈ σ(A), µ ∈ σ(B)} = σ(∆C2). Proof. The subsets X = σ(A), Y = σ(B) of ℓ2(J) are compact in the weak topology of ℓ2(J), the functions τj(λ) := λj, ηj(µ) := µj, j ∈ J, are contin- uous so that the function F (λ, µ) := λ · µ = X τj(λ)ηj (µ) j belongs to the Varopoulos algebra V (X, Y ) = C(X) ⊗C(Y ). B(C2(H)) by letting for Φ(x, y) = P∞ i=1 fi(x)gi(y) Define now representations τ : V (X, Y ) → B(B(H)) and π : V (X, Y ) → τ (Φ)(T ) = ∞X i=1 fi(A)T gi(B), T ∈ B(H), and π(Φ) = τ (Φ)C2(H). Let Ψ2 : C2(H) → B(H) be the inclusion map. Then τ (Φ)Ψ2 = Ψ2π(Φ). By Corollary 2.5, σ(∆) = σ(F ) = σ(∆C2 ). (cid:3) In the next section we will show that the statement of Proposition 4.4 is false for more general class of multiplication operators with commuting normal coefficients. 5. Mutiplication operators with the Haagerup condition It is known that a multiplication operator (4) is well defined and bounded on B(H) if its coefficients satisfy a more general condition: kX j < ∞, kX AjA∗ B∗ j Bj < ∞, (5) j∈J j∈J where the convergence of the series and the expression for the multiplication operators is in the weak* topology. In this generality it is possible that ∆ does not preserve the Schatten-von Neumann ideals Cp, but if Aj and Bj are normal then this is true. Indeed in this case the operator e∆ is also bounded on B(H) whence by duality ∆ preserve C1 and the invariance of all Cp can be easily seen by using a complex interpolation argument. Let now the coefficient families A and B be commutative and consist of normal operators. Realizing them as families of multiplication operators on H1 = L2(Y, ν) and H2 = L2(X, µ) respectively, i.e. Ajg(y) = gj(y)g(y), Bjf (x) = fj(x)f (x), 10 V. S. SHULMAN AND L. TUROWSKA we obtain (6) ess supx∈X X fj(x)2 < ∞ ess supy∈Y X gj(y)2 < ∞. Hence F = Pj∈J fj ⊗gj is an element of the weak∗ Haagerup tensor product j∈J j∈J V ∞(X, Y ) := L∞(X, µ) ⊗w∗h L∞(Y, ν) (see [1] for the definition of this tensor product). It is known that elements i=1 ai ⊗ bi ∈ V ∞(X, Y ) can be identified with (marginally equivalence ϕ = P∞ classes of) functions F : X × Y → C, ϕ(x, y) = ∞X i=1 ai(x)bi(y). Recall that a subset E ⊂ X × Y is called marginally null (with respect to µ × ν) if E ⊂ (X1 × Y ) ∪ (X × Y1) and µ(X1) = ν(Y1) = 0. Set Γ(X, Y ) = L2(X, µ) ⊗L2(Y, ν). We can also identify Γ(X, Y ) with the space of all (marginally equivalence classes of) functions h : X × Y → C which admit representation h(x, y) = ∞X ui(x)vi(y), i=1 where ui ∈ L2(X, µ), vi ∈ L2(Y, ν), such that P∞ 2 < ∞, and P∞ i=1 vi2 2 < ∞. We write khkΓ for the projective norm of h ∈ Γ(X, Y ). It is known that B(L2(X, µ), L2(Y, ν)) is dual to Γ(X, Y ) and the duality is given by i=1 ui2 hX, f ⊗ gi = hXf, ¯gi, for X ∈ B(H1, H2), f ∈ L2(X, µ), g ∈ L2(Y, ν). One has V ∞(X, Y ) = {ϕ ∈ L∞(X × Y ) : ϕh ∈µ×ν Γ(X, Y ) ∀h ∈ Γ(X, Y )}, here ϕh ∈µ×ν Γ(X, Y ) means that ϕh differs from a function in Γ(X, Y ) on a µ × ν-null set. Let A(R) = FL1(R) be the Fourier algebra of R, where F is the Fourier transform. Then the map P : Γ(R, R) → A(R) (with the Lebesque measure on R) given by (7) where f (t) = f (−t), is a contractive surjection. P (f ⊗ g)(t) = g ∗ f (t), If B(R) is the Fourier-Stiltjes algebra of the group R then, for each h ∈ B(R), the function h(x − y) belongs to V ∞(R, R) (with respect to the Lebesgue measure). Recall that B(R) is isomorphic to the measure algebra M (R) via the Fourier-Stiltjes transform µ(t) = RR e−itxdµ(x), µ ∈ M (R). The classical fact of harmonic analysis, referred to as the Wiener-Pitt phenomen, is the existence of a function g ∈ B(R) such that g(x) ≥ 1, x ∈ R, and 1/g 6∈ B(R) (see e.g. [5, chapter 4]). ON THE SPECTRUM OF MULTIPLICATION OPERATORS 11 Let F (x, y) = g(y − x). Then by the above F (x, y) = ∞X j=1 fj(x)gj(y) with fj, gj ∈ L∞(R) satisfying (6). Let Aj, Bj be the multiplication opera- tors by gj and fj respectively and ∆(X) = P∞ Proposition 5.1. σ(∆) 6= {g(x) : x ∈ R} = σ(∆C2). j=1 AjXBj. Proof. By the choice of the function g we have that 0 6∈ {g(x) : x ∈ R}. To prove the claim it is enough to show that ∆ is not invertible. Assume to the contrary that ∆ has an inverse ∆′. Then, given Ψ ∈ Γ(R, R), we have h∆∆′(X), Ψi = h∆′(X), F Ψi = hX, Ψi, where h·, ·i is the pairing between B(L2(R)) and Γ(R, R). As null F = ∅, we have by [12, Corollary 4.3] that the space R(SF ) := {F Ψ : Ψ ∈ Γ(R, R)} is dense in Γ(R, R). Let SF : Γ(R, R) → Γ(R, R), Ψ 7→ F Ψ and SF −1 be the linear operator defined on R(SF ) by SF −1(F Ψ) = Ψ. From the above equality we have h∆′(X), F Ψi = hX, SF −1(F Ψ)i, SF −1 is bounded and ∆′ = (SF −1)∗. Write SF −1 also for the extension of SF −1 to Γ(R, R). Then SF −1 is the multiplication by 1/F . Let P : Γ(R, R) → A(R) be the surjective contraction given by (7), then P ((1/F )Ψ) = (1/g)P (Ψ) for any Ψ ∈ Γ(R, R). Therefore, 1/g is a multiplier of the Fourier algebra A(R) and hence 1/g ∈ B(R) by [10, Theorem 3.8.1]. A contradiction. (cid:3) Observe that we always have σ(∆C2) ⊂ σ(∆) due to Corollary 2.2 (ii). Similar arguments can be applied to prove the following Proposition 5.2. The spectrum of a C2-positive operator ∆ with coefficient families, satisfying (5) and consisting of commuting normal operators, is not necessarily contained in R+. Proof. We note first that there exists g ∈ B(R) such that g(x) ≥ 1, x ∈ R, but 1/g /∈ B(R). Indeed, assuming contrary to the statement that any such positive g ∈ B(R) is invertible we obtain that whenever f = f1 + if2 ∈ B(R) such that f ≥ 1 and f1 and f2 are real valued, f1, f2 ∈ B(R), f 2 = 1 + f 2 f 2 Let now h = g − 1 ∈ B(R) and set G(x, y) = g(x − y), H(x, y) = h(x − y), (x, y) ∈ R. We have σ(∆HC2 ) = {h(x) : x ∈ R} ⊂ R+. However ∆H + 1 = ∆G is not invertible by the arguments from the proof of the previous proposition. Hence −1 ∈ σ(∆H ). (cid:3) 2 ∈ B(R) and 1/f = (f1 − if2)/f 2 ∈ B(R), a contradiction. 12 V. S. SHULMAN AND L. TUROWSKA References [1] D. P. Blecher and R. R. Smith, The dual of the Haagrup tensor product, J. London Math. Soc. (2) 4 (1992), 126-144. [2] Colojoara and Foias, Theory of generalized spectral operators. Mathematics and its Applications, Vol. 9. Gordon and Breach, Science Publishers, New York-London- Paris, 1968. xvi+232 pp. [3] R. Curto and L. Fialkow, The spectral picture of [LA, RB], J. Funct. Anal. 71(1987), 371-392. [4] M.G. Krein, On linear bounded operators in functional spaces with two norms, Akad. Nauk Ukrain. RSR. Zbirnik Prac Inst. Mat. 9 (1947), 104-129. [5] K.B. Laursen and M.M. Neumann, An introduction to local spectral theory, Lon- don Mathematical Society Monographs. NewSeries, 20. The Clarendon Press, Oxford University Press, New York, 2000. [6] B. Magajna, Sums of products of positive operators and spectra of Luders operators, Proc. Amer. Math. Soc. 141 (2013), 1349-1360. [7] G. Nagy, On spectra of Luders operations, J. Math. Phys. 49 (2008), no.2, 022110. [8] V. V. Peller, Hankel operators in the theory of perturbations of unitary and selfad- joint operators, Funct. Anal. Appl. 19 (1985), no. 2, 111-123. [9] D.B. Pitts, Norming algebras and automatic complete boundedness of isomorphisms of operator algebras, Proc. Amer. Math. Soc. 136 (2008), 1757-1768. [10] W. Rudin, Fourier analysis on groups, Wiley-Interscience, 1990. [11] G.I. Russu, On the spectrum of operator in a pair of Banach spaces Mathematika, Izvestia VUZ'ov 318(1988), 11, 42 -- 49, in Russian. [12] V.S. Shulman and L. Turowska, Operator synthesis I. Synthetic sets, bilattices and tensor algebras, J. Funct. Anal. 209 (2004), 293-331. [13] V.S. Shulman and L. Turowska, Operator synthesis II. Individual synthesis and linear operator equations J. Reine Angew. Math. 590 (2006), 143 -- 187. [14] D. Voiculescu, Remarks on Hilbert-Schmidt perturbation of almost-normal opera- tors, Topics in modern operator theory (Timioara/Herculane 1980), Oper. Th. Adv. Appl. 2, Birkhuser, Basel-Boston, Mass. (1981), 311-318. [15] J. Zemanek, On spectral characterization of two-sided ideals in Banach algebras, Studia Math. 67 (1980), 1 -- 12. Department of Mathematics, Vologda State University, Vologda, Russia E-mail address: [email protected] Department of Mathematical Sciences, Chalmers University of Technol- ogy and the University of Gothenburg, Gothenburg SE-412 96, Sweden E-mail address: [email protected]
1806.08945
1
1806
2018-06-23T11:10:46
A note on homogeneous Sobolev spaces of fractional order
[ "math.FA", "math.AP" ]
We consider a homogeneous fractional Sobolev space obtained by completion of the space of smooth test functions, with respect to a Sobolev--Slobodecki\u{\i} norm. We compare it to the fractional Sobolev space obtained by the $K-$method in real interpolation theory. We show that the two spaces do not always coincide and give some sufficient conditions on the open sets for this to happen. We also highlight some unnatural behaviors of the interpolation space. The treatment is as self-contained as possible.
math.FA
math
A NOTE ON HOMOGENEOUS SOBOLEV SPACES OF FRACTIONAL ORDER LORENZO BRASCO AND ARIEL SALORT Abstract. We consider a homogeneous fractional Sobolev space obtained by completion of the space of smooth test functions, with respect to a Sobolev–Slobodeckiı norm. We compare it to the fractional Sobolev space obtained by the K−method in real interpolation theory. We show that the two spaces do not always coincide and give some sufficient conditions on the open sets for this to happen. We also highlight some unnatural behaviors of the interpolation space. The treatment is as self-contained as possible. Contents 1. Introduction 1.1. Motivations 1.2. Aims 1.3. Results 1.4. Plan of the paper 2. Preliminaries 2.1. Basic notation 2.2. Sobolev spaces 2.3. A homogeneous Sobolev–Slobodeckiı space 2.4. Another space of functions vanishing at the 1 1 3 4 5 5 5 6 6 boundary Interpolation VS. Sobolev-Slobodeckiı 7 8 3. An interpolation space 11 4. 11 4.1. General sets 14 4.2. Convex sets 20 4.3. Lipschitz sets and beyond 23 5. Capacities 6. Double-sided estimates for Poincar´e constants 26 Appendix A. An example 27 Appendix B. One-dimensional Hardy inequality 30 31 Appendix C. A geometric lemma References 32 1. Introduction 1.1. Motivations. In the recent years there has been a great surge of interest towards Sobolev spaces of fractional order. This is a very classical topic, essentially initiated by the Russian school in the 50s of the last century, with the main contributions given by Besov, Lizorkin, Nikol'skiı, Slobodeckiı and their collaborators. Nowadays, we have a lot of monographies at our disposal on 2010 Mathematics Subject Classification. 46E35, 46B70. Key words and phrases. Nonlocal operators, fractional Sobolev spaces, real interpolation, Poincar´e inequality. 1 2 BRASCO AND SALORT the subject. We just mention the books by Adams [1, 2], by Nikol'skiı [25] and those by Triebel [30, 31, 32]. We also refer the reader to [31, Chapter 1] for an historical introduction to the subject. The reason for this revival lies in the fact that fractional Sobolev spaces seem to play a funda- mental role in the study and description of a vast amount of phenomena, involving nonlocal effects. Phenomena of this type have a wide range of applications, we refer to [10] for an overview. There are many ways to introduce fractional derivatives and, consequently, Sobolev spaces of fractional order. Without any attempt of completeness, let us mention the two approaches which are of interest for our purposes: • a concrete approach, based on the introduction of explicit norms, which are modeled on the case of Holder spaces. For example, by using the heuristic δs hu(x) := u(x + h) − u(x) hs ∼ "derivative of order s ", for x, h ∈ RN , a possible choice of norm is (cid:18)(cid:90) and more generally(cid:18)(cid:90) (cid:107)δs hu(cid:107)q Lp (cid:19) 1 p , dh hN Lp , for 1 ≤ q ≤ ∞. hu(cid:107)p (cid:107)δs (cid:19) 1 q dh hN Observe that the integral contains the singular kernel h−N , thus functions for which the norm above is finite must be better than just merely s−Holder regular, in an averaged sense; • an abstract approach, based on the so-called interpolation methods. The foundations of these methods were established at the beginning of the 60s of the last century, by Calder´on, Gagliardo, Krejn, Lions and Petree, among others. A comprehensive treatment of this approach can be found for instance in the books [4, 3, 29] and references therein In a nutshell, the idea is to define a scale of "intermediate spaces" between Lp and the standard Sobolev space W 1,p, by means of a general abstract construction. The main advantage of this second approach is that many of the properties of the spaces constructed in this way can be extrapolated in a direct way from those of the two "endpoint" spaces Lp and W 1,p. As mentioned above, actually other approaches are possible: a possibility is to use the Fourier transform. Another particularly elegant approach consists in taking the convolution with a suitable kernel (for example, heat or Poisson kernels are typical choices) and looking at the rate of blow-up of selected Lp norms with respect to the convolution parameter. However, we will not consider these constructions in the present paper, we refer the reader to [31] for a wide list of definitions of this type. In despite of the explosion of literature on Calculus of Variations settled in fractional Sobolev spaces of the last years, the abstract approach based on interpolation seems to have been completely neglected or, at least, overlooked. For example, the well-known survey paper [14], which eventually became a standard reference on the field, does not even mention interpolation techniques. HOMOGENEOUS SOBOLEV SPACES 3 1.2. Aims. The main scope of this paper is to revitalize some interest towards interpolation theory in the context of fractional Sobolev spaces. In doing this, we will resist the temptation of any unnecessary generalization. Rather, we will focus on a particular, yet meaningful, question which can be resumed as follows: Given a concrete fractional Sobolev space of functions vanishing "at the boundary" of a set, does it coincide with an interpolation space? We can already anticipate the conclusions of the paper and say that this is not always true. Let Our concerns involve the so-called homogeneous fractional Sobolev-Slobodeckiı spaces Ds,p us now try to enter more in the details of the present paper. 0 (Ω). Given an open set Ω ⊂ RN , an exponent 1 ≤ p < ∞ and a parameter 0 < s < 1, it is defined as the completion of C∞ 0 (Ω) with respect to the norm u (cid:55)→ [u]W s,p(RN ) := RN×RN (cid:18)(cid:90)(cid:90) (cid:19) 1 p . u(x) − u(y)p x − yN +s p dx dy Such a space is the natural fractional counterpart of the homogeneous Sobolev space D1,p defined as the completion of C∞ 0 (Ω) with respect to the norm 0 (Ω), The space D1,p D1,p 0 (Ω) is a natural setting for studying variational problems of the type 0 (Ω) has been first studied by Deny and Lions in [13], among others. We recall that (cid:18)(cid:90) Ω u (cid:55)→ ∇up dx . (cid:19) 1 p (cid:90) Ω (cid:26) 1 (cid:90) p Ω inf ∇up dx − f u dx , supplemented with Dirichlet boundary conditions, in absence of regularity assumptions on the boundary ∂Ω. In the same way, the space Ds,p 0 (Ω) is the natural framework for studying minimiza- tion problems containing functionals of the type (1.1) 1 p u(x) − u(y)p x − yN +s p dx dy − RN×RN f u dx, Ω in presence of nonlocal Dirichlet boundary conditions, i.e. the values of u are prescribed on the whole complement RN \ Ω. Observe that even if this kind of boundary conditions may look weird, these are the correct ones when dealing with energies (1.1), which take into account interactions "from infinity". that for u ∈ C∞ 0 (Ω) and Ds,p 0 (Ω), we have (see [5] and [26, Corollary 1.3]) The connection between the two spaces D1,p 0 (Ω) is better appreciated by recalling (1 − s) lim s(cid:37)1 RN×RN u(x) − u(y)p x − yN +s p dx dy = αN,p ∇up dx, with αN,p = (cid:104)ω, e1(cid:105)p dHN−1(ω), SN−1 e1 = (1, 0, . . . , 0). On the other hand, as s (cid:38) 0 we have (see [24, Theorem 3]) s lim s(cid:38)0 RN×RN u(x) − u(y)p x − yN +s p dx dy = βN,p up dx, (cid:90)(cid:90) (cid:90)(cid:90) (cid:90) (cid:90)(cid:90) 1 p (cid:27) (cid:90) (cid:90) Ω (cid:90) Ω 4 with BRASCO AND SALORT βN,p = 2 N ωN , p and ωN is the volume of the N−dimensional unit ball. These two results reflect the "interpolative" nature of the space Ds,p lation space X s,p Indeed, one of our goals is to determine whether Ds,p 0 (Ω), which will be however discussed in more detail in the sequel. (Ω) defined as the completion of C∞ 0 (Ω) coincides or not with the real interpo- 0 (Ω) with respect to the norm 0 (cid:32)(cid:90) +∞ (cid:32) (cid:33)p (cid:33) 1 p K(t, u, Lp(Ω),D1,p 0 (Ω) := (cid:107)u(cid:107)X s,p 0 (Ω)) is the K−functional associated to the spaces Lp(Ω) and D1,p 0 (Ω)) dt t ts 0 . 0 (Ω), see Here K(t,·, Lp(Ω),D1,p Section 3 below for more details. In particular, we will be focused on obtaining double-sided norm inequalities leading to answer our initial question, i.e. estimates of the form [u]W s,p(RN ) ≤ (cid:107)u(cid:107)X s,p 0 (Ω) ≤ C [u]W s,p(RN ), u ∈ C∞ 0 (Ω). 1 C Moreover, we compute carefully the dependence on the parameter s of the constant C. Indeed, we will see that C can be taken independent of s. 1.3. Results. We now list the main achievements of our discussion: 1. the space Ds,p 0 (Ω) is always larger than X s,p 0 (Ω) (see Proposition 4.1) and they do not coincide for general open sets, as we exhibit with an explicit example (see Example 4.4); 2. they actually coincide on a large class of domains, i.e. bounded convex sets (Theorem 4.7), convex cones (Corollary 4.8), Lipschitz sets (Theorem 4.10); are equivalent for the classes of sets at point 2 (Theorem 6.1). More precisely, by setting 3. the Poincar´e constants for the embeddings Ds,p 0 (Ω) (cid:44)→ Lp(Ω) and (cid:110) D1,p 0 (Ω) (cid:44)→ Lp(Ω), (cid:111) : (cid:107)u(cid:107)Lp(Ω) = 1 , 0 < s < 1, λs p(Ω) = inf u∈C∞ 0 [u]p W s,p(Ω) (cid:26)(cid:90) (cid:17)s ≤ s (1 − s) λs Ω 0 λ1 p(Ω) = inf u∈C∞ (cid:16) 1 C λ1 p(Ω) ∇up dx : (cid:107)u(cid:107)Lp(Ω) = 1 , (cid:16) p(Ω) ≤ C λ1 p(Ω) . (cid:27) (cid:17)s and we have Moreover, on convex sets the constant C > 0 entering in the relevant estimate is universal, i.e. it depends on N and p only. On the other hand, we show that this equivalence fails if we drop any kind of regularity assumptions on the sets (see Remark 6.3). As a byproduct of our discussion, we also highlight some weird and unnatural behaviors of the interpolation space X s,p (Ω): • the "extension by zero" operator X s,p sets (see Remark 4.5). This is in contrast with what happens for the spaces Lp(Ω), D1,p and Ds,p (RN ) is not continuous for general open 0 (Ω) (Ω) (cid:44)→ X s,p 0 0 0 0 (Ω); • the sharp Poincar´e interpolation constant HOMOGENEOUS SOBOLEV SPACES (cid:111) , 5 0 < s < 1 (cid:110)(cid:107)u(cid:107)pX s,p 0 Λs p(Ω) = inf u∈C∞ 0 (Ω) : (cid:107)u(cid:107)Lp(Ω) = 1 (Ω) is sensitive to removing sets with zero capacity. In other words, if we remove a compact set E (cid:98) Ω having zero capacity in the sense of X s,p p(Ω \ E) > Λs Λs Again, this is in contrast with the case of D1,p 0 p(Ω). 0 (Ω) and Ds,p (Ω), then (see Lemma 5.4) 0 (Ω). Remark 1.1. As recalled at the beginning, nowadays there is a huge literature on Sobolev spaces of fractional order. Nevertheless, to the best of our knowledge, a detailed discussion on the space Ds,p 0 (Ω) in connection with interpolation theory seems to be missing. For this reason, we believe that our discussion is of independent interest. We also point out that for Sobolev spaces of functions not necessarily vanishing at the bound- ary, there is a very nice paper [11] by Chandler-Wilde, Hewett and Moiola comparing "concrete" constructions with the interpolation one. 1.4. Plan of the paper. In Section 2 we present the relevant Sobolev spaces, constructed with the concrete approach based on the so-called Sobolev-Slobodeckiı norms. Then in Section 3 we introduce the homogeneous interpolation space we want to work with. Essentially, no previous knowledge of interpolation theory is necessary. The comparison between the concrete space and the interpolation one is contained in Section 4. This in turn is divided in three subsections, each one dealing with a different class of open sets. We point out here that we preferred to treat convex sets separately from Lipschitz sets, for two reasons: the first one is that for convex sets the comparison between the two spaces can be done "by hands", without using any extension theorem. This in turn permits to have a better control on the relevant constants entering in the estimates. The second one is that in proving the result for Lipschitz sets, we actually use the result for convex sets. In order to complement the comparison between the two spaces, in Section 5 we compare the two relevant notions of capacity, naturally associated with the norms of these spaces. Finally, Section 6 compares the Poincar´e constants. The paper ends with 3 appendices: the first one contains the construction of a counter-example used throughout the whole paper; the second one proves a version of the one-dimensional Hardy inequality; the last one contains a geometric expedient result dealing with convex sets. Acknowledgments. The first author would like to thank Yavar Kian and Antoine Lemenant for useful discussions on Stein's and Jones' extension theorems. Simon Chandler-Wilde is gratefully acknowledged for some explanations on his paper [11]. This work started during a visit of the second author to the University of Ferrara in October 2017. 2. Preliminaries 2.1. Basic notation. In what follows, we will always denote by N the dimension of the ambient space. For an open set Ω ⊂ RN , we indicate by Ω its N−dimensional Lebesgue measure. The symbol Hk will stand for the k−dimensional Hausdorff measure. Finally, we set BR(x0) = {x ∈ RN : x − x0 < R}, and ωN = B1(0). 6 BRASCO AND SALORT 2.2. Sobolev spaces. For 1 ≤ p < ∞ and an open set Ω ⊂ RN , we use the classical definition This is a Banach space endowed with the norm W 1,p(Ω) := u ∈ Lp(Ω) : ∇up dx < +∞ . (cid:90) Ω (cid:16)(cid:107)u(cid:107)p (cid:27) (cid:17) 1 p . (cid:107)u(cid:107)W 1,p(Ω) = Lp(Ω) + (cid:107)∇u(cid:107)p Lp(Ω) (cid:26) (cid:90) Ω We also denote by D1,p with respect to the norm 0 (Ω) the homogeneous Sobolev space, defined as the completion of C∞ 0 (Ω) If the open set Ω ⊂ RN supports the classical Poincar´e inequality u (cid:55)→ (cid:107)∇u(cid:107)Lp(Ω). then D1,p We will set 0 (Ω) is indeed a functional space and it coincides with the closure in W 1,p(Ω) of C∞ 0 (Ω). up dx ≤ ∇up dx, for every u ∈ C∞ 0 (Ω), (cid:110)(cid:107)∇u(cid:107)p (cid:111) Lp(Ω) : (cid:107)u(cid:107)Lp(Ω) = 1 . λ1 p(Ω) = inf u∈C∞ 0 (Ω) (cid:90) c Ω It occurs λ1 p(Ω) = 0 whenever Ω does not support such a Poincar´e inequality. Remark 2.1. We remark that one could also consider the space 0 0 (Ω) := {u ∈ W 1,p(RN ) : u = 0 a.e. in RN \ Ω}. 0 (Ω) (cid:44)→ Lp(Ω). W 1,p It is easy to see that D1,p 0 (Ω) ⊂ W 1,p continuous, then both spaces are known to coincide, thanks to the density of C∞ see [20, Theorem 1.4.2.2]. 2.3. A homogeneous Sobolev–Slobodeckiı space. Given 0 < s < 1 and 1 ≤ p < ∞, the fractional Sobolev space W s,p(RN ) is defined as If in addition ∂Ω is (Ω), (Ω), whenever D1,p 0 (Ω) in W 1,p W s,p(RN ) :=(cid:8)u ∈ Lp(RN ) : [u]W s,p(RN ) < +∞(cid:9) , (cid:19) 1 where the Sobolev–Slobodeckiı seminorm [· ]W s,p(RN ) is defined as u(x) − u(y)p x − yN +s p dx dy [u]W s,p(RN ) := 0 p This is a Banach space endowed with the norm (cid:18)(cid:90)(cid:90) (cid:16)(cid:107)u(cid:107)p RN×RN (cid:107)u(cid:107)W s,p(RN ) = Lp(RN ) + [u]p W s,p(RN ) (cid:17) 1 p . . In what follows, we need to consider nonlocal homogeneous Dirichlet boundary conditions, outside an open set Ω ⊂ RN . In this setting, it is customary to consider the homogeneous Sobolev–Slobodeckiı space Ds,p 0 (Ω). The latter is defined as the completion of C∞ 0 (Ω) with respect to the norm u (cid:55)→ [u]W s,p(RN ). Observe that the latter is indeed a norm on C∞ following Poincar´e inequality (cid:90)(cid:90) (cid:90) c Ω up dx ≤ u(x) − u(y)p x − yN +s p dx dy, RN×RN for every u ∈ C∞ 0 (Ω), 0 (Ω). Whenever the open set Ω ⊂ RN admits the HOMOGENEOUS SOBOLEV SPACES 7 we get that Ds,p with the closure in W s,p(RN ) of C∞ 0 (Ω) is a functional space continuously embedded in Lp(Ω). In this case, it coincides We also define λs p(Ω) = inf u∈C∞ 0 (Ω) 0 (Ω) with the norm 0 (Ω). We endow the space Ds,p (cid:107)u(cid:107)Ds,p 0 (Ω) := [u]W s,p(RN ). (cid:110)(cid:107)u(cid:107)pDs,p (cid:111) 0 (Ω) : (cid:107)u(cid:107)Lp(Ω) = 1 , i.e. this is the sharp constant in the relevant Poincar´e inequality. Some embedding properties of the space Ds,p 0 (Ω) are investigated in [18]. Remark 2.2. As in the local case, one could also consider the space (Ω) := {u ∈ W s,p(RN ) : u = 0 a.e. in RN \ Ω}. W s,p It is easy to see that Ds,p 0 (Ω) ⊂ W s,p is continuous, then both spaces are known to coincide, again thanks to the density of C∞ W s,p 0 (Ω) (cid:44)→ Lp(Ω). As before, whenever ∂Ω 0 (Ω) in (Ω), whenever Ds,p (Ω), see [20, Theorem 1.4.2.2]. 0 0 0 2.4. Another space of functions vanishing at the boundary. Another natural fractional Sobolev space of functions "vanishing at the boundary" is given by the completion of C∞ 0 (Ω) with respect to the localized norm (cid:18)(cid:90)(cid:90) [u]W s,p(Ω) = u(x) − u(y)p x − yN +s p dx dy Ω×Ω (cid:19) 1 p . We will denote this space by Ds,p(Ω). We recall the following Lemma 2.3. Let 1 < p < ∞ and 0 < s < 1. For every Ω ⊂ RN open bounded Lipschitz set, we have: • if s p > 1, then • if s p ≤ 1, then there exists a sequence {un}n∈N ⊂ C∞ 0 (Ω) such that Ds,p 0 (Ω) = Ds,p(Ω); (cid:107)un(cid:107) Ds,p(Ω) (cid:107)un(cid:107)Ds,p 0 (Ω) lim n→∞ = 0. Proof. The proof of the first fact is contained in [7, Theorem B.1]. As for the case s p ≤ 1, in [15, Section 2] Dyda constructed a sequence {un}n∈N ⊂ C∞ 0 (Ω) such that By observing that for such a sequence we have n→∞(cid:107)un(cid:107) Ds,p(Ω) = 0 lim n→∞(cid:107)un − 1Ω(cid:107)Lp(Ω) = 0. lim 0 (Ω) ≥(cid:16) and (cid:17) 1 p n→∞(cid:107)un(cid:107)Ds,p lim λs p(Ω) n→∞(cid:107)un(cid:107)Lp(Ω) = lim (cid:16) p(Ω)Ω(cid:17) 1 p λs , we get the desired conclusion. In the inequality above, we used that λs set, thanks to [8, Corollary 5.2]. p(Ω) > 0 for an open bounded (cid:3) Remark 2.4. Clearly, we always have (cid:107)u(cid:107) Ds,p(Ω) ≤ (cid:107)u(cid:107)Ds,p 0 (Ω), for every u ∈ C∞ 0 (Ω). 8 BRASCO AND SALORT As observed in [16], the reverse inequality (2.1) (cid:107)u(cid:107)Ds,p 0 (Ω) ≤ C (cid:107)u(cid:107) Ds,p(Ω), (cid:33) (cid:32)(cid:90) is equivalent to the validity of the Hardy-type inequality dx ≤ C x − y−N−s p y u(x)p (cid:90)(cid:90) (cid:90) Ω RN\Ω for every u ∈ C∞ 0 (Ω), u(x) − u(y)p x − yN +s p dx dy. Ω×Ω A necessary and sufficient condition for this to happen is proved in [16, Proposition 2]. We also observe that the failure of (2.1) implies that in general the "extension by zero" operator T0 : Ds,p(Ω) → Ds,p(RN ), is not continuous. We refer to [16] for a detailed discussion of this issue. Remark 2.5. The space Ds,p(Ω) is quite problematic in general, especially in the case s p ≤ 1 where it may fail to be a functional space. A more robust variant of this space is By definition, this is automatically a functional space, continuously contained in W s,p(Ω). It is a classical fact that if Ω is a bounded open set with smooth boundary, then 0 (Ω) in W s,p(Ω)". (cid:101)Ds,p(Ω) = "closure of C∞ (cid:101)Ds,p(Ω) = W s,p(Ω), (cid:101)Ds,p(Ω) = W s,p (cid:101)Ds,p(Ω) = Ds,p 0 (Ω), (Ω), 0 for s p < 1, for s p > 1, for s p (cid:54)= 1, and see Theorem [32, Theorem 3.4.3]. Moreover, we also have see for example [7, Proposition B.1]. 3. An interpolation space Let Ω ⊂ RN be an open set. If X(Ω) and Y (Ω) are two normed vector spaces containing C∞ as a dense subspace, we define for every t > 0 and u ∈ C∞ 0 (Ω) (cid:110)(cid:107)u − v(cid:107)X(Ω) + t(cid:107)v(cid:107)Y (Ω) (cid:111) 0 (Ω) the K−functional . (3.1) We are interested in the following specific case: let us take 0 < s < 1 and 1 < p < ∞, we choose K(t, u, X(Ω), Y (Ω)) := inf v∈C∞ 0 (Ω) X(Ω) = Lp(Ω) and Y (Ω) = D1,p 0 (Ω). Then we use the notation (cid:107)u(cid:107)X s,p 0 (Ω) := (cid:32)(cid:90) +∞ (cid:32) 0 It is standard to see that this is a norm on C∞ the completion of C∞ 0 (Ω) with respect to this norm. (cid:33)p (cid:33) 1 p dt t K(t, u, Lp(Ω),D1,p 0 (Ω)) , u ∈ C∞ 0 (Ω). ts 0 (Ω), see [4, Section 3.1]. We will indicate by X s,p 0 (Ω) The first result is the Poincar´e inequality for the interpolation space X s,p (Ω). The main focus is on the explicit dependence of the constant on the local Poincar´e constant λ1 p. 0 HOMOGENEOUS SOBOLEV SPACES 9 Lemma 3.1. Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open set. Then for every u ∈ C∞ (3.2) 0 (Ω) we have (cid:0)λ1 p(Ω)(cid:1)s (cid:107)u(cid:107)p (cid:90) +∞ (cid:107)u(cid:107)p Lp(Ω) (cid:46) 0 Lp(Ω) ≤ s (1 − s)(cid:107)u(cid:107)pX s,p (cid:18) K(t, u, Lp(Ω), Lp(Ω)) 0 (Ω). (cid:19)p dt ts , t Proof. We proceed in two stages: we first prove that and then we show that the last integral is estimated from above by the norm X s,p First stage. Let us take u ∈ C∞ 0 0 (Ω), for every t ≥ 1 and v ∈ C∞ 0 (Ω) (cid:107)u(cid:107)Lp(Ω) ≤ (cid:107)u − v(cid:107)Lp(Ω) + t(cid:107)v(cid:107)Lp(Ω). (Ω). By taking the infimum, we thus get (cid:107)u(cid:107)Lp(Ω) ≤ K(t, u, Lp(Ω), Lp(Ω)). By integrating with respect to the singular measure dt/t, we then get (cid:90) +∞ (cid:18) K(t, u, Lp(Ω), Lp(Ω)) (cid:19)p dt (cid:90) +∞ 1 (3.3) We now pick 0 < t < 1, by triangle inequality we get for every v ∈ C∞ t(cid:107)u(cid:107)Lp(Ω) ≤ t(cid:107)u − v(cid:107)Lp(Ω) + t(cid:107)v(cid:107)Lp(Ω) ≤ (cid:107)u − v(cid:107)Lp(Ω) + t(cid:107)v(cid:107)Lp(Ω). Lp(Ω) ts t 1 0 (Ω) ≥ t−s p (cid:107)u(cid:107)p (cid:107)u(cid:107)p Lp(Ω) s p . dt t = By taking the infimum over v ∈ C∞ 0 (Ω), we obtain for u ∈ C∞ t(cid:107)u(cid:107)Lp(Ω) ≤ K(t, u, Lp(Ω), Lp(Ω)). 0 (Ω) and 0 < t < 1 By integrating again, we get this time (cid:90) 1 (cid:18) K(t, u, Lp(Ω), Lp(Ω)) (3.4) 0 By summing up (3.3) and (3.4), we get the estimate 0 ts (cid:19)p dt (cid:90) +∞ t ≥ (cid:90) 1 (cid:18) K(t, u, Lp(Ω), Lp(Ω)) tp−s p (cid:107)u(cid:107)p Lp(Ω) dt t (3.5) (cid:107)u(cid:107)p Lp(Ω) ≤ s (1 − s) (cid:107)u(cid:107)p (1 − s) p Lp(Ω) . = (cid:19)p dt . 0 ts t Second stage. Given u ∈ C∞ (3.2) trivially holds. By definition of λ1 0 (Ω), we take v ∈ C∞ p(Ω) we have that 0 (Ω). We can suppose that λ1 p(Ω) > 0, otherwise (cid:107)u − v(cid:107)Lp(Ω) + t(cid:107)v(cid:107)Lp(Ω) ≤ (cid:107)u − v(cid:107)Lp(Ω) + t (λ1 p (cid:107)∇v(cid:107)Lp(Ω). − 1 p(Ω)) If we recall the definition (3.1) of the K−functional, we get (cid:107)u − v(cid:107)Lp(Ω) + K(t, u, Lp(Ω), Lp(Ω))p ≤ (cid:32) t (cid:107)∇v(cid:107)Lp(Ω) 1 p (λ1 p(Ω)) (cid:32) (cid:33)p (cid:33)p , and by taking infimum over v ∈ C∞ 0 (Ω) and multiplying by t−s p, we get t−s pK(t, u, Lp(Ω), Lp(Ω))p ≤ t−s p K t (λ1 p(Ω)) 1 p , u , Lp(Ω),D1,p 0 (Ω) . 10 BRASCO AND SALORT We integrate over t > 0, by performing the change of variable τ = t/(λ1 p(Ω)) 1 p we get By using this in (3.5), we prove the desired inequality (3.2). We will set (cid:90) +∞ 0 t ts ≤ (cid:18) K(t, u, Lp(Ω), Lp(Ω)) (cid:19)p dt (cid:110)(cid:107)u(cid:107)pX s,p (cid:17)s ≤ s (1 − s) Λs p(Ω) = inf u∈C∞ (cid:16) 0 (Ω) λ1 p(Ω) 0 Λs 0 1 (λ1 p(Ω))s(cid:107)u(cid:107)pX s,p (cid:111) (Ω) : (cid:107)u(cid:107)Lp(Ω) = 1 , (Ω). (cid:3) i.e. this is the sharp constant in the relevant Poincar´e inequality. As a consequence of (3.2), we obtain (3.6) Proposition 3.2 (Interpolation inequality). Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open set. For every u ∈ C∞ p(Ω). (3.7) In particular, we also obtain 0 (Ω) we have s (1 − s)(cid:107)u(cid:107)pX s,p 0 (3.8) Proof. We can assume that u (cid:54)≡ 0, otherwise there is nothing prove. K−functional K(t, u, Lp(Ω),D1,p 0 (Ω)) we take v = τ u for τ > 0, thus we obtain . s (1 − s) Λs In the definition of the K(t, u, Lp(Ω),D1,p 0 (Ω)) ≤ inf By integrating for t > 0, we get (cid:107)u(cid:107)pX s,p 0 (Ω) ≤ (cid:90) +∞ 0 (Ω) ≤ (cid:107)u(cid:107)p (1−s) Lp(Ω) (cid:107)∇u(cid:107)s p Lp(Ω). (cid:105) λ1 p(Ω) (cid:17)s p(Ω) ≤(cid:16) (cid:104)1 − τ(cid:107)u(cid:107)Lp(Ω) + t τ (cid:107)∇u(cid:107)Lp(Ω) (cid:110)(cid:107)u(cid:107)Lp(Ω), t(cid:107)∇u(cid:107)Lp(Ω) (cid:110)(cid:107)u(cid:107)p (cid:90) (cid:107)u(cid:107)Lp (Ω) (cid:90) +∞ Lp(Ω), tp (cid:107)∇u(cid:107)p tp (1−s) dt t (cid:111) (cid:111) (cid:107)∇u(cid:107)Lp (Ω) dt t Lp(Ω) ts p 0 . τ >0 = min min = (cid:107)∇u(cid:107)p Lp(Ω) + (cid:107)u(cid:107)p Lp(Ω) (cid:107)u(cid:107)Lp (Ω) (cid:107)∇u(cid:107)Lp (Ω) t−s p dt t (cid:20) (cid:21) = (cid:107)u(cid:107)p (1−s) Lp(Ω) (cid:107)∇u(cid:107)s p Lp(Ω) 1 p (1 − s) + 1 s p . We thus get the desired conclusion (3.7). The estimate (3.8) easily follows from the definition of (cid:3) Poincar´e constant. From (3.6) and (3.8), we get in particular the following Corollary 3.3 (Equivalence of Poincar´e constants). Let 1 < p < ∞ and 0 < s < 1. For every Ω ⊂ RN open set we have (cid:16) (cid:17)s s (1 − s) Λs p(Ω) = λ1 p(Ω) . HOMOGENEOUS SOBOLEV SPACES 11 In particular, there holds D1,p 0 (Ω) (cid:44)→ Lp(Ω) Remark 3.4 (Extensions by zero in X s,p zero" operators 0 ⇐⇒ X s,p 0 (Ω) (cid:44)→ Lp(Ω). ). We observe that by interpolating the "extension by T0 : D1,p 0 (Ω) → D1,p 0 (RN ) and T0 : Lp(Ω) → Lp(RN ) which are both continuous, one obtains the same result for the interpolating spaces. In other words, we have (cid:107)u(cid:107)pX s,p 0 (RN ) ≤ (cid:107)u(cid:107)pX s,p 0 (Ω), for every u ∈ C∞ 0 (Ω). 0 (Ω) ⊂ C∞ This can be also seen directly: it is sufficient to observe that C∞ ately get 0 (RN ), thus we immedi- K(t, u, Lp(RN ),D1,p 0 (RN )) ≤ K(t, u, Lp(Ω),D1,p 0 (Ω)), since in the K−functional on the left-hand side the infimum is performed on a larger class. By integrating, we get the conclusion. However, differently from the case of D1,p 0 (Ω), Lp(Ω) and Ds,p 0 (Ω), in general for u ∈ C∞ 0 (Ω) we have (cid:107)u(cid:107)pX s,p 0 (RN ) < (cid:107)u(cid:107)pX s,p 0 (Ω). In other words, even if u ≡ 0 outside Ω, passing from Ω to RN has an impact on the interpolation norm. Actually, if Ω has not smooth boundary, the situation can be much worse than this. We refer to Remark 4.5 below. 4. Interpolation VS. Sobolev-Slobodeckiı 0 (Ω) and X s,p 4.1. General sets. We want to compare the norms of Ds,p simplest estimate, which is valid for every open set. Proposition 4.1 (Comparison of norms I). Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open set, then for every u ∈ C∞ (Ω). We start with the 0 0 (Ω) we have 1 (4.1) (cid:107)u(cid:107)pDs,p In particular, we have the continuous inclusion X s,p Proof. To prove (4.1), we take h ∈ RN \ {0} and ε > 0, then there exists v ∈ C∞ 0 (Ω) ≤ (cid:107)u(cid:107)pX s,p (Ω) ⊂ Ds,p 2p (1−s) N ωN 0 (Ω). (Ω). 0 0 0 (Ω) such that (4.2) (cid:107)u − v(cid:107)Lp(Ω) + h(cid:107)v(cid:107)D1,p 0 (Ω) ≤ (1 + ε) K(h, u, Lp(Ω),D1,p 0 (Ω)). 12 BRASCO AND SALORT Thus for h (cid:54)= 0 we get1 (cid:18)(cid:90) u(x + h) − u(x)p hN +s p RN (cid:19) 1 p ≤ dx (cid:18)(cid:90) (cid:18)(cid:90) RN u(x + h) − v(x + h) − u(x) + v(x)p hN +s p (cid:19) 1 p v(x + h) − v(x)p hN +s p dx p −s (cid:107)u − v(cid:107)Lp(Ω) p +s (cid:107)∇v(cid:107)Lp(Ω) p −s (cid:18) (cid:107)u − v(cid:107)Lp(Ω) + + RN ≤ 2h− N + h1− N ≤ 2h− N h 2 0 (Ω) (cid:107)v(cid:107)D1,p (cid:33)p 0 (Ω)) (cid:19) 1 p dx (cid:19) . 1 hN . (cid:33)p By using (4.2), we then obtain u(x + h) − u(x)p hN +s p dx ≤ 2p (1 + ε)p K(h/2, u, Lp(Ω),D1,p hs (cid:90) RN (cid:90)(cid:90) We now integrate with respect to h ∈ RN . This yields u(x + h) − u(x)p hN +s p RN×RN dx dh ≤ 2p (1 + ε)p RN (cid:32) (cid:90) (cid:32) (cid:90) +∞ (cid:32) K(h/2, u, Lp(Ω),D1,p 0 (Ω)) hs K(t/2, u, Lp(Ω),D1,p 0 (Ω)) 0 ts dh hN (cid:33)p dt t . = 2p (1 + ε)p N ωN By making the change of variable t/2 = τ and exploiting the arbitariness of ε > 0, we eventually (cid:3) reach the desired estimate. Corollary 4.2 (Interpolation inequality for Ds,p 0 ). Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open set. For every u ∈ C∞ (4.3) 0 (Ω) ≤ 2p (1−s) N ωN (cid:107)u(cid:107)p (1−s) s (1 − s)(cid:107)u(cid:107)pDs,p Lp(Ω) (cid:107)∇u(cid:107)s p 0 (Ω) we have (cid:3) Proof. It is sufficient to combine Propositions 4.1 and 3.2. Remark 4.3. For p (cid:38) 1, the previous inequality becomes [7, Proposition 4.2]. In this case, the constant in (4.3) is sharp for N = 1. Lp(Ω). For a general open set Ω ⊂ RN , the converse of inequality (4.1) does not hold. This means that in general we have X s,p 0 and X s,p 0 (Ω) (cid:54)= Ds,p 0 (Ω), 1In the second inequality, we use the classical fact ϕ(x + h) − ϕ(x)p dx = (cid:104)∇ϕ(x + t h), h(cid:105) dt (cid:90) RN 0 (Ω) (Ω) ⊂ Ds,p (cid:90) (cid:12)(cid:12)(cid:12)(cid:12)(cid:90) 1 (cid:90) 1 (cid:90) (cid:18)(cid:90) (cid:90) 1 RN 0 0 0 RN RN ≤ hp = hp (cid:12)(cid:12)(cid:12)(cid:12)p dx (cid:19) ∇ϕ(x + t h)p dt dx ∇ϕ(x + t h)p dx dt = hp (cid:107)∇ϕ(cid:107)Lp(RN ). HOMOGENEOUS SOBOLEV SPACES 13 the inclusion being continuous. We use the construction of Appendix A in order to give a counter- example. Example 4.4. With the notation of Appendix A, let us take2 0 (E) such that , (F + z) (cid:33) (cid:32) (cid:91) 0 ((cid:101)Ωn) ⊂ C∞ p((cid:101)Ωn) + ε (cid:20) (cid:17) (cid:16) Ω + z = E = RN \ z∈ZN For every ε > 0, we take un ∈ C∞ [un]p W s,p(RN ) < λs Here the set (cid:101)Ωn is defined by (cid:91) (cid:101)Ωn = z∈ZN n On the other hand, we have (cid:107)un(cid:107)pX s,p 0 (E) ≥ −n − 1 2 (cid:17)s (cid:90) E (cid:16) λ1 p(E) s (1 − s) (cid:20) (cid:21)N−1 × {0}. , 1 4 − 1 4 with F = unp dx = 1. and , n + 1 2 E (cid:90) (cid:21)N \ (cid:91) (cid:16) z∈ZN n (F + z). (cid:17)s unp dx ≥ µp(Q; F ) s (1 − s) , where we also used (A.4). By Lemma A.1, we have that λs p(Ωn) converges to 0 for s p < 1, so that n→∞ (cid:107)un(cid:107)pX s,p lim inf 0 (E) ≥ 1 C and lim sup n→∞ [un]p W s,p(RN ) ≤ ε. Thus by the arbitrariness of ε, we obtain (cid:107)un(cid:107)pDs,p (cid:107)un(cid:107)pX s,p 0 0 (E) (E) lim n→∞ = 0, for 1 < p < ∞ and s < 1 p . Remark 4.5 (Extensions by zero in X s,p {un}n∈N ⊂ C∞ 0 0 (E) as in Example 4.4. We have seen that (Ω)...reprise). We take the set E ⊂ RN and the sequence By using Proposition 4.6 we obtain lim n→∞ lim n→∞ (cid:107)un(cid:107)X s,p (cid:107)un(cid:107)Ds,p 0 0 (E) (E) = +∞. (cid:107)un(cid:107)X s,p (cid:107)un(cid:107)X s,p 0 0 (E) (RN ) = +∞, as well, still for s p < 1. Thus the "extension by zero" operator in general is not continuous. 2In dimension N = 1, we simply take E = R \ Z. 14 BRASCO AND SALORT 4.2. Convex sets. We now prove the converse of (4.1), under suitable assumptions on Ω. We start with the case of a convex set. The case Ω = RN is simpler and instructive, thus we give a separate statement. The proof can be found for example in [28, Lemma 35.2]. We reproduce it, for the reader's convenience. We also single out an explicit determination of the constant. Proposition 4.6 (Comparison of norms II: RN ). Let 1 < p < ∞ and 0 < s < 1. For every u ∈ C∞ 0 (RN ) we have (cid:107)u(cid:107)pX s,p In particular, we have that Ds,p Proof. Let u ∈ C∞ 0 (RN ), we set 0 2p N ωN (cid:107)u(cid:107)pDs,p 0 (RN ). (cid:17)p N (N + 1) (RN ). (RN ) ≤(cid:16) 0 0 (RN ) = X s,p (cid:18)(cid:90) (cid:19) 1 p U (h) = RN u(x + h) − u(x)p dx , h ∈ RN , and observe that by construction (cid:90) RN We also define U () = 1 N ωN N−1 thus by Jensen's inequality we have (cid:90) +∞ (cid:18) U (cid:19)p (4.4) 0 s W s,p(RN ). U (h)p hN +s p dh = [u]p (cid:90) {h∈RN : h=} U dHN−1,  > 0, (cid:32)(cid:90) (cid:90) +∞ (cid:90) 0 d  ≤ 1 N ωN 1 = N ωN RN {h∈RN : h=} 1 U (h)p hN +s p dh = N ωN (cid:33) d N +s p U p dHN−1 [u]p W s,p(RN ). We now take the compactly supported Lipschitz function ψ(x) = N + 1 ωN (1 − x)+, where (· )+ stands for the positive part. Observe that ψ has unit L1 norm, by construction. We then define (cid:16) x (cid:17) t From the definition of the K−functional, we get ψt(x) = 1 tN ψ , for t > 0. K(t, u, Lp(RN ),D1,p 0 (RN )) ≤ (cid:107)u − ψt ∗ u(cid:107)Lp(Ω) + t(cid:107)∇ψt ∗ u(cid:107)Lp(Ω), by observing that ψt ∗ u ∈ C∞ Minkowski inequality we get HOMOGENEOUS SOBOLEV SPACES 15 0 (RN ). We estimate the two norms separately: for the first one, by (cid:13)(cid:13)(cid:13)(cid:13)(cid:90) (cid:90) (cid:90) RN (cid:18)(cid:90) RN RN [u(·) − u(· − y)] ψt(y) dy u(x) − u(x − y)p dx U (−y) ψt(y) dy ≤ N + 1 ωN tN RN (cid:13)(cid:13)(cid:13)(cid:13)Lp(RN ) (cid:19) 1 (cid:90) p Bt(0) (cid:107)u − ψt ∗ u(cid:107)Lp(RN ) = ≤ = = ψt(y) dy U (−y) dy (cid:90) t 0 N (N + 1) U N−1 d ≤ N (N + 1) t U d. tN (cid:90) ∇ψt(y) dy = 0, ∇ψt ∗ u = (∇ψt) ∗ u = ∇ψt(y) [u(x − y) − u(x)] dy. For the norm of the gradient, we first observe that thus we can write RN Consequently, by Minkowski inequality we get (cid:13)(cid:13)(cid:13)(cid:13)(cid:90) (cid:90) RN (cid:18)(cid:90) (cid:107)∇ψt ∗ u(cid:107)Lp(RN ) = ≤ (cid:90) t 0 (cid:90) RN dt t t (cid:55)→ ≤(cid:16) (cid:90) t dt t 0 ≤ ≤ ∇ψt(y) [u(· − y) − u(·)] dy (cid:90) RN RN ≤ N + 1 ωN tN +1 Bt(0) u(x − y) − u(x)p dx p ∇ψt(y) dy U (−y) dy ≤ N (N + 1) U d. (cid:90) t 0 (cid:13)(cid:13)(cid:13)(cid:13)Lp(RN ) (cid:19) 1 t2 (cid:90) t (cid:18)(cid:90) t (cid:17)p (cid:90) T 0 U d. 0 0 In conclusion, we obtained for every t > 0 K(t, u, Lp(RN ),D1,p (4.5) (cid:32) (cid:90) T K(t, u, Lp(RN ),D1,p 0 (RN )) 0 ts 0 (RN )) ≤ 2 N (N + 1) (cid:33)p t 2 N (N + 1) If we integrate on (0, T ), the previous estimate gives If we now use Lemma B.1 with α = p + s p for the function (cid:19)p U d t−p−s p dt t . we get (cid:32) (cid:90) T 0 K(t, u, Lp(Ω),D1,p 0 (Ω)) ts (cid:33)p U d, (cid:18) 2 N (N + 1) (cid:18) 2 N (N + 1) s + 1 (cid:19)p (cid:90) T (cid:19)p 0 1 (cid:18) U (cid:19)p ts dt t s + 1 N ωN [u]p W s,p(RN ), 16 BRASCO AND SALORT where we used (4.4) in the second inequality. By letting T going to +∞, we get the desired (cid:3) estimate. We denote by the inradius of an open set Ω ⊂ RN . This is the radius of the largest open ball inscribed in Ω. We introduce the eccentricity of an open bounded set Ω ⊂ RN , defined by RΩ = sup x∈Ω dist(x, ∂Ω), E(Ω) = diam (Ω) 2 RΩ . By generalizing the construction used in [9, Lemma A.6] for a ball, we have the following. Theorem 4.7 (Comparison of norms II: bounded convex sets). Let 1 < p < ∞ and 0 < s < 1. If Ω ⊂ RN is an open bounded convex set, then for every u ∈ C∞ (4.6) 0 (Ω), for a constant C = C(N, p,E(Ω)) > 0. In particular, we have X s,p Proof. The proof runs similarly to that of Proposition 4.6 for RN , but now we have to pay attention to boundary issues. Indeed, the function ψt ∗ u is not supported in Ω, unless t is sufficiently small, depending on u itself. In order to avoid this, we need to perform a controlled scaling of the function. By keeping the same notation as in the proof of Proposition 4.6, we need the following modification: we take a point x0 ∈ Ω such that (Ω) ≤ C (cid:107)u(cid:107)pDs,p (Ω) = Ds,p 0 (Ω) we have (cid:107)u(cid:107)pX s,p 0 0 (Ω). 0 Without loss of generality, we can assume that x0 = 0. Then we define the rescaled function dist(x0, ∂Ω) = RΩ. (cid:19) (cid:18) RΩ RΩ − t ut = u x , 0 < t < RΩ 2 . We observe that and by Lemma C.1, we have dist support(ut) = (cid:19) Ω, ∂Ω ≥ (cid:18) RΩ − t RΩ Ω, RΩ − t RΩ (cid:18) 1 − RΩ − t RΩ (cid:19) RΩ = t. This implies that We can now estimate the K−functional by using the choice v = ψt ∗ ut, that is ψt ∗ ut ∈ C∞ 0 (Ω), for every 0 < t < RΩ 2 . K(t, u, Lp(Ω),D1,p 0 (Ω)) ≤ (cid:107)u − ψt ∗ ut(cid:107)Lp(Ω) + t(cid:107)∇ψt ∗ ut(cid:107)Lp(Ω), for every 0 < t < RΩ 2 . Let us set then we have that for every x ∈ Ω, Ωt = {x ∈ RN : dist(x, Ω) < t}, y (cid:55)→ ψt(x − y) has support contained in Ωt. HOMOGENEOUS SOBOLEV SPACES 17 (cid:18) x − y (cid:19) t dy dx. By using this and Jensen's inequality, we obtain Thus by using a change of variable and Fubini Theorem we get (cid:90) RΩ/2 0 t Ω Ωt ts ≤ (cid:90) Lp(Ω) ≤ (cid:107)u − ψt ∗ ut(cid:107)p (cid:12)(cid:12)(cid:12)(cid:12)u(x) − u (cid:18) RΩ (cid:90) (cid:19)p dt (cid:18)(cid:107)u − ψt ∗ ut(cid:107)Lp(Ω) (cid:12)(cid:12)(cid:12)(cid:12)u(x) − u (cid:90) (cid:90) (cid:90) RΩ/2 (cid:18) RΩ − t (cid:19)N (cid:90) RΩ/2 (cid:90) (cid:90) (cid:32)(cid:90) RΩ/2 (cid:90) (cid:90) (cid:101)Ω Ωt ⊂(cid:101)Ω := 2 ΩRΩ/2, u(x) − u(z)p t−s p−N ψ RΩ RΩ−t Ωt t−s p RΩ ≤ = Ωt Ω Ω Ω 0 0 0 RΩ RΩ − t (cid:18) R R − t y (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)p 1 (cid:18) x − z tN ψ (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)p 1 (cid:18) x − y (cid:19) (cid:18) x (cid:33) (cid:19) dt t tN ψ y tN ψ RΩ − t t t−s p u(x) − u(z)p 1 dy dx dt t − RΩ − t RΩ t (cid:19) z dz dx dt t where we used that ψ We now observe that (cid:18) x − z (cid:12)(cid:12)(cid:12)(cid:12) x − z (cid:12)(cid:12)(cid:12)(cid:12) ≥ 1 + i.e. for every x ∈ Ω and z ∈(cid:101)Ω, thus in particular z RΩ + if t t (cid:54)= 0 (cid:19) (cid:12)(cid:12)(cid:12)(cid:12) z RΩ (cid:12)(cid:12)(cid:12)(cid:12) if 1 + 0 < t ≤ x − z z RΩ (cid:19) dt (cid:18) x − z This implies that for x ∈ Ω and z ∈(cid:101)Ω we get (cid:90) +∞ (cid:90) RΩ/2 (cid:90) +∞ (cid:90) +∞ t−s p−N ψ z RΩ ≤ x−z z 1+ RΩ + = t t 0 0 + z RΩ t t dz dx, RΩ 2 . (cid:12)(cid:12)(cid:12)(cid:12) < 1, (cid:19) z RΩ + for 0 < t < t + (cid:12)(cid:12)(cid:12)(cid:12) x − z (cid:18) x − z (cid:18) x − z ψ t + z RΩ t z RΩ (cid:19) = 0, = 0. ⇐⇒ then then ψ t−s p−N ψ t−s p−N ψ t + (cid:18) x − z (cid:18) x − z (cid:18) x − z diam((cid:101)Ω) + t t z RΩ z RΩ t (cid:19) dt (cid:19) dt (cid:19) dt (cid:33)N +s p z RΩ + t t ≤ x−z diam((cid:101)Ω) RΩ 1+ t−s p−N ψ (cid:32) ≤ N + 1 ωN (N + s p) 1 + RΩ x − z−N−s p. 18 Thus, we obtain(cid:90) RΩ/2 0 (4.7) BRASCO AND SALORT (cid:18)(cid:107)u − ψt ∗ ut(cid:107)Lp(Ω) (cid:32) (cid:32) ωN (N + s p) N + 1 ≤ ts N + 1 t (cid:19)p dt diam((cid:101)Ω) diam((cid:101)Ω) RΩ 1 + ωN (N + s p) RΩ 1 + ≤ (cid:33)N +s p (cid:90) (cid:33)N +s p (cid:90) (cid:101)Ω (cid:107)u(cid:107)pDs,p Ω Observe that by construction u(x) − u(z)p x − zN +s p dx dz diam((cid:101)Ω) = 2 diam(ΩRΩ/2) ≤ 2 (cid:33)p (cid:90) RΩ/2 0 (Ω) tp (cid:32)(cid:107)ψt ∗ ut(cid:107)D1,p (cid:18) RΩ (cid:90) ts u RN RΩ − t ∇ψt ∗ ut(x) = 0 (Ω). (cid:16) (cid:17) . diam(Ω) + RΩ dt t 0 (Ω). ≤ C (cid:107)u(cid:107)pDs,p (cid:18) x − y (cid:19) dy, t We now need to show that (4.8) 0 We first observe that (cid:90) (cid:90) (cid:18)(cid:90) RN RN (cid:18)(cid:90) RN 1 L1(RN ) RN tN +1 (cid:107)∇ψ(cid:107)p−1 tp−1 (cid:107)∇ψ(cid:107)p−1 tp−1 L1(RN ) ≤ × = ≤ RΩ − t y x − u RΩ − t (cid:12)(cid:12)(cid:12)(cid:12)u (cid:18) RΩ (cid:19) (cid:18) RΩ (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dy (cid:19)p−1 (cid:18) x − y (cid:12)(cid:12)(cid:12)(cid:12)u (cid:19) (cid:18) RΩ (cid:90) (cid:90) (cid:90) (cid:90) RΩ − t RN RN x t u (z) − u (w)p RN RN t (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) x − y (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)p (cid:18) RΩ − t 1 y tN +1 1 (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)p (cid:18) RΩ (cid:12)(cid:12)(cid:12)(cid:12)∇ψ 1 RΩ − t (cid:19) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dy (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) x − y (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dz dw. t tN +1 (z − w) tN +1 RΩ t − u dx (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dy dx y (cid:19) 1 tN +1 ∇ψ (cid:19) (cid:18) x − y (cid:18) RΩ (cid:19)(cid:21) dy = 0. t RΩ − t y − u (cid:18) x − y (cid:19) t dy, 1 tN +1 ∇ψ and by the Divergence Theorem (cid:90) (cid:20) Thus we obtain as well (cid:90) −∇ψt ∗ ut(x) = u RN 1 RN tN +1 ∇ψ (cid:19) (cid:18) RΩ RΩ − t x and by Holder's inequality (cid:107)ψt ∗ ut(cid:107)p = ∇utp dx D1,p 0 (Ω) HOMOGENEOUS SOBOLEV SPACES 19 This yields(cid:90) RΩ/2 tp 0 (4.9) (cid:33)p (cid:90) ts 0 (Ω) t−s p (cid:32)(cid:107)ut(cid:107)D1,p (cid:90) RΩ/2 (cid:90) (cid:90) (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) RΩ − t RΩ t RN RN 0 ≤ C = C dt t (cid:90) RN RN (z − w) u(z) − u(w)p u(z) − u(w)p 1 tN t−s p 1 tN (cid:32)(cid:90) RΩ/2 (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:54)= 0 ⇐⇒ 0 As above, we now observe that thus in particular for 0 < t < RΩ/2 we have =⇒ z − w > 1 This implies that for z, w ∈ RN we have (cid:90) RΩ/2 0 t−s p 1 tN 1 2 (cid:12)(cid:12)(cid:12)(cid:12)∇ψ t (cid:18) RΩ − t RΩ t (z − w) (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) RΩ − t (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) RΩ − t RΩ t RΩ t (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dz dw (cid:33) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dt t (z − w) (z − w) dt t dz dw. ∇ψ (cid:18) RΩ − t (cid:90) +∞ RΩ t RΩ − t RΩ (z − w) z − w < 1, t = 0. (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)∇ψ (cid:18) RΩ − t z−w t−s p 1 tN RΩ t z − w−N−s p. 2 N + 1 ωN (N + s p) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dt t ≤ ≤ (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) dt t (z − w) By inserting this estimate in (4.9), we now get (4.8). We are left with estimating the integral of the K−functional on (RΩ/2, +∞): for this, we can use the trivial decomposition which gives (cid:90) +∞ (cid:32) K(t, u, Lp(Ω),D1,p 0 (Ω)) RΩ 2 ts u = (u − 0) + 0, (cid:33)p (cid:90) +∞ (cid:107)u(cid:107)p Lp(Ω) ts p (cid:18) RΩ 2 RΩ 2 (cid:107)u(cid:107)p Lp(Ω) s p ≤ dt t = ≤ 2s p s p (cid:107)u(cid:107)pDs,p 0 (Ω) dt t (cid:19)−s p (cid:18) 1 p(Ω) Rs p λs Ω (cid:19) , 0 (Ω). By recalling that for a convex set with finite inradius where we used Poincar´e inequality for Ds,p we have (see [8, Corollary 5.1]) Ω ≥ p(Ω) Rs p λs (cid:33)p for a constant C = C(N, p) > 0, we finally obtain (cid:32) (cid:90) +∞ K(t, u, Lp(Ω),D1,p 0 (Ω)) RΩ 2 ts C s (1 − s) , dt t ≤ 2s p p (cid:107)u(cid:107)pDs,p 0 (Ω) (cid:18) 1 − s C (cid:19) . By using this in conjunction with (4.7) and (4.8), we finally conclude the proof. (cid:3) 20 BRASCO AND SALORT For general unbounded convex sets, the previous proof does not work anymore. However, for convex cones the result still holds. We say that a convex set Ω ⊂ RN is a convex cone centered at x0 ∈ RN if for every x ∈ Ω and τ > 0, we have x0 + τ (x − x0) ∈ Ω. Then we have the following Corollary 4.8 (Comparison of norms II: convex cones). Let 1 < p < ∞ and 0 < s < 1. If Ω ⊂ RN is an open convex cone centered at x0 ∈ RN , then for every u ∈ C∞ 0 (Ω) we have (cid:107)u(cid:107)pX s,p (Ω) ≤ C (cid:107)u(cid:107)pDs,p 0 (Ω), 0 for a constant C = C(N, p,E(Ω ∩ B1(x0))) > 0. In particular, we have X s,p Proof. We assume for simplicity that x0 = 0 and take u ∈ C∞ we have that u ∈ C∞ 0 (Ω ∩ BR(0)), for R large enough. From the previous result, we know that (Ω) = Ds,p 0 (Ω). 0 (Ω). Since u has compact support, 0 (cid:107)u(cid:107)pX s,p (Ω∩BR(0)) ≤ C (cid:107)u(cid:107)Ds,p 0 (Ω∩BR(0)) = C (cid:107)u(cid:107)Ds,p 0 (Ω). 0 We recall that the constant C depends on the eccentricity of Ω∩ BR(0). However, since Ω is a cone, we easily get E(Ω ∩ BR(0)) = E(Ω ∩ B1(0)), for every R > 0, i.e. the constant C is independent of R. Finally, by observing that we get the desired conclusion. (cid:107)u(cid:107)pX s,p (Ω) ≤ (cid:107)u(cid:107)pX s,p 0 0 (Ω∩BR(0)), (cid:3) Remark 4.9 (Rotationally symmetric cones). Observe that if Ω is the rotationally symmetric convex cone Ω = {x ∈ RN : (cid:104)x − x0, ω(cid:105) > β x − x0}, for some 0 ≤ β < 1, x0 ∈ RN and ω ∈ SN−1, we have E(Ω ∩ B1(0)) = 1 2 max (cid:110) 2 (cid:112) (cid:111)(cid:32) (cid:33) , 1(cid:112)1 − β2 1 − β2, 1 1 + by elementary geometric considerations. In particular, when Ω is a half-space (i.e. β = 0), then we have E(Ω ∩ B1(0)) = 2. 4.3. Lipschitz sets and beyond. In this section we show that the norms of X s,p are equivalent on open bounded Lipschitz sets. We also make some comments on more general sets, see Remark 4.11 below. and Ds,p 0 0 By generalizing the idea of [22, Theorem 11.6] (see also [6, Theorem 2.1]) for p = 2 and smooth sets, we can rely on the powerful extension theorem for Sobolev functions proved by Stein and obtain the following Theorem 4.10 (Comparison of norms II: Lipschitz sets). Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open bounded set, with Lipschitz boundary. Then for every u ∈ C∞ 0 (Ω) we have (cid:107)u(cid:107)pX s,p (Ω) ≤ C1 (cid:107)u(cid:107)pDs,p 0 (Ω), 0 for a constant C1 > 0 depending on N, p, diam(Ω) and the Lipschitz constant of ∂Ω. In particular, we have X s,p (Ω) = Ds,p 0 (Ω) in this case as well. 0 HOMOGENEOUS SOBOLEV SPACES 21 Proof. We take an open ball B ⊂ RN with radius diam(Ω) and such that Ω (cid:98) B. We then take a linear and continuous extension operator such that (4.10) T : W 1,p(B \ Ω) → W 1,p(B),  (cid:107)T (u)(cid:107)Lp(B) ≤ eΩ (cid:107)u(cid:107)Lp(B), (cid:107)∇T (u)(cid:107)Lp(B) ≤ eΩ (cid:107)u(cid:107)W 1,p(B), where eΩ > 0 depends on N, p, ε, δ and diam(Ω). We observe that such an operator exists, thanks to the fact that Ω has a Lipschitz boundary, see [27, Theorem 5, page 181]. We also observe that the first estimate in (4.10) is not explicitly stated by Stein, but it can be extrapolated by having a closer look at the proof, see [27, page 192]. For every v ∈ C∞ 0 (B), we define the operator R(v) = v − T (v), and observe that R(v) ≡ 0 in B \ Ω and R(v) ∈ W 1,p(B). Since Ω has continuous boundary, this implies that R(v) ∈ D1,p u ∈ C∞ (cid:17) 1 0 (B) and every ε > 0, we take ϕε ∈ C∞ 0 (RN ), for every v ∈ C∞ p (cid:107)ϕε − R(v)(cid:107)Lp(Ω) ≤ (cid:107)∇ϕε − ∇R(v)(cid:107)Lp(Ω) < ε. (cid:16) λ1 p(Ω) 0 (Ω) such that 0 (Ω), see Remark 2.1. We now fix This is possible, thanks to the definition of D1,p K−functional as follows 0 (Ω). Then for t > 0 we can estimate the relevant K(t,R(u), Lp(Ω),D1,p 0 (Ω)) ≤ (cid:107)R(u) − ϕε(cid:107)Lp(Ω) + t(cid:107)∇ϕε(cid:107)Lp(Ω) ≤ (cid:107)R(u) − R(v)(cid:107)Lp(Ω) + (cid:107)R(v) − ϕε(cid:107)Lp(Ω) + t(cid:107)∇R(v)(cid:107)Lp(Ω) + t(cid:107)∇ϕε − ∇R(v)(cid:107)Lp(Ω) ≤ (cid:107)R(u − v)(cid:107)Lp(Ω) + t(cid:107)∇R(v)(cid:107)Lp(Ω) + ε ≤ (cid:107)u − v(cid:107)Lp(Ω) + (cid:107)T (u − v)(cid:107)Lp(Ω) (cid:16) p(cid:17) 1 +(cid:0)λ1 p(Ω)(cid:1)− 1 p(cid:17) 1 +(cid:0)λ1 p(Ω)(cid:1)− 1 0 (Ω)) ≤ (1 + eΩ)(cid:107)u − v(cid:107)Lp(B) + t(cid:0)(cid:107)∇v(cid:107)Lp(B) + eΩ (cid:107)v(cid:107)W 1,p(B) + t(cid:0)(cid:107)∇v(cid:107)Lp(Ω) + (cid:107)∇T (v)(cid:107)Lp(Ω) (cid:1) + ε (cid:16) (cid:1) . By applying (4.10), we then get K(t,R(u), Lp(Ω),D1,p We now use that (cid:107)v(cid:107)W 1,p(B) = (cid:16)(cid:107)v(cid:107)p (cid:16) 1 +(cid:0)λ1 + ε p(cid:17) p(Ω)(cid:1)− 1 (cid:17) 1 . Lp(B) + (cid:107)∇v(cid:107)p Lp(B) p ≤ (cid:107)∇v(cid:107)Lp(B) (cid:18) 1 + (cid:19) 1 p , 1 λ1 p(B) 22 BRASCO AND SALORT thanks to Poincar´e inequality. By spending this information in the previous estimate and using the arbitrariness of ε, we get K(t,R(u), Lp(Ω),D1,p (cid:32) 0 (Ω)) ≤ (1 + eΩ)(cid:107)u − v(cid:107)Lp(B) 1 λ1 p(B) 1 + eΩ 1 + + t (cid:18) p(cid:33) (cid:19) 1 (cid:107)∇v(cid:107)Lp(B). We set for simplicity (cid:18) γΩ = 1 + eΩ 1 + (cid:19) 1 p , 1 λ1 p(B) then by taking the infimum over v ∈ C∞ 0 (B) K(t,R(u), Lp(Ω),D1,p 0 (Ω)) ≤ γΩ K(t, u, Lp(Ω),D1,p 0 (Ω)). As usual, we integrate in t, so to get (cid:107)R(u)(cid:107)pX s,p (4.11) We now observe that if u ∈ C∞ for the convex set B, we get 0 (Ω) ≤ γp Ω (cid:107)u(cid:107)pX s,p 0 (B), for u ∈ C∞ 0 (RN ). 0 (Ω), then we have R(u) = u. Thus from (4.11) and Theorem 4.7 (cid:107)u(cid:107)pX s,p 0 (Ω) ≤ C γp Ω (cid:107)u(cid:107)pDs,p 0 (RN ) = C γp Ω (cid:107)u(cid:107)pDs,p 0 (Ω), where C only depends on N and p. This concludes the proof. for every u ∈ C∞ 0 (Ω), (cid:3) Remark 4.11 (More general sets). It is not difficult to see that the previous proof works (and thus X s,p 0 (Ω) are equivalent), whenever the set Ω is such that there exists a linear and continuous extension operator (Ω) and Ds,p 0 T : W 1,p(B \ Ω) → W 1,p(B), such that (4.10) holds. Observe that there is a vicious subtility here: the first condition in (4.10) is vital and, in general, it may fail to hold for an extension operator. For example, there is a beautiful extension result by Jones [21, Theorem 1], which is valid for very irregular domains (possibly having a fractal boundary): however, the construction given by Jones does not assure that the first estimate in (4.10) holds true, see the statement of [21, Lemma 3.2]. In order to complement the discussion of Remarks 3.4 and 4.5 on "extensions by zero" in X s,p , 0 we explicitly state the consequence of (4.11). Corollary 4.12. Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open bounded set as in Theorem 4.10. Then for every u ∈ C∞ 0 (Ω), there holds (cid:107)u(cid:107)pX s,p 0 (Ω) ≤ C (cid:107)u(cid:107)pX s,p 0 (RN ), for a constant C = C(N, p, ε, δ) > 0. (cid:111) , (cid:111) . HOMOGENEOUS SOBOLEV SPACES 23 5. Capacities Let 1 < p < ∞ and 0 < s < 1 be such that3 s p < N . For every compact set F ⊂ RN , we define the (s, p)−capacity of F inf u∈C∞ 0 (RN ) and the interpolation (s, p)−capacity of F caps,p(F ) = [u]p (cid:110) (cid:110)(cid:107)u(cid:107)pX s,p 0 W s,p(RN ) : u ≥ 0 and u ≥ 1F int caps,p(F ) = inf u∈C∞ 0 (RN ) (RN ) : u ≥ 0 and u ≥ 1F As a straightforward consequence of Propositions 4.1 and 4.6, we have the following Corollary 5.1 (Comparison of capacities). Let 1 < p < ∞ and 0 < s < 1 be such that s p < N . Let F ⊂ RN be a compact set, then we have for a constant C = C(N, p) > 1. In particular, it holds caps,p(F ) ≤ int caps,p(F ) ≤ C caps,p(F ), 1 C caps,p(F ) = 0 if and only if int caps,p(F ) = 0. Proposition 5.2. Let 1 < p < ∞ and 0 < s < 1 be such that s p < N . For every E, F ⊂ RN compact sets, we have caps,p(E ∪ F ) ≤ caps,p(E) + caps,p(F ). Proof. We fix n ∈ N \ {0} and choose two non-negative functions ϕn, ψn ∈ C∞ 0 (RN ) such that [ϕn]p [ψn]p W s,p(RN ) ≤ caps,p(E) + W s,p(RN ) ≤ caps,p(F ) + (cid:16) max{ϕn, ψn}(cid:17) ∗ ε, 1 n Un,ε = 1 n , ϕn ≥ 1E, , ψn ≥ 1F . 0 < ε (cid:28) 1, and We then set we thus get where {ε}ε>0 is a family of standard Friedrichs mollifiers. We observe that for every n ∈ N \ {0}, it holds that Un,ε ∈ C∞ 0 (RN ). Moreover, by construction we have By observing that Jensen's inequality implies Un,ε ≥ 1E∪F . [Un,ε]W s,p(RN ) ≤(cid:104) max{ϕn, ψn}(cid:105)p caps,p(E ∪ F ) ≤ [Un,ε]W s,p(RN ) ≤(cid:104) , W s,p(RN ) max{ϕn, ψn}(cid:105)p W s,p(RN ) . By using the submodularity of the Sobolev-Slobodeckıi seminorm (see [19, Theorem 3.2 & Remark 3.3]), we obtain caps,p(E ∪ F ) ≤ [ϕn]p W s,p(RN ) + [ψn]p W s,p(RN ). 3As usual, the restriction s p < N is due to the scaling properties of the relevant energies. It is not difficult to see that for s p ≥ N , both infima are identically 0. 24 BRASCO AND SALORT Finally, thanks to the choice of ϕn and ψn, we get the desired conclusion by the arbitrariness of (cid:3) n. Proposition 5.3. Let 1 < p < ∞ and 0 < s < 1 be such that s p < N . Let Ω ⊂ RN be an open set. We take a compact set E (cid:98) Ω such that Then we have (5.1) and (5.2) caps,p(E) = 0. Hτ (E) = 0 for every τ > N − s p, p(Ω \ E) = λs λs p(Ω). Proof. To prove (5.1), we can easily adapt the proof of [17, Theorem 4, page 156], dealing with the local case. In order to prove (5.2), we first assume Ω to be bounded. Let ε > 0, we take uε ∈ C∞ 0 (Ω) such that (cid:90) We further observe that the boundedness of Ω implies that (cid:107)uε(cid:107)pDs,p 0 (Ω) < (1 + ε) λs p(Ω) uεp dx = 1. and Ω (cid:110)(cid:107)u(cid:107)pDs,p (cid:111) 0 (Ω) : (cid:107)u(cid:107)Lp(Ω) = 1 , λs p(Ω) = min u∈Ds,p 0 (Ω) and that any solution of this problem has norm L∞(Ω) bounded by a universal constant, see [7, Theorem 3.3]. Thus, without loss of generality, we can also assume that Since E has null (s, p)−capacity, there exists ϕε ∈ C∞ (cid:107)uε(cid:107)L∞(Ω) ≤ M, for 0 < ε (cid:28) 1. 0 (Ω) such that [ϕε]p W s,p(RN ) < ε, ϕε ≥ 0 and ϕε ≥ 1E. We set ψε = ϕε/(cid:107)ϕε(cid:107)L∞(RN ) and observe that (cid:107)ϕε(cid:107)L∞(RN ) ≥ 1. The function uε (1 − ψε) is p(Ω \ E), then by using the triangle inequality we admissible for the variational problem defining λs have (cid:16) (cid:17) 1 p ≤ [uε (1 − ψε)]W s,p(RN ) p(Ω \ E) λs (cid:107)uε (1 − ψε)(cid:107)Lp(Ω\E) (cid:107)uε (1 − ψε)(cid:107)Lp(Ω\E) ≤ [uε]W s,p(RN ) (cid:107)1 − ψε(cid:107)L∞(RN ) (cid:107)uε(cid:107)L∞ [ψε]W s,p(RN ) (cid:107)uε (1 − ψε)(cid:107)Lp(Ω\E) + . From the first part of the proof, we know that E has Lebesgue measure 0, thus the Lp norm over Ω \ E is the same as that over Ω. If we now take the limit as ε goes to 0 and use the properties of uε, together with4 [ψε]W s,p(RN ) < ε (cid:107)ϕε(cid:107)L∞(RN ) ≤ ε, 4Observe that, from the first condition, we get that ψε converges to 0 strongly in Lp(Ω), by Sobolev inequality. Since the family {uε} is bounded in L∞(Ω), this is enough to infer (cid:90) lim ε→0 Ω uεp 1 − ψεp dx = 1. HOMOGENEOUS SOBOLEV SPACES 25 and we get lim ε→0 (cid:107)1 − ψε(cid:107)L∞(RN ) = lim ε→0 (cid:16) (cid:17) 1 p ≤(cid:16) p(Ω \ E) λs λs p(Ω) (1 − ψε) ≤ 1, sup RN (cid:17) 1 0 (Ω \ E) ⊂ C∞ . p 0 (Ω), thus we get the In order to remove the last assumption, we consider the sets ΩR = Ω ∩ BR(0). For R large The reverse inequality simply follows from the fact that C∞ conclusion when Ω is bounded. enough, this is a non-empty open bounded set and E (cid:98) ΩR as well. We thus have p(ΩR \ E) = λs λs p(ΩR). By taking the limit5 as R goes to +∞, we get the desired conclusion in the general case as well. (cid:3) The previous result giving the link between the Poincar´e constant and sets with null capacity does (Ω). Indeed, we have the following result, which shows not hold true in the interpolation space X s,p that the interpolation Poincar´e constant is sensitive to removing sets with null (s, p)−capacity. Lemma 5.4. Let 1 < p < N and 0 < s < 1. Let Ω ⊂ RN be an open set and E (cid:98) Ω a compact set such that 0 int caps,p(E) = 0 < capp(E). Then we have Proof. By Corollary 3.3, we know that p(Ω \ E) = Λs It is now sufficient to use that λ1 p−capacity. p(Ω). p(Ω \ E) > Λs Λs (cid:17)s (cid:16) p(Ω \ E) λ1 p(Ω \ E) > λ1 and (cid:16) λ1 p(Ω) (cid:17)s . Λs p(Ω) = p(Ω), as a consequence of the fact that E has positive (cid:3) Remark 5.5. As an explicit example of the previous situation, we can take s p < 1 and the (N − 1)−dimensional set F = [−a, a]N−1 × {0}. Observe that capp(F ) > 0 by [17, Theorem 4, page 156]. On the other hand, we have int caps,p(F ) = 0. Indeed, we set We then take the usual sequence of Friedrichs mollifiers {ε}ε>0 ⊂ C∞ 0 (RN ) and define Fε = {x ∈ RN : dist(x, F ) < ε}. ϕε = 1Fε ∗ ε ∈ C∞ 0 (RN ). Observe that by construction we have ϕε ≡ 1 on Fε and ϕε ≡ 0 on RN \ F2 ε. 5Such a limit exists by monotonicity. 26 BRASCO AND SALORT By definition of (s, p)−capacity and using the interpolation estimate (4.3), we get (cid:18)(cid:90) caps,p(F ) ≤ [ϕε]p (cid:18)(cid:90) ≤ C Q W s,p(RN ) ϕεp dx (cid:19)1−s (cid:18)(cid:90) (cid:19)1−s (cid:18)(cid:90) Q 1Fεp dx ≤ C ≤ C Fε ε−s p ≤ C ε1−s p. Q (cid:19)s (cid:19)s(cid:18)(cid:90) ∇ϕεp dx 1Fεp dx Q (cid:19)s p ∇ε dx Q We then observe that the last quantity goes to 0 as ε goes to 0, thanks to the fact that s p < 1. By Corollary 5.1, we have int caps,p(F ) = caps,p(F ) = 0. as desired. 6. Double-sided estimates for Poincar´e constants We already observed that for an open set Ω ⊂ RN we have (cid:16) (cid:17)s s (1 − s) Λs p(Ω) = λ1 p(Ω) . We now want to compare λ1 Theorem 6.1. Let 1 < p < ∞ and 0 < s < 1. Let Ω ⊂ RN be an open set, then p with the sharp Poincar´e constant for the embedding Ds,p 0 (Ω) (cid:44)→ Lp(Ω). s (1 − s) λs p(Ω) ≤ 2p (1−s) N ωN λ1 p(Ω) . (cid:17)s If in addition: • Ω ⊂ RN is bounded with Lipschitz boundary, then we also have the reverse inequality and C = C(N, p) > 0 is the same constant as in the Hardy inequality for Ds,p Theorem 1.1]). 0 (Ω) (see [8, Proof. The first inequality (6.1) is a direct consequence of the interpolation inequality (4.3). Indeed, by using the definition of λs p(Ω), we obtain from this inequality Lp(Ω) ≤ C (cid:107)u(cid:107)(1−s) p p(Ω)(cid:107)u(cid:107)p s (1 − s) λs Lp(Ω) (cid:107)∇u(cid:107)s p Lp(Ω), for every u ∈ C∞ over C∞ 0 (Ω), we get the claimed inequality. 0 (Ω). By simplifying the factor (cid:107)u(cid:107)p Lp(Ω) on both sides and taking the infimum (6.1) (6.2) (6.3) (cid:16) (cid:17)s ≤ s (1 − s) λs (cid:17)s ≤ s (1 − s) λs (cid:17)s (cid:16) λ1 p(B1(0)) C , (cid:16) (cid:16) 1 C1 1 C2 C2 = λ1 p(Ω) p(Ω), where C1 > 0 is the same constant as in Theorem 4.10; • Ω ⊂ RN is convex, then we also have the reverse inequality λ1 p(Ω) p(Ω), where C2 is the universal constant given by HOMOGENEOUS SOBOLEV SPACES 27 In order to prove (6.2), for every ε > 0 we take ϕ ∈ C∞ 0 (Ω) such that (cid:107)ϕ(cid:107)pDs,p (cid:107)ϕ(cid:107)p 0 (Ω) Lp(Ω) < λs p(Ω) + ε, then we use Theorem 4.10 to infer 1 C1 This in turn implies (cid:107)ϕ(cid:107)pX s,p (cid:107)ϕ(cid:107)p 0 Lp(Ω) (Ω) < λs p(Ω) + ε. p(Ω) ≤ λs Λs p(Ω), 1 C1 by arbitrariness of ε > 0. A further application of Corollary 3.3 leads to the desired conclusion. Finally, if Ω ⊂ RN is convex, we can proceed in a different way. We first observe that we can always suppose that RΩ < +∞, otherwise both λ1 p(Ω) vanish and there is nothing to prove. Then (6.3) comes by joining the simple estimate p(Ω) and λs p(Ω) ≤ λ1 λ1 p(B1(0)) Rp Ω , which follows from the monotonicity and scaling properties of λ1 5.1], i.e. p, and the estimate of [8, Corollary s (1 − s) λs p(Ω) ≥ C . Rs p Ω The latter is a consequence of the Hardy inequality in convex sets for Ds,p 0 . (cid:3) Remark 6.2. For p = 2, the double-sided estimate of Theorem 6.1 is contained in [12, Theorem 4.5]. The proof in [12] relies on probabilistic techniques. In [12] the result is proved by assuming that Ω verifies a uniform exterior cone condition. Remark 6.3. Inequality (6.2) can not hold for a general open set Ω ⊂ RN , with a constant independent of Ω. Indeed, one can construct a sequence {Ωn}n∈N ⊂ RN such that lim n→∞ see Lemma A.1 below. λ1 p(Ωn) λs p(Ωn) = +∞, for 1 < p < ∞ and s < 1 p , (cid:16) (cid:16) (cid:17)s (cid:17)s In this section, we construct a sequence of open bounded sets {Ωn}n∈N ⊂ RN with rough boundaries and fixed diameter, such that we have Appendix A. An example lim n→∞ (A.1) The sets Ωn are obtained by removing from an N−dimensional cube an increasing array of regular (N − 1)−dimensional cracks. . = +∞, for 1 < p < ∞ and s < 1 p λ1 p(Ωn) λs p(Ωn) 28 BRASCO AND SALORT Figure 1. The set Ωn in dimension N = 2, for n = 3. (cid:21)N , 1 2 (cid:20) − 1 4 , 1 4 and F = For N ≥ 1, we set6 − 1 2 For every n ∈ N, we also define Q = (cid:20) (cid:110) Finally, we consider the sets Ω = Q \ F, (cid:101)Ωn = and (cid:21)N−1 × {0}. (cid:111) (cid:21)N \ (cid:91) z∈ZN n ZN n = z = (z1, . . . , zN ) ∈ ZN : max{z1, . . . ,zN} ≤ n . (cid:17) (cid:20) (cid:16) (cid:91) (cid:101)Ωn = RN \ (cid:91) (cid:91) Ω + z = n z∈ZN E = n∈N (F + z). z∈ZN −n − 1 2 , n + 1 2 (F + z), Then (A.1) is a consequence of the next result. Lemma A.1. With the notation above, for 1 < p < ∞ and s < 1/p we have for every n ∈ N, (A.2) p((cid:101)Ωn) ≥ C = C(N, p, F ) > 0, λ1 and (A.3) In particular, the new sequence of rescaled sets {Ωn}n∈N ⊂ RN defined by n→∞ λs lim p((cid:101)Ωn) = 0. (cid:21)N \ (cid:91) , − 1 2 1 2 (cid:20) z∈ZN n (F + z) 2 n + 1 , Ωn = (cid:101)Ωn− 1 N (cid:101)Ωn = 6For N = 1, the set F simply coincides with the point {0}. HOMOGENEOUS SOBOLEV SPACES is such that diam(Ωn) = √ N , for every n ∈ N and lim n→∞ 29 (cid:0)λ1 p(Ωn)(cid:1)s λs p(Ωn) = +∞. Proof. We divide the proof in two parts, for ease of readability. Of course, it is enough to prove (A.2) and (A.3). Indeed, the last statement is a straightforward consequence of these facts and of the scaling properties of the diameter and of the Poincar´e constants. Proof of (A.2). For 1 < p < ∞ we define ∇up dx up dx : u = 0 on F  . (cid:90) (cid:90) Q  Q (cid:27) µp(Q; F ) = min u∈W 1,p(Q)\{0} We first observe that F is a compact set with positive (N − 1)−dimensional Hausdorff measure, thus by [17, Theorem 4, page 156] we have capp(F ; Q) = inf u∈C∞ 0 (Q) Q ∇up dx : u ≥ 1F > 0, for every 1 < p < ∞. We can thus infer existence of a constant C = C(N, p, F ) > 0 such that up dx ≤ ∇up dx, for every u ∈ W 1,p(Q) such that u = 0 on F, (cid:26)(cid:90) (cid:90) (cid:90) 1 C Q Q see [23, Theorem 10.1.2]. This shows that µp(Q; F ) > 0. For every ε > 0, we consider uε ∈ C∞ 0 (E) \ {0} such that ∇uεp dx uεp dx E λ1 p(E) + ε > . (cid:90) (cid:90) E (cid:90) (cid:90) ∇uεp dx ≥ µ(Q, F ) uεp dx, Q+z Q+z We now observe that for every z ∈ ZN , there holds thanks to the fact that un vanishes on (the relevant translated copy of) F and to the fact that µp(Q, F ) = µp(Q + z, F + z). By using this information, we get (cid:90) E ∇uεp dx = (cid:90) (cid:88) z∈ZN Q+z ≥ µp(Q, F ) ∇uεp dx (cid:90) (cid:88) z∈ZN Q+z uεp dx = µp(Q, F ) (cid:90) E uεp dx. By recalling the choice of uε, we then get p(E) + ε ≥ µp(Q; F ). λ1 Thanks to the arbitrariness of ε > 0 and to the fact that (cid:101)Ωn ⊂ E, this finally gives p(E) ≥ µp(Q; F ), for every n ∈ N, p((cid:101)Ωn) ≥ λ1 λ1 (A.4) as desired. 30 BRASCO AND SALORT Proof of (A.3). We recall that (cid:101)Ωn = (cid:91) z∈ZN n (cid:16) (cid:17) = Ω + z (cid:20) −n − 1 2 , n + (F + z), z∈ZN n 1 2 (cid:21)N \ (cid:91)  = 0.  (cid:91) z∈ZN n caps,p (F + z) and that each (N − 1)−dimensional set F + z has null (s, p)−capacity, thanks to Remark 5.5. By using Proposition 5.2, we also obtain Then by Proposition 5.3, we get p((cid:101)Ωn) = λs p λs (cid:32)(cid:20) −n − 1 2 , n + 1 2 (cid:21)N(cid:33) = (2 n + 1)−s p λs p(Q). This is turn gives the desired conclusion (A.3). (cid:3) Appendix B. One-dimensional Hardy inequality We used the following general form of the one-dimensional Hardy inequality (the classical case corresponds to α = p−1 below). This can be found for example in [23]. For the sake of completeness, we give a sketch of a proof based on Picone's inequality.7 Lemma B.1. Let 1 < p < ∞ and α > 0. For every f ∈ C∞ Proof. We take 0 < β < α/(p − 1) and consider the function ϕ(t) = tβ. Observe that this solves (B.1) 0 p (cid:18) α (cid:19)p (cid:90) T −(cid:0)ϕ(cid:48)(t)p−2 ϕ(cid:48)(t) tp−α−1(cid:1)(cid:48) (cid:90) T βp−1 (α − β (p − 1)) (cid:90) T 0 ((0, T ]) we have f(cid:48)(t)p tp dt t . 0 tα f (t)p tα ≤ dt t = βp−1 (α − β (p − 1)) tβ (p−1)−α−1 = βp−1 (α − β (p − 1)) t−α−1 ϕ(t)p−1. (cid:90) T ϕp−1 tα ψ dt t = 0 ϕ(cid:48)p−2 ϕ(cid:48) 0 tα ψ(cid:48) tp dt t . Thus, for every ψ ∈ C∞ 0 ((0, T ]) we have the weak formulation We take ε > 0 and f ∈ C∞ 0 ((0, T ]) non-negative, we insert the test function 7For u, v differentiable functions with v ≥ 0 and u > 0, we have the pointwise inequality ψ = f p (ε + ϕ)p−1 , (cid:19)(cid:48) u(cid:48)p−2 u(cid:48) (cid:18) vp up−1 ≤ v(cid:48)p. = ≤ 0 (cid:90) T (cid:90) T (cid:90) T 0 ≤ 0 (cid:90) T 0 f p tα dt t tα f(cid:48)(t)p tα tp dt t . f(cid:48)(t)p tα tp dt t . HOMOGENEOUS SOBOLEV SPACES 31 in the previous integral identity. By using Picone's inequality, we then obtain f p ϕ(cid:48)p−2 ϕ(cid:48) ϕp−1 βp−1 (α − β (p − 1)) (ε + ϕ)p−1 f p tα dt t (ε + ϕ)p−1 (cid:18) (cid:19)(cid:48) tp dt t (cid:90) T 0 If we take the limit as ε goes to 0, by Fatou's Lemma we get βp−1 (α − β (p − 1)) The previous inequality holds true for every 0 < β < α/(p− 1) and βp−1 (α − β (p− 1)) is maximal (cid:3) for β = α/p. This concludes the proof. Appendix C. A geometric lemma When comparing the norms of X s,p (Ω) and Ds,p 0 (Ω) for a convex set, we used the following 0 geometric result. We recall that RΩ = sup x∈Ω dist(x, ∂Ω), is the inradius of Ω, i.e. the radius of the largest ball inscribed in Ω. Lemma C.1. Let Ω ⊂ RN be an open convex set such that RΩ < +∞. Let x0 ∈ Ω be a point such that Then for every 0 < t < 1 we have dist(x0, ∂Ω) = RΩ. dist(cid:0)x0 + t (Ω − x0), ∂Ω(cid:1) ≥ (1 − t) RΩ. dist(cid:0)∂(t Ω), ∂Ω(cid:1) ≥ (1 − t) RΩ. Proof. Without loss of generality, we can assume that 0 ∈ Ω and that x0 = 0. Clearly, it is sufficient to prove that Every point of ∂(t Ω) is of the form t z, with z ∈ ∂Ω. We now take the cone Cz, obtained as the convex envelope of BRΩ (0) and the point z. By convexity of Ω, we have of course Cz ⊂ Ω. We thus obtain dist(t z, ∂Ω) ≥ dist(t z, ∂Cz). (C.1) We now distinguish two cases: (i) z = RΩ; (ii) z > RΩ. When alternative i) occurs, then Cz = BRΩ (0) and thus dist(t z, ∂Cz) = dist(t z, BRΩ (0)) = t z − z = (1 − t)z = (1 − t) RΩ. By using this in (C.1), we get the desired estimate. If on the contrary we are in case ii), then by elementary geometric considerations we have t z − z = see Figure 2. This gives again the desired conclusion. dist(t z, ∂Cz) RΩz , (cid:3) 32 BRASCO AND SALORT Figure 2. The case (ii) in the proof of Lemma C.1. Colored in red, the distance of t z from ∂Cz. References [1] R. A. Adams, Sobolev spaces. Pure and Applied Mathematics, Vol. 65. Academic Press, New York-London, 1975. 2 [2] R. A. Adams, J. J. F. Fournier, Sobolev spaces. Second edition. Pure and Applied Mathematics (Amsterdam), 140. Elsevier/Academic Press, Amsterdam, 2003. 2 [3] C. Bennett, R. Sharpley, Interpolation of operators. Pure and Applied Mathematics, 129. Academic Press, Inc., Boston, MA, (1988). 2 [4] J. Bergh, J. Lofstrom, Interpolation spaces. An introduction. Grundlehren der Mathematischen Wissenschaften, 223. Springer-Verlag, Berlin-New York, 1976. 2, 8 [5] J. Bourgain, H. Brezis, P. Mironescu, Another look at Sobolev spaces. Optimal control and partial differential equations, 439–455, IOS, Amsterdam, 2001. 3 [6] J. H. Bramble, Interpolation between Sobolev spaces in Lipschitz domains with an application to multigrid theory, Math. Comp., 64 (1995), 1359–1365. 20 [7] L. Brasco, E. Lindgren, E. Parini, The fractional Cheeger problem, Interfaces Free Bound., 16 (2014), 419–458. 7, 8, 12, 24 [8] L. Brasco, E. Cinti, On fractional Hardy inequalities in convex sets, to appear on Discrete Contin. Dyn. Syst. [9] L. Brasco, F. Santambrogio, A sharp estimate `a la Calder´on-Zygmund for the p−Laplacian, Commun. Contemp. Ser. A (2018), available at http://cvgmt.sns.it/paper/3560/ 7, 19, 26, 27 Math., 20 (2018), 1750030, 24 pp. 16 [10] C. Bucur, E. Valdinoci, Nonlocal Diffusion and Applications, Lecture Notes of the Unione Matematica Italiana, 20. Springer, [Cham]; Unione Matematica Italiana, Bologna, 2016. 2 [11] S. N. Chandler-Wilde, D. P. Hewett, A. Moiola, Interpolation of Hilbert and Sobolev spaces: quantitative estimates and counterexamples, Mathematika, 61 (2015), 414–443. 5 [12] Z.-Q. Chen, R. Song, Two-sided eigenvalue estimates for subordinate processes in domains, J. Funct. Anal., 226 (2005), 90–113. 27 [13] J. Deny, J.-L. Lions, Les espaces du type de Beppo Levi, Ann. Inst. Fourier, 5 (1954), 305–370. 3 [14] E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. 2 [15] B. Dyda, A fractional order Hardy inequality, Illinois J. Math., 48 (2004), 575–588. 7 HOMOGENEOUS SOBOLEV SPACES 33 [16] B. Dyda, A. V. Vahakangas, Characterizations for fractional Hardy inequality, Adv. Calc. Var., 8 (2015), 173–182. 8 [17] L. C. Evans, R. Gariepy, Measure theory and fine properties of functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1992. 24, 25, 29 [18] G. Franzina, Non-local torsion functions and embeddings, to appear on Appl. Anal. (2018), available at http://cvgmt.sns.it/paper/3748/, doi:10.1080/00036811.2018.1463521 7 [19] N. Gigli, S. Mosconi, The abstract Lewy-Stampacchia inequality and applications, J. Math. Pures Appl., 104 (2015), 258–275. 23 [20] P. Grisvard, Elliptic problems in nonsmooth domains. Monographs and Studies in Mathematics, 24. Pitman (Advanced Publishing Program), Boston, MA, 1985. 6, 7 [21] P. W. Jones, Quasiconformal mappings and extendability of functions in Sobolev spaces, Acta Math., 147 (1981), 71–88. 22 [22] J.-L. Lions, E. Magenes, Non-homogeneous boundary value problems and applications. Vol. I. Translated from the French by P. Kenneth. Die Grundlehren der mathematischen Wissenschaften, Band 181. Springer-Verlag, New York-Heidelberg, 1972. 20 [23] V. G. Maz'ja, Sobolev spaces. Translated from the Russian by T. O. Shaposhnikova. Springer Series in Soviet Mathematics. Springer-Verlag, Berlin, 1985. 29, 30 [24] V. Maz'ya, T. Shaposhnikova, On the Bourgain, Brezis, and Mironescu theorem concerning limiting embeddings of fractional Sobolev spaces, J. Funct. Anal., 195 (2002), 230–238. 3 [25] S. M. Nikol'skiı, Approximation of functions of several variables and imbedding theorems. Translated from the Russian by John M. Danskin, Jr. Die Grundlehren der Mathematischen Wissenschaften, Band 205. Springer- Verlag, New York-Heidelberg. 1975. 2 [26] A. Ponce, A new approach to Sobolev spaces and connections to Γ−convergence, Calc. Var. Partial Differential Equations, 19 (2004), 229–255. 3 [27] E. Stein, Singular integrals and differentiability properties of functions. Princeton Mathematical Series, 30. Princeton University Press, Princeton, N.J. 1970. 21 [28] L. Tartar, An introduction to Sobolev spaces and interpolation spaces. Lecture Notes of the Unione Matematica Italiana, 3. Springer, Berlin; UMI, Bologna, 2007. 14 [29] H. Triebel, Interpolation theory, function spaces, differential operators, North-Holland Publishing Co., Amsterdam-New York, 1978 2 [30] H. Triebel, Theory of function spaces. III. Monographs in Mathematics, 100. Birkhauser Verlag, Basel, 2006. 2 [31] H. Triebel, Theory of function spaces. II. Monographs in Mathematics, 84. Birkhauser Verlag, Basel, 1992. 2 [32] H. Triebel, Theory of function spaces. Monographs in Mathematics, 78. Birkhauser Verlag, Basel, 1983. 2, 8 (L. Brasco) Dipartimento di Matematica e Informatica Universit`a degli Studi di Ferrara Via Machiavelli 35, 44121 Ferrara, Italy E-mail address: [email protected] (A. Salort) Departamento de Matem´atica, FCEN Universidad de Buenos Aires and IMAS CONICET, Buenos Aires, Argentina E-mail address: [email protected]
1102.3268
1
1102
2011-02-16T09:07:52
Exact observability, square functions and spectral theory
[ "math.FA", "eess.SY", "math.OC" ]
In the first part of this article we introduce the notion of a backward-forward conditioning (BFC) system that generalises the notion of zero-class admissibiliy introduced in [Xu,Liu,Yung]. We can show that unless the spectum contains a halfplane, the BFC property occurs only in siutations where the underlying semigroup extends to a group. In a second part we present a sufficient condition for exact observability in Banach spaces that is designed for infinite-dimensional output spaces and general strongly continuous semigroups. To obtain this we make use of certain weighted square function estimates. Specialising to the Hilbert space situation we obtain a result for contraction semigroups without an analyticity condition on the semigroup.
math.FA
math
EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY BERNHARD H. HAAK AND EL MAATI OUHABAZ Abstract. In the first part of this article we introduce the notion of a backward- forward conditioning (BFC) system that generalises the notion of zero-class admissibiliy introduced in [21]. We can show that unless the spectum contains a halfplane, the BFC property occurs only in siutations where the underly- ing semigroup extends to a group. In a second part we present a sufficient condition for exact observability in Banach spaces that is designed for infinite- dimensional output spaces and general strongly continuous semigroups. To obtain this we make use of certain weighted square function estimates. Spe- cialising to the Hilbert space situation we obtain a result for contraction semi- groups without an analyticity condition on the semigroup. In this article we study exact observability of linear systems (A, C) on Banach 1. Introduction spaces of the form x′(t) + Ax(t) = 0 = x0 x(0) y(0) = Cx(t)   We suppose throughout this article that −A is the generator of a strongly continuous semigroup T (t)t≥0 on a Banach space X. For details on semigroup theory used frequently in this article we refer to e.g. to the textbooks [5, 7, 16]. Since we deal with unbounded operators in general, we will note D(A) the domain of A and R(A) its range. Let Y be another Banach space and suppose that the observation operator C : D(A) → Y is bounded and linear when D(A) is endowed with the graph norm kxkD(A) = kxk + kAxk. Here we denote by k k the norm of X. Since the observation operator C is generally unbounded, the concept of admissibility is introduced. It means that the output y of the system (usually measured in L2 norm) depends continuously on the initial value x0. Definition 1.1. We say that C is L2-admissible in time τ > 0 (for A or for T (t)t≥0) if there exists a constant M (τ ) > 0 such that x∈D(A),kxk=1Z τ sup 0 kCT (t)xk2 Y dt =: M (τ )2 < ∞. Date: July 20, 2018. 1991 Mathematics Subject Classification. 93B07, 43A45, 47A60. Key words and phrases. Exact observability, admissibility, spectral theory of semigroups and groups, H∞ functional calculus, square function estimates. The first author was partially supported by the ANR project ANR-09-BLAN-0058-01. 1 2 BERNHARD H. HAAK AND EL MAATI OUHABAZ Definition 1.2. We say that C is exactly L2-observable for A (or for T (t)) in time η > 0 if there exists a constant m(η) > 0 such that x∈D(A),kxk=1Z η inf 0 kCT (t)xk2 Y dt =: m(η)2 > 0. For more information the notion of admissible observation (or control) operators we refer the reader to the overview article [8] or, both for admissibility and observ- ability issues to the recent book [20] and references therein. We summarise some well-known facts and notations: When there is no risk of confusion, 'admissible' means L2 admissible in some finite time τ > 0 and 'exact observable' means ex- actly L2-observable for A in some finite time η > 0. We say that C is infinite-time admissible if M (∞) < ∞ and exactly observable in infinite time if m(∞) > 0. Finite-time admissibility does not depend on the choice of τ > 0. Nevertheless it turns out to be useful to study the (clearly non-decreasing) functions t 7→ m(t) and t 7→ M (t). In the 'dual' situation of a control operator B the quantity m(η)−1 is often referred to as control cost of a system. We refer e.g. to [13, 17, 18] and references therein for more details. Independence of the time τ > 0 of the notion of admissibility means that a lack of admissibility expresses either by M (τ ) = ∞ for all τ > 0 or by M (τ ) < ∞ for finite τ while M (∞) = ∞. On the other hand, a lack of exact observability expresses by m(η) = 0 for 0 < η < η0 for η0 ∈ (0, ∞]. We remark that Example 2.3 below satisfies m(η) = 0 for 0 < η < 2, m(2) = 1 while m(η) → +∞ for η → +∞. Since most parabolic equations like for example the heat equation are not exactly observable unless very special observations are chosen whereas exact observability appears frequently for hyperbolic systems such as the wave equation, it appears natural to study necessary spectral conditions of the generator −A that make ex- act observability possible or impossible. In this direction we extend and complete former results of [21]. We introduce the notion of backward-forward condition- ing BFC-systems. These are admissible and exactly observable systems for which M (η) < m(τ ) for some η < τ . We analyse spectral properties of the generator −A of the semigroup of such systems. In particular, we prove that the approximate point spectrum of A is contained in a vertical strip. Therefore, the boundary of the spectrum is also contained in a strip. We prove in addition that if (A, C) is an admissible BFC-system such that the spectrum of A does not contain a half-plane then the semigroup actually extends to a group. Note that every bounded group with an admissible operator C is a BFC-system. Since (BFC) is a frequent property that typically is more likely to hold the more 'regular' the operators A and C are, this shows that exact observability is considerably rare outside the group context. A second part of this paper is devoted to a new sufficient criterion for exact ob- servability. Under an assumption of square function type estimate we prove that a condition like kCA−αxkY ≥ δkxk, implies exact observability. Here α ∈ (0, 1) and δ is a positive constant. Without any further assumption, we show that if −A is the generator of a contrac- tion semigroup on a Hilbert space and C : D(A) → Y is such that kCA−1/2xkY ≥ δkxk, then (A, C) is exactly observable. In order to state and prove our criterion EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 3 we make a heavy use of square function estimate of type kxk2 ≤ K 2Z ∞ 0 k(tA)−β(T (2t2β) − T (t2β))xk2 dt t , where K is a positive constant, β ∈ (0, 1) and T (t) denotes the semigroup generates In the case where β = 1/2, this corresponds to a lower square function by −A. estimate kxk2 ≤ K 2Z ∞ 0 kϕ(tA)xk2 dt t , where ϕ(z) := z−1/2(e−2z − e−z). On Hilbert spaces, it is well known that such estimate is related to the holomorphic functional calculus of the operator A. The needed results on this functional calculus and associated square function estimates for sectorial operators will be sketched in the last two sections. As we will explain later our criterion applies for bounded analytic semigroups on Hilbert spaces whose generator admits a bounded H ∞-calculus -- but the first part of this paper reveals this to be impossible for a large class of systems unless A is bounded. One important aspect of the criterion that might also help in other situations is therefore how to avoid making use of analyticity assumption of the semigroup. We discuss at the end of this papers two examples. 2. BFC-systems Let X and Y be Banach spaces with norms k·k and k·kY , respectively. Through- out this section, (T (t))t≥0 is a strongly continuous semigroup on X whose generator is denoted by −A. Definition 2.1. An admissible observation operator C for A is called zero-class admissible, if limτ →0+ M (τ ) = 0. Note that if the semigroup T (t) is bounded analytic with generator −A then for all α ∈ [0, 1/2), Aα is zero-class admissible. This follows from the inequality kAαT (t)k ≤ M t−α. Consequently, if C is bounded on such a fractional domain space D(Aα) and thus kCT (t)xkY ≤ M [kAαT (t)xk + kT (t)xk] , C is zero-class admissible. It is also a obvious fact that every bounded operator C : X → Y is zero-class ad- missible. Here (T (t))t≥0 is merely a strongly continuous semigroup on X. Consider now the linear operator eΨτ : X → L2(0, τ ; Y ) defined by eΨτ x = CT (·)x. Then admissibility (i.e., M (τ ) < ∞) means that eΨτ is a bounded operator. If in addition m(τ ) > 0, then eΨτ is injective and has closed range. Therefore, we may consider the operator Ψτ : X → R(eΨτ ), Ψτ = eΨτ . We have kxk kxk 1 (2.1) = sup m(τ ) x∈D(A),kxk=1 kΨτ xk = sup x∈D(A),kxk6=0 kΨτ xk = kΨ−1 τ k. We introduce the following definition. Definition 2.2. We say that the system (A, C) has the backward-forward con- ditioning property or shortly that (A, C) is a BFC-system if there exists some 0 < η < τ such that C is admissible and exactly observable in time τ and if (BFC) kΨ−1 τ k kΨηk < 1. 4 BERNHARD H. HAAK AND EL MAATI OUHABAZ The condition (BFC) is clearly a conditioning property for the output operator with different times η and τ which correspond to a backward and forward evolution of the system. It also follows from (2.1) that (BFC) is equivalent to (2.2) M (η) < m(τ ) for some η < τ. Therefore, if C is exactly observable in some time τ and of zero-class, then (2.2) holds trivially by letting η sufficiently small. Hence, the system is BFC. If C is admissible at any τ > 0 and if m(t) → +∞ for t → +∞, then (2.2) holds and again the system is BFC. Zero-class admissible operators are introduced and studied in [21]. See also [9] from which we borrow a concrete example leading to an BFC-system in which C is not zero-class. Example 2.3. The following example is taken from Jacob, Partington and Pott [9, Example 3.9]. We shall use Ingham inequalities to prove that our system is BFC. Similar ideas could be used in a more general class of examples. Consider an undamped wave equation on [0, 1] with Dirichlet boundary conditions and Neumann type observation of the form ∂ 2 ∂t2 z(x, t) = ∂ 2 z(0, t) = z(1, t) = 0 z(x, 0) = z0(x) y(t) = ∂ ∂x z(0, t) for x ∈ (0, 1), t ≥ 0 for t ≥ 0 for x ∈ (0, 1) ∂ ∂t z(x, 0) = z1(x) ∂x2 z(x, t) and We rewrite the system as a first order Cauchy problem   t ≥ 0 ∂ ∂t U = −AU (t), U (0) = (z0, z1) CU (x, t) = ∂   0 (cid:19) and U = (f, g). The latter Cauchy problem is consid- ∂x f (0) −I − ∂ 2 ∂x2 where A =(cid:18) 0 qR 1 0 f ′2dx +R 1 a norm on H 1 ered on the Hilbert space H = H 1 0 g2dx. Note that by the Poincar´e inequality, qR 1 0 (0, 1) × L2(0, 1) endowed with the norm k(f, g)k = 0 f ′2dx defines 0 (0, 1) which is equivalent to the usual one. It is a standard fact that −A generates a strongly continuous semigroup T (t) on H. It is easy to see that A has compact resolvent and the eigenvalues are λn = −inπ, n ∈ Z \ {0} with normalised eigenfunctions Un(x) = (cid:16) sin(nπx) which form an orthonormal basis of H. Fix (f, g) ∈ H and denote by αn = h(f, g), UniH (the scalar product in H). Then , sin(nπx)(cid:17) inπ k(f, g)k2 = Xn∈Z,n6=0 αn2 and Using the well known Ingham inequalities (see e.g. 4.3]) we obtain the following estimates for all τ > 2, CT (t)(f, g) =Xn αneinπtCUn = −iXn αneinπt(cid:12)(cid:12)(cid:12)(cid:12) 0 (cid:12)(cid:12)(cid:12)(cid:12)Xn αn2 ≤Z τ 2 m(τ )2Xn dt ≤ M (τ )2Xn αn2 αneinπt. [22, p. 162] or [10, Theorem EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 5 with m(τ )2 ≥ 2τ π (cid:18)1 − 4 τ 2(cid:19) , M (τ )2 ≤ 8τ π (cid:18)1 + 4 τ 2(cid:19) . This shows that C is admissible at any time τ > 0 and exactly observable in time τ > 2 with constant m(τ ) → +∞ as τ → +∞. This shows (2.2) and hence the system (A, C) is backward-forward conditioning. In order to see that C is not zero-class, we consider small τ > 0 and f ∈ H 1 with Fourier coefficients αn and note that 2 0 (0, 1) where χ[0,τ ] denote the indicator function of [0, τ ]. From this equality it is clear that dt = kχ[0,τ ]f k2 2, 0 (cid:12)(cid:12)(cid:12)(cid:12)Xn Z τ αneinπt(cid:12)(cid:12)(cid:12)(cid:12) 0 (cid:12)(cid:12)(cid:12)(cid:12)Xn kf k2=1Z τ 2 αneinπt(cid:12)(cid:12)(cid:12)(cid:12) dt = M (τ )2. 1 = sup kf k2=1 kχ[0,τ ]f k2 2 = sup Therefore, the right hand side does not converge to 0 as τ → 0. Remark 2.4. The above example is a special case of the following situation: let C be admissible in some arbitrary time τ > 0 and exactly observable in some time η > 0 for a group U (t)t∈R. Observe that kxk = kU (−t)U (t)xk ≤ kU (−t)kkU (t)xk whence kU (t)xk ≥ kU (−t)k−1kxk. From Z nη 0 kCU (t)xk2 dt = kCU (t)U (jη)xk2 dt 0 Z η n−1Xj=0 ≥ m(η)2(cid:18)n−1Xj=0 kU (−jη)k−2(cid:19)kxk2, we then infer m(nη) → +∞ for n → +∞ whenever the sum in the last expres- sion diverges. This is in particular the case for bounded groups U (t)t∈R. By the admissibility of the system (A, C) one then obtains (BFC) by letting n sufficiently large. We thank Hans Zwart for pointing out this remark to us. 3. Spectral properties of BFC-systems We consider the same notation X, Y , A, (T (t))t≥0 and C : D(A) → Y as in the previous section. Or aim here is to study spectral properties of BFC-systems. We will extend some results which have been proved in [21] in the context of zero-class operators. We note also that related ideas and results were obtained previously by Nikolski [14] in the particular case of bounded observation operators C on X. Let us introduce the classical function ε : R+ → R+ defined by ε(t) := inf kxk=1 kT (t)xk. It is clear that ε(t) is strictly positive for all t > 0 if this holds for a single t0 > 0. Indeed, from kT (s)k kT (t)xk ≥ kT (t + s)xk ≥ ε(t)kT (s)xk ≥ ε(t)ε(s)kxk one infers that ε(t)kT (s)k ≥ ε(t + s) ≥ ε(t)ε(s), 6 BERNHARD H. HAAK AND EL MAATI OUHABAZ for all t, s ≥ 0. For this reason we distinguish the cases that ε(t) is strictly positive for all t > 0 of that it vanishes for all t > 0 and we note this by ε(t) > 0 or ε(t) = 0 respectively. The following lemma is essentially contained in [14] and [21]. Lemma 3.1. If (A, C) is an admissible and exactly observable BFC-system, then ε(t) > 0. Proof. By the definition of BFC-system, there exist 0 < η < τ such that By the semigroup property, δ := m(τ )2 − M (η)2 > 0. m(τ )2kxk2 ≤Z τ =Z η 0 0 kCT (t)xk2 Y dt kCT (t)xk2 Y dt +Z τ −η 0 kCT (t)T (η)xk2 Y dt ≤ M (η)2kxk2 + M (τ −η)2kT (η)xk2 which immediately yields kT (η)xk2 ≥ M (τ −η)−2(m(τ )2 − M (η)2)kxk2 ≥ M (τ −η)−2δkxk2. Therefore, ε(η) > 0, and hence ε(t) > 0 for all t > 0. (cid:3) Lemma 3.2. Suppose that (A, C) is an admissible and exactly observable BFC- system. Then T (t)∗ is injective for one (and thus all) t > 0 if and only if T (t) extends to a group on X. Proof. We know by Lemma 3.1 that ε(t) > 0. This implies that T (t) is injective and has closed image for all t ≥ 0. Thus, T (t) is bijective if and only if T (t)∗ is injective. The latter is clearly independent of t > 0 by the semigroup law. Indeed, if T (t0)∗ is injective for some t0 > 0, so are all T (s)∗ for s < t0 since T (t0)∗ = T (t0−s)∗T (s)∗. If s > t0 then we find n ∈ N, δ ∈ [0, t0[ such that T (s)∗ = (T (t0)∗)nT (δ)∗, and the injectivity of T (s)∗ follows from that of T (t0)∗ and T (δ)∗. We saw that T (t)∗ is injective for one (and thus all) t > 0 if and only T (t) is bijective which in turn by S(t) :=(cid:26) T (t) T (t)−1 if if t ≥ 0 t < 0 is equivalent to a group extension of T (t) on X. (cid:3) For a closed operator S on X recall the notions of point spectrum σP (S) =(cid:8)λ ∈ C : ker(λI − S) 6= {0}(cid:9), the approximate point spectrum and the residual spectrum inf x∈D(S),kxk=1 kλx − Sxk = 0(cid:9) σA(S) =(cid:8)λ ∈ C : σR(S) =(cid:8)λ ∈ C : range(λ − S) is not dense in X(cid:9). It is easy to see that σR(S) = σ(S)\σA(S). Of course, σP (S) ⊆ σA(S). EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 7 Proposition 3.3. Let (A, C) be an admissible and exactly observable BFC-system. Then there exist no approximate point spectrum of A with arbitrary large real parts. In particular, if C is an admissible zero-class operator and A has a sequence of approximate point spectrum with arbitrary large real parts then C is not exactly observable. Proof. Recall that exp(−tσA(A)) ⊆ σA(T (t)) (see [5, p. 276], note that −A is the generator). If we find a sequence λn ∈ σA(A) with Re(λn) → +∞, then e−tλn ∈ σA(T (t)). Hence, inf kxk=1 kT (t)x − e−tλn xk = 0. By ε(t)kxk ≤ kT (t)xk ≤ kT (t)x − e−tλn xk + e−tRe(λn)kxk we get ε(t) = 0 which is incompatible with exact observability by Lemma 3.1. (cid:3) Since −A is the generator of a strongly continuous semigroup, σ(A) is contained in a right-half plane. On the other hand, it is well known that the boundary of the spectrum ∂σ(A) is contained in σA(A). We obtain from the previous proposition that Re(∂σ(A)) is bounded, i.e., ∂σ(A) is contained in a vertical strip. Thus Corollary 3.4. Let (A, C) be an admissible BFC-system. Then Re(∂σ(A)) := {Re(λ) : λ ∈ ∂σ(A)} is bounded. The following lemma is known. Lemma 3.5. Let S ∈ B(X) satisfy kSxk ≥ γkxk for some γ > 0 and all x ∈ X and assume 0 ∈ σ(S). Then there exists δ > 0 such that B(0, δ) ⊆ σR(S). The main result in [21] which states that if C is zero-class admissible and σR(A) is empty, then T (t) extends to a group. The next propositions extend this result. Proposition 3.6. Let (A, C) be an admissible BFC-system. {Reλ : λ ∈ σR(A)} is bounded, then (T (t))t≥0 extends to a group on X. If Re(σR(A)) := Proof. We know by Lemma 3.1 that ε(t) > 0. If T (t) was not boundedly invertible for some t > 0, then 0 ∈ σ(T (t)). By Lemma 3.5, there exists δt > 0 such that B(0, δt) ⊆ σR(T (t)). Since σR(T (t))\{0} = exp(−tσR(A)) (see [5, p. 276]) we obtain B(0, δt)\{0} ⊆ exp(−tσR(A)), Therefore, there exists a real sequence (λn) ∈ σR(A) such that Reλn → +∞. This contradicts the assumption. (cid:3) Proposition 3.7. Assume that (A, C) is an admissible BFC-system. If σ(A) does not contain a half-plane then (T (t))t≥0 extends to a group on X. Proof. By Corollary 3.4 we see that σ(A) is either contained in vertical strip or contains a half-plane. Now we apply Proposition 3.6 to conclude. (cid:3) Considering the right shift semigroup T (t) on L2(R+) with the identity obser- vation C = Id provides an example of a BFC-system (even a zero-class admissible one, see [21, Remark 3.1]) for which no group extension is possible. In this exam- ple, it is not difficult to check the spectrum of A satisfies σ(A) = σR(A) = C+ (the right half-plane). This shows that the spectral condition in Proposition 3.7 cannot 8 BERNHARD H. HAAK AND EL MAATI OUHABAZ be omitted. Assume that (A, C) is an admissible BFC-system. If T (t) is analytic, differentiable or merely eventually continuous then by [5, p. 113] σ(A) does not contain a half-plane. Thus, we conclude by Proposition 3.7 that (T (t))t≥0 extends to group on X. Proposition 3.8. Assume that (A, C) is an admissible BFC-system. If T (t) is compact for some t > 0, then X has finite dimension. Proof. If T (t) is compact for some t > 0 then σ(A) = σP (A) is discrete. It follows from Proposition 3.3 that σ(A) is bounded. We conclude by Proposition 3.6 that (T (t))t≥0 extends to a group on X. Thus, I = T (t)T (−t) is compact on X and therefore X has finite dimension. (cid:3) 4. Sufficient conditions for exact observability Our aim in this section is to derive conditions on C and A which imply exact observability. Our condition reads as follows (4.1) kCA−(1−β)xkY ≥ δkxk for all x ∈ D(A) ∩ R(A). Here β ∈ (0, 1) and δ > 0 are constants. Since we shall assume that A is injective, it may be convenient to understand (4.1) in the sense kCA−1xkY ≥ δkA−βxk. Of course, R(A) ∩ R(Aβ ) = R(A). In the sequel we need some basic properties of the H ∞ functional calculus for sectorial operators. This functional calculus goes back to the work of McIntosh[12]. More recent publications of the meanwhile rich theory can be found in [6] or [11] and the references given therein. We briefly sketch the needed results and definitions. Definition 4.1. We denote by Sω the open sector {z ∈ C∗ : arg(z) < ω} and by Sω the closure of Sω in C. We call a closed operator A on X sectorial of angle ω if A is densely defined having its spectrum in Sω such that λR(λ, A) := λ(λ−A)−1 of A is uniformly bounded on the complement of each strictly larger sectors Sθ, θ > ω. Notice that if −A generates a bounded semigroup T (t)t≥0, then A is sectorial of angle π/2 by the Hille-Yosida theorem. Moreover, the semigroup is (bounded) analytic if and only if A is sectorial of angle < π/2. Let H ∞(Sω) denote the holo- morphic and bounded functions on Sω and Let H ∞(Sω) denote the holomorphic and bounded functions on Sω that are continuous and bounded on Sω. We further consider the ideal H ∞ 0 (Sω)) of all functions f ∈ H ∞(Sω) (re- 0 (Sω)) that allow an estimate f (z) ≤ M max(zε, z−ε). The class spectively H ∞ of H ∞ 0 (Sω) functions admits a natural functional calculus for sectorial operators A of angle ω. Indeed, if f ∈ H ∞ 0 (Sθ) for some θ > ω and if Γ = ∂Sθ denotes the orientated path with strictly decreasing imaginary part, the Cauchy integral 0 (Sω) (respectively H ∞ f (A) := f (λ) R(λ, A) dλ 1 2πiZΓ converges absolutely in norm and defines therefore a bounded operator f (A). If A has, say, dense range, one obtains then a functional calculus for all functions f ∈ H ∞ 0 (Sθ). EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 9 Definition 4.2. We say that A admits a bounded H ∞(Sω) (respectively H ∞(Sω)) functional calculus if f (A) is a bounded operator on X and there exists a constant M such that (4.2) kf (A)k ≤ M kf k∞ for all f ∈ H ∞(Sθ) and θ > ω (respectively for all f ∈ H ∞(Sω)). Let us mention that by an approximation argument, if (4.2) holds for all f ∈ 0 (Sθ) then it holds for all f ∈ H ∞(Sθ). Therefore, it is enough to check the H ∞ 0 (Sθ) to obtain bounded H ∞(Sω) functional calculus. validity of (4.2) for all f ∈ H ∞ Similarly, if (4.2) holds for for all f ∈ H ∞ 0 (Sω) we obtain a H ∞(Sω) functional calculus. We say that A admits upper square function estimates on Sθ if there is a constant M > 0 such that (4.3) ∀x ∈ X : Z ∞ 0 kϕ(tA)xk2 dt t ≤ M 2kxk2 0 (Sθ). In the same way we speak of lower square function estimates for all ϕ ∈ H ∞ on Sθ if one has (4.4) ∀x ∈ X : kxk2 ≤ K 2Z ∞ 0 kϕ(tA)xk2 dt t 0 (Sµ) functions for all µ > ω are equivalent to a bounded H ∞ If X = H is a Hilbert space, then upper square function estimates for A and for A∗ for H ∞ 0 (Sµ) functional calculus for all µ > ω. Moreover, by an approximate identity argument and a duality estimate, lower square function estimates for A follow from upper estimate of its adjoint A∗. We will go into some details on the Hilbert space theory of the functional calculus in the last section and refer at this point to [12, 3] for more details. Before stating our first result of this section we discuss the following estimate for A (SQβ) kxk2 ≤ K 2Z ∞ 0 k(tA)−β(T (2t2β) − T (t2β))xk2 dt t that we will need to formulate the theorem. Here, K is a positive constant and β ∈ (0, 1). In case β = 1/2, letting ϕ(z) := z−1/2 (e−2z − e−z), this corresponds to a lower square function estimate (4.4) for ϕ ∈ H ∞ 0 (Sπ/2 ). As mentioned above, (4.4) follows from H ∞−functional calculus when X is a Hilbert space. We will discuss again this in the next section. Assume that −A is the generator of bounded strongly continuous semigroup on X and is injective. If we let ϕ(z) := z−β(e−2z − e−z), then (SQβ) can be seen as kxk2 ≤ K 2Z ∞ = K 2Z ∞ 2β Z ∞ K 2 t 0 (cid:13)(cid:13)(tA)−β(T (2t2β) − T (t2β))x(cid:13)(cid:13)2 dt 0 (cid:13)(cid:13)t2β2−βϕ(t2βA)x(cid:13)(cid:13)2 dt 0 (cid:13)(cid:13)s−(1/2−β)ϕ(sA)x(cid:13)(cid:13)2 ds s t , (letting s = t2β) = i.e. as 'weighted' lower square function estimate for ϕ ∈ H ∞ 0 (Sπ/2). 10 BERNHARD H. HAAK AND EL MAATI OUHABAZ By [6, Theorem 6.4.6], the completion X1/2−β,2 of X with respect to the seminorm is independent of the choice of ϕ and coincides (with equivalent norms) with the real interpolation space: [x]β =(cid:18)Z ∞ s (cid:19)1/2 0 (cid:13)(cid:13)s−(1/2−β)ϕ(sA)x(cid:13)(cid:13)2 ds X1(A)(cid:17) 3 = X1/2−β,2. 4 − β 2 ,2 (cid:16) X−1(A), (4.5) (4.6) Here X−1(A) is the completion of R(A) with respect to kA−1xk and X1(A) is the completion of D(A) with respect to kAxk. From (4.5), it follows that (SQβ) is equivalent to the continuous embedding (cid:16) X−1(A), X1(A)(cid:17) 3 4 − β 2 ,2 ֒→ X. For the rest of this discussion, we assume for simplicity that A is invertible. In this X1(A) = D(A) and X−1(A) = X−1. It is a known fact that the semigroup case, (T (t)) extends to a strongly continuous semigroup (T−1(t)) on the extrapolation space X−1, whose (negative) generator A−1 is an extension of A (see [5] Chapter II, Section 5). In addition A is the part of A−1 on X and hence D(A2 −1) = D(A) with equivalent norms. Indeed, x ∈ D(A2 −1) ⇔ x ∈ D(A−1) = X, A−1x ∈ D(A−1) ⇔ x ∈ X, A−1x ∈ D(A−1) ⇔ x ∈ D(A). The fact that A−1 : X → X−1 is an isometry implies that kAxkX = kA2 Assume now that β < 1 2 . We have −1xkX−1 . (X−1, D(A)) 3 4 − β −1)(cid:1) 3 2 ,2 =(cid:0)X−1, D(A2 =(cid:0)D(A−1), D(A2 −1)(cid:1) 1 4 − β = (X, D(A)) 1 2 ,2 2 −β,2 ֒→ X. 2 −β,2 Note that the second equality follows from [19, p. 105]. Hence for β < 1 2 (4.7) 4 − β which means that (SQβ) always holds for β < 1 2 . X1/2−β,2 = (X−1, D(A)) 3 2 ,2 ֒→ X, Let us finally mention that for β > 1 2 , (SQβ) never holds for non-negative self- adjoint operators with compact resolvent in infinite dimension separable Hilbert spaces. Indeed, consider such an operator A. The spectrum is discrete σ(A) = {λn} with λn → +∞. Applying (SQβ) to a normalised eigenvector x = ϕn (associated with λn) yields (e−2t2β − e−t2β )2 dt t 1 ≤ K 2Z ∞ 0 (cid:13)(cid:13)(tA)−β ϕn(cid:13)(cid:13)2 n Z ∞ n Z ∞ K 2 λ2β 0 K 2λn 2βλ2β = = 0 (e−2s − e−s)2 ds s2 t−2β(e−2t2β λn − e−t2β λn )2 dt t EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 11 = Cλ1−2β n . For β > 1 cannot hold. 2 , the last term goes to 0 as n tends to +∞. This shows that (SQβ) Now we come to our main result of this section concerning exact observability. Theorem 4.3. Let −A be the generator of a bounded semigroup on the Banach space X and assume that A is injective and has dense range. Let C : D(A) → Y be bounded and suppose that there exists β ∈ (0, 1) such that the lower square function estimate (SQβ) and (4.1) are satisfied. Then there exists a constant m > 0 such that (4.8) for all x ∈ D(A) ∩ R(A). m2kxk2 ≤Z ∞ 0 kCT (t)xk2 Y dt Proof. Fix x ∈ D(A) ∩ R(A) and apply the lower square function estimate (SQβ) to obtain 0 kxk2 ≤ K 2Z ∞ δ2 Z ∞ δ2 Z ∞ = K 2 ≤ K 2 0 0 k(tA)−β(T (2t2β) − T (t2β))xk2 dt t kCA−(1−β)(tA)−β (T (2t2β) − T (t2β))xk2 Y dt t kCA−1(T (2t2β) − T (t2β))xk2 Y dt t1+2β . Using the fact that x ∈ D(A), we can write CA−1(T (2t2β) − T (t2β))x = −CA−1Z 2t2β t2β Using this and the previous estimates yields AT (s)x ds = −Z 2t2β t2β CT (s)x ds. 0 t2β t2β ≤ K 2 kxk2 ≤ K 2 δ2 Z ∞ δ2 Z ∞ δ2 Z ∞ δ2 Z 2 kZ 2t2β 0 (cid:18)Z 2t2β 0 Z 2t2β 1 Z ∞ 2βδ2 Z ∞ This shows (4.8) with m =plog(2)/2β K 0 = log(2) K 2 ≤ K 2 = K 2 t2β 0 δ . CT (s)x dsk2 Y dt t1+2β 1 · kCT (s)xkY ds(cid:19)2 dt t1+2β kCT (s)xk2 Y ds dt t kCT (t2βu)xk2 Y t2β−1 dt du kCT (r)xk2 Y dr. Notice that if 0 ∈ (A) then R(A) = X and hence D(A) ∩ R(A) = D(A). In this case (4.8) holds for all x ∈ D(A) and this means that C is exactly observable for A. Corollary 4.4. Let −A be the generator of a bounded semigroup on the Banach space X and assume that A is boundedly invertible. Let C : D(A) → Y be a bounded (cid:3) 12 BERNHARD H. HAAK AND EL MAATI OUHABAZ operator and assume that there exists β ∈ (0, 1) such that the lower square function estimate (SQβ) and (4.1) are satisfied. Then C is exactly observable for A. Corollary 4.5. Let −A be the generator of a bounded semigroup on the Banach space X and assume that A is injective and has dense range. Let C : D(A) → Y be infinite-time admissible for A and assume that there exists β ∈ (0, 1) such that the lower square function estimate (SQβ) and (4.1) are satisfied. Then C is exactly observable for A. Proof. It remains now to extend the estimate (4.8) from Theorem 4.3 for all x ∈ D(A). This is an easy task. For x ∈ D(A) the sequence xn := n(n+A)−1x − n−1(n−1+A)−1x ∈ D(A) ∩ R(A) converges to x in D(A) (for the graph norm). Therefore, Cxn converges in Y to Y dt since C is supposed to (cid:3) be infinite-time admissible. We obtain (4.8) for all x ∈ D(A). Y dt converges to R ∞ Cx and R ∞ 0 kCT (t)xnk2 0 kCT (t)xk2 In the next corollary we obtain a criterion for finite time exact observability. Corollary 4.6. Let −A be the generator of a bounded semigroup on the Banach space X and assume that there exists a constant ω > 0 such that A + ω satisfies (SQβ). Assume that C : D(A) → Y is bounded and that kC(ω + A)−(1−β)xkY ≥ δkxk for x ∈ D(A). Then C is exactly observable in finite time. Proof. We apply the previous theorem to ω + A and obtain kxk2 ≤ MZ ∞ 0 kCe−ωtT (t)xk2 Y dt for all x ∈ D(A). We split the right hand side into two parts and write kCe−ωtT (t)xk2 kCe−ωtT (t)xk2 kCe−ωtT (t)xk2 Y dt Z ∞ 0 0 Y dt =Z τ ≤Z τ =Z τ 0 0 Y dt +Z ∞ τ Y dt +Z ∞ Y dt + e−ωτZ ∞ 0 0 kCT (t)xk2 kCe−ω(τ +t)T (t + τ )xk2 Y dt kCT (t)xk2 kCe−ωtT (t)T (τ )xk2 Y dt. Since C is infinite-time admissible for e−ωtT (t) and the semigroup T (t) is bounded we have for some constants M ′, M ′′ Z ∞ 0 kCe−ωtT (t)T (τ )xk2 Y dt ≤ M ′kT (τ )xk2 ≤ M ′′kxk2. Therefore, kxk2 ≤ MZ τ 0 kCT (t)xk2 Y dt + M M ′′e−ωτ kxk2. If we choose τ large enough such that M M ′′e−ωτ < 1 we obtain the desired in- equality. (cid:3) As explained at the beginning of this section, (SQβ) holds for all β < 1 2 and all generators of bounded semigroup (see (4.7)). Therefore, applying Theorem 4.3 with β = 1 2 − ε, we obtain the following corollary. EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 13 Corollary 4.7. Let −A be the generator of a bounded semigroup (T (t))t≥0 on a X and assume that A is invertible. Then, if kCA−1/2+εxk ≥ δkxk for some ε, δ > 0 and all x ∈ D(A), C is infinite-time exactly observable for A. 5. Exact observability on Hilbert spaces Proposition 5.1. Let X and Y be Hilbert spaces. Then, if (A, C) is exactly observ- able and admissible in infinite time, T (t)t≥0 is similar to a contraction semigroup. Proof. Denote by hx, yiY the scalar product of Y and define for x, y ∈ D(A) hx, yi∼ X :=Z ∞ 0 hCT (t)x, CT (t)yiY dt. This is clearly a bilinear (or sesquilinear) form on D(A) × D(A). Admissibility and exact observability imply that kxkX and kxk∼ X are equivalent. By density, we extend this to all x ∈ X and k · k∼ X is associated with a scalar product on X. With respect to the new norm, kT (t)xk∼ X =(cid:16)Z ∞ 0 kCT (t+s)xk2 ds(cid:17)1/2 =(cid:16)Z ∞ t kCT (s)xk2 ds(cid:17)1/2 ≤ kxk∼ X and so T (t)t≥0 is a contraction semigroup with respect to k · k∼ X . (cid:3) Now we turn back to Theorem 4.3. As mentioned at the beginning of the previous section, the lower square function estimate (SQβ) holds for small β for all generators of bounded strongly continuous semigroups. It is then tempting to use the theorem for small β in order to include a large class of semigroups T (t). On the other hand, if we assume that 0 ∈ (A) and kCA−α−εxk ≥ δkxk, one also has kCA−αxk = kCA−α−εAεxk ≥ δkAεxk ≥ δ′kxk. That is, the invertibility condition on CA−(1−β) becomes more restrictive when β decreases. To admit more observation operators C one therefore seeks for values of β large enough. Combining both conditions forces to play with different values of β in different situations. In the following corollary we choose β = 1/2. Corollary 5.2. Let −A be the generator of a semigroup of contractions (T (t))t≥0 on a Hilbert space X. If A has dense range and kCA−1/2xk ≥ δkxk for all x ∈ D(A) ∩ R(A), then (5.1) m2kxk2 ≤Z ∞ 0 kCT (t)xk2 Y dt for all x ∈ D(A) ∩ R(A). In addition, if either A is invertible or C is infinite-time admissible for A then C is infinite-time exactly observable for A. Notice that in view of Proposition 5.1, the hypothesis of a semigroup of contrac- tions is necessity to be able to conclude in the case that C is admissible. Proof. By the Lumer-Phillips theorem, A is an accretive operator, i.e. RehAx, xi ≥ 0 for all x ∈ D(A). Since A has dense range and X is reflexive, A is actually injective (cf. [3, Theorem. 3.8]). We need to verify that A admits lower square function estimates (4.4) with functions ϕ that are bounded holomorphic on Sπ/2 and continuous on Sπ/2. We shall explain how this follows from the functional calculus. First notice that A has a bounded H ∞(Sπ/2) functional calculus. This 14 BERNHARD H. HAAK AND EL MAATI OUHABAZ means that for every bounded holomorphic function ϕ on Sπ/2 and continuous on Sπ/2, ϕ(A) is well defined and kϕ(A)kL(X) ≤ M sup z∈Sπ/2 ϕ(z). This is essentially von Neumann's inequality for contractions on a Hilbert space. Indeed, if A is accretive, T := (A − 1)(A + 1)−1 is a contraction and von Neumann's inequality states kp(T )k ≤ kpkH∞(D) for every polynomial p. From this, the H ∞- calculus can be derived by approximation arguments. Two different direct proofs for the boundedness of the H ∞(Sπ/2) calculus are given in [6, Theorem 7.1.7]. One uses a dilation theorem of the semigroup into a unitary C0-group due to Sz.-Nagy. The second exploits accretivity of A from a 'numerical range' viewpoint and can be seen as the most simple case of the Crouzeix-Delyon theorems [4, 2]. Having the boundedness of the functional calculus on Sπ/2 in hands we certainly have upper 0 (Sθ) when θ > π/2. This is well known square function estimates for functions in H ∞ and proved by McIntosh [12]. For the particular functions ψα(z) := zα/(1 + z) for α ∈ (0, 1) this means that for some positive constants kα kαZ ∞ kψα(tA)xk2 dt t ≤ kxk2. (5.2) 0 Given now a function ψ ∈ H ∞ M max(z2ε, z−2ε) and write 0 (Sπ/2), we choose ε > 0 small such that ψ(z) ≤ ψ(z) = ψε(z) × z−εψ(z) + ψ1−ε(z) × zεψ(z). Notice that z±εψ(z) ∈ H ∞ 0 (Sπ/2). Therefore, upper square function estimate for A with the function ψ follow from (5.2) using the boundedness of the H ∞(Sπ/2) calculus. All what we explain here works also for the adjoint A∗. By a duality argument as in [12] or [3], we pass from upper square function estimates for A∗ to lower square function estimates for A. As a particular case, (SQβ) holds with β = 1 (cid:3) 2 . We then apply Theorem 4.3. Again as in Corollary 4.6, we can obtain observability in finite time by adding a constant w to A. That is Corollary 5.3. Let −A be the generator of a semigroup (T (t)) on the Hilbert space X. Suppose that there exists a constant ω > 0 such that A + ω is accretive and that C : D(A) → Y is bounded. If kC(A + ω)−1/2xk ≥ δkxk for all x ∈ D(A) and some δ > 0, then C is exactly observable in finite time. Examples 5.4. 1- Consider the Schrodinger equation = i∆u = 0 on ∂Ω ∂u ∂t u y(t) = Cu(t) = ∇u(t)   with X = L2(Ω), Y = (L2(Ω))d and Ω is any open subset of Rd with boundary ∂Ω. In this problem, ∆ denotes the Laplacian with Dirichlet boundary conditions. It is (the negative of ) the associated operator with the symmetric form a(u, v) :=ZΩ ∇u∇vdx, D(a) = W 1,2 0 (Ω). EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 15 Since k(−i∆)1/2uk2 2 = k(−∆)1/2uk2 2 = kCuk2 2 we may conclude from Corollary 5.2 that (5.3) δkf k2 2 ≤Z ∞ 0 k∇eit∆f k2 2 dt for all f ∈ D(∆) ∩ R(∆). Note that we can replace here the Dirichlet boundary condition by the Neumann one. The arguments are the same, we just need to replace D(a) = W 1,2 (Ω) by D(a) = W 1,2(Ω). The same method applies for more general boundary conditions. 2- Let A be the uniformly elliptic operator 0 A = − dXj,k=1 ∂ ∂xk(cid:0)ajk ∂ ∂xj(cid:1) with bounded measurable coefficients ajk ∈ L∞(Rd). The operator A is defined by sesquilinear form techniques (see for example [15]) and note that A is not necessar- ily self-adjoint. It is a standard fact that −A generates a contraction semigroup on L2(Rd). Con- sider the problem (cid:26) ∂u = −Au ∂t y(t) = Cu(t) = ∇u(t) with X = L2(Rd) and Y = (L2(Rd)d. By the solution of the Kato's square root problem (see [1]), it is known that (5.4) k∇uk2 ≈ kA 1/2ku ∀u ∈ W 1,2(Rd). On the other hand, the accretivity of A implies an H ∞ functional calculus on L2(Rd). Hence it satisfies upper and lower square function estimate (4.3) and (4.4). In particular, (5.5) kf k2 2 ≈Z ∞ 0 kA 1/2e−tAf k2 2 dt. This implies that C = ∇ is both admissible and exactly observable for A in infinite time. Note that the semigroup e−tA is bounded holomorphic on some sector Sω and e−teiwA is a contraction on L2(Rd) (see [15]). Taking the maximal angle w, we ob- tain a contraction semigroup e−teiw A which is not holomorphic. Since (eiwA) 1/2 = 1/2 we see that (5.4) holds with eiwA in place of A. By Corollary 5.2 we have eiw/2A (5.6) δkf k2 2 ≤Z ∞ 0 k∇e−teiwAf k2 2 dt for f ∈ D(A) ∩ R(A). We may consider the same problem on a bounded Lipschitz domain instead of Rd. In this case, A is invertible. Hence eiwA is also invertible and we obtain (5.6) for all f ∈ D(A). This means that C is exactly observable for eiwA. 16 BERNHARD H. HAAK AND EL MAATI OUHABAZ References [1] Pascal Auscher, Steve Hofmann, Michael Lacey, Alan McIntosh, and Ph. Tchamitchian, The solution of the Kato square root problem for second order elliptic operators on Rn, Ann. of Math. (2) 156 (2002), no. 2, 633 -- 654. [2] Catalin Badea, Michel Crouzeix, and Bernard Delyon, Convex domains and K-spectral sets, Math. Z. 252 (2006), no. 2, 345 -- 365. [3] Michael Cowling, Ian Doust, Alan McIntosh, and Atsushi Yagi, Banach space operators with a bounded H∞ functional calculus, J. Austral. Math. Soc. Ser. A 60 (1996), no. 1, 51 -- 89. [4] Michel Crouzeix and Bernard Delyon, Some estimates for analytic functions of strip or sectorial operators, Arch. Math. (Basel) 81 (2003), no. 5, 559 -- 566. [5] Klaus-Jochen Engel and Rainer Nagel, One-parameter semigroups for linear evolution equa- tions, Graduate Texts in Mathematics, vol. 194, Springer-Verlag, New York, 2000, With contributions by S. Brendle, M. Campiti, T. Hahn, G. Metafune, G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R. Schnaubelt. [6] Markus Haase, The functional calculus for sectorial operators, Operator Theory, Advances and Applications, vol. 169, Birkhauser Verlag, 2006. [7] Einar Hille and Ralph S. Phillips, Functional analysis and semi-groups, American Mathemat- ical Society Colloquium Publications, vol. 31, American Mathematical Society, Providence, R. I., 1957, rev. ed. [8] Birgit Jacob and Jonathan R. Partington, Admissibility of control and observation operators for semigroups: a survey, Current trends in operator theory and its applications, Oper. Theory Adv. Appl., vol. 149, Birkhauser, Basel, 2004, pp. 199 -- 221. [9] Birgit Jacob, Jonathan R. Partington, and Sandra Pott, Zero-class admissibility of observa- tion operators, Systems Control Lett. 58 (2009), no. 6, 406 -- 412. [10] Vilmos Komornik and Paola Loreti, Fourier series in control theory, Springer Monographs in Mathematics, Springer-Verlag, New York, 2005. [11] Peer C. Kunstmann and Lutz Weis, Maximal L p-regularity for parabolic equations, Fourier multiplier theorems and H∞-functional calculus, Functional analytic methods for evolution equations, Lecture Notes in Math., vol. 1855, Springer, Berlin, 2004, pp. 65 -- 311. [12] Alan McIntosh, Operators which have an H∞ functional calculus, Miniconference on opera- tor theory and partial differential equations (North Ryde, 1986), Proc. Centre Math. Anal. Austral. Nat. Univ., vol. 14, Austral. Nat. Univ., Canberra, 1986, pp. 210 -- 231. [13] Luc Miller, A direct Lebeau-Robbiano strategy for the observability of heat-like semigroups, Discrete Contin. Dyn. Syst. Ser. B 14 (2010), no. 4, 1465 -- 1485. [14] Nikolai Nikolski, Rudiments de la thorie du contrle : vision operationnelle, Publication de l'Ecole Doctorale de mathmatiqes de Bordeaux, 1996, Lecture notes of a course given 1994- 1995 in Bordeaux. [15] El Maati Ouhabaz, Analysis of heat equations on domains, London Mathematical Society Monographs Series, vol. 31, Princeton University Press, Princeton, NJ, 2005. [16] A. Pazy, Semigroups of linear operators and applications to partial differential equations, Applied Mathematical Sciences, vol. 44, Springer-Verlag, New York, 1983. [17] Thomas I. Seidman, On uniform null controllability and blowup estimates, Control theory of partial differential equations, Lect. Notes Pure Appl. Math., vol. 242, Chapman & Hall/CRC, Boca Raton, FL, 2005, pp. 213 -- 227. [18] Tenenbaum, G´erald and Tucsnak, Marius, On the null-controllability of diffusion equations, ESAIM: COCV (2010). [19] Hans Triebel, Interpolation theory, function spaces, differential operators, second ed., Johann Ambrosius Barth, Heidelberg, 1995. [20] Marius Tucsnak and George Weiss, Observation and Control for Operator Semigroups, Birkhauser Verlag, Basel, 2009. [21] Gen Qi Xu, Chao Liu, and Siu Pang Yung, Necessary conditions for the exact observability of systems on Hilbert spaces, Systems Control Lett. 57 (2008), no. 3, 222 -- 227. [22] Robert M. Young, An introduction to nonharmonic Fourier series, first ed., Academic Press Inc., San Diego, CA, 2001. EXACT OBSERVABILITY, SQUARE FUNCTIONS AND SPECTRAL THEORY 17 Institut de Math´ematiques de Bordeaux, Universit´e Bordeaux 1, 351, cours de la Lib´eration, 33405 Talence CEDEX, FRANCE E-mail address: [email protected] Institut de Math´ematiques de Bordeaux, Universit´e Bordeaux 1, 351, cours de la Lib´eration, 33405 Talence CEDEX, FRANCE E-mail address: [email protected]
1005.3634
1
1005
2010-05-20T08:53:42
Li-Yorke and distributionally chaotic operators
[ "math.FA" ]
We study Li-Yorke chaos and distributional chaos for operators on Banach spaces. More precisely, we characterize Li-Yorke chaos in terms of the existence of irregular vectors. Sufficient "computable" criteria for distributional and Li-Yorke chaos are given, together with the existence of dense scrambled sets under some additional conditions. We also obtain certain spectral properties. Finally, we show that every infinite dimensional separable Banach space admits a distributionally chaotic operator which is also hypercyclic.
math.FA
math
Li-Yorke and distributionally chaotic operators1 T. Berm´udeza, A. Bonillaa, F. Mart´ınez-Gim´enezb, A. Perisb aDepartamento de An´alisis Matem´atico, Universidad de La Laguna, 38271, La Laguna (Tenerife), Spain bIUMPA, Universitat Polit`ecnica de Val`encia, Departament de Matem`atica Aplicada, Edifici 7A, 46022 Val`encia, Spain. Abstract We study Li-Yorke chaos and distributional chaos for operators on Banach spaces. More pre- cisely, we characterize Li-Yorke chaos in terms of the existence of irregular vectors. Sufficient "computable" criteria for distributional and Li-Yorke chaos are given, together with the exis- tence of dense scrambled sets under some additional conditions. We also obtain certain spectral properties. Finally, we show that every infinite dimensional separable Banach space admits a distributionally chaotic operator which is also hypercyclic. Keywords: Li-Yorke chaos, distributional chaos, irregular vector, distributionally irregular vectors, weighted shift operators 2000 MSC: 47A16 1. Introduction During the last years many researchers paid attention to the "wild behaviour" of orbits gov- erned by linear operators on infinite dimensional spaces (more especially, on Banach or Fr´echet spaces). One of the most significant cases being the hypercyclicity, that is, the existence of vec- tors x ∈ X such that the orbit Orb(T, x) := {x, T x, T 2x, . . . } under a (continuous and linear) operator T : X → X on a topological vector space (usually, Banach or Fr´echet space) X, is dense in X. We refer the reader to the recent books [2] and [16] for a thorough account of the subject. This notion from operator theory joined chaos after the definition of Devaney [11], which (in our context) requires hypercyclicity and density of the set of periodic points of T in X. The concept of "chaos" appeared for the first time in the Mathematical literature in the paper of Li and Yorke [20] of mid 70's. Definition 1. Let (X, d) be a metric space. A continuous map f : X → X is called Li-Yorke chaotic if there exists an uncountable subset Γ ⊂ X such that for every pair x, y ∈ Γ of distinct points we have d( f nx, f ny) = 0 and lim sup d( f nx, f ny) > 0. lim inf n n In this case, Γ is a scrambled set and {x, y} ⊂ Γ a Li-Yorke pair. 1Dedicated to Professor Antonio Martin´on on his 60th birthday Email addresses: [email protected] (T. Berm´udez), [email protected] (A. Bonilla), [email protected] (F. Mart´ınez-Gim´enez), [email protected] (A. Peris) Preprint submitted to Elsevier October 27, 2018 Li-Yorke chaos was studied for linear operators, for instance, in [12, 14]. Incidentally, every hypercyclic operator T : X → X on a Fr´echet space X is Li-Yorke chaotic with respect to any (continuous) translation invariant metric d on X: E.g., fix a hypercyclic vector x ∈ X and consider the segment Γ := {λx ; λ ≤ 1}, which is a scrambled set for T . The notion of distributional chaos was introduced by Schweizer and Smital in [28]. The definition was stated for interval maps with the intention to unify different notions of chaos for continuous maps on intervals. This concept was widely studied by several authors and can be formulated in any metric space. In particular, it is considered for linear operators defined on Banach or Fr´echet spaces in [25, 22, 17, 18, 19]. For any pair {x, y} ⊂ X and for each n ∈ N, the distributional function Fn xy : R+ → [0, 1] is defined by Fn xy(τ) = 1 n card{0 ≤ i ≤ n − 1 : d( f ix, f iy) < τ} where card{A} denotes the cardinality of the set A. Define Fxy(τ) = lim inf n→∞ F ∗ xy(τ) = lim sup n→∞ Fn xy(τ) Fn xy(τ). Definition 2 ([28, 26]). A continuous map f : X → X on a metric space X is distributionally chaotic if there exist an uncountable subset Γ ⊂ X and ε > 0 such that for every τ > 0 and each pair of distinct points x, y ∈ Γ, we have that F ∗ xy(τ) = 1 and Fxy(ε) = 0. The set Γ is a distributionally ε-scrambled set and the pair {x, y} a distributionally chaotic pair. Moreover, f exhibits dense distributional chaos if the set Γ may be chosen to be dense. Generally speaking, for any distinct x, y ∈ Γ the iterations of these points are arbitrary close and ε separated alternatively, but additionally there are time intervals where any of these exclud- ing possibilities is much more frequent than the other. Given A ⊂ N, its upper and lower density are defined by dens(A) = lim sup n card{A ∩ [1, n]} n , and dens(A) = lim inf n card{A ∩ [1, n]} n , respectively. With these concepts in mind, one can equivalently say that f is distributionally chaotic on Γ if there exists ε > 0 such that for any x, y ∈ Γ, x , y, we have dens{n ∈ N ; d( f nx, f ny) < ε} = 0, and dens{n ∈ N ; d( f nx, f ny) < τ} = 1, for every τ > 0. From now on X will be a Banach space and T : X → X a bounded operator. In this case the associated distance is d(x, y) = kx − yk, with x, y ∈ X, where k·k is the norm of X. We recall the following concept from operator theory. Definition 3 (Beauzamy [3]). A vector x ∈ X is said to be irregular for T if lim infnkT nxk = 0 and lim supnkT nxk = ∞. Inspired by this definition, and by the notion of distributional chaos, we consider the follow- ing stronger property. Definition 4. A vector x ∈ X is said to be distributionally irregular for T if there are increasing sequences of integers A = (nk)k and B = (mk)k such that dens(A) = dens(B) = 1, limkkT nk xk = 0 and limkkT mk xk = ∞. 2 2. Li-Yorke chaotic operators We first discuss Li-Yorke chaos for operators with the following result that establishes the equivalence between Li-Yorke chaos and the existence of an irregular vector. Theorem 5. Let T : X → X be an operator. The following assertions are equivalent: (i) T is Li-Yorke chaotic. (ii) T admits a Li-Yorke pair. (iii) T admits an irregular vector. Proof. (i) implies (ii) is clear; in order to prove (ii) implies (iii) suppose that T admits y, z ∈ X with lim inf n d(T ny, T nz) = 0 and lim sup d(T ny, T nz) > 0. n If we set x = y − z then lim infnkT nxk = 0 and there is δ > 0 such that lim supnkT nxk > δ. If lim supnkT nxk = ∞ then x is an irregular vector. Otherwise, M := lim supnkT nxk < ∞. We observe that kT k > 1 since, given n, m ∈ N, n < m, with kT nxk < δ/2 and kT mxk > δ, we have that kT km−n > 2. We select a strictly increasing sequence of integers (nk)k such that the sequence of vectors (T nk x)k tends to 0 fast enough so that kT n2k xk < 4−k, k Xj=0 kT n2 j+n2k+1 xk 4 jkT kn2 j−1 kT n2 j xk < kT n2k xk, and δ 4kkT kn2k−1 kT n2k xk − k−1 Xj=0 M 4 jkT kn2 j−1 kT n2 j xk > k, (1) (2) for all k ∈ N, where n−1 = n0 := 0. We now set u = ∞ Xj=0 1 4 jkT kn2 j−1 kT n2 j xk T n2 j x. We will see that u is an irregular vector for T . Indeed, kT n2k+1 uk ≤ kT n2 j+n2k+1 xk 4 jkT kn2 j−1 kT n2 j xk ∞ Xj=0 < kT n2k xk + ∞ Xj=k+1 kT n2k+1 kkT n2 j xk 4 jkT kn2 j−1 kT n2 j xk < 4−k + ∞ Xj=k+1 kT kn2k+1 4 jkT kn2 j−1 < 2−k, by (1) and the selection of T n2k x, k ∈ N. On the other hand, let (mk)k be a strictly increasing sequence of integers such that kT mk xk > δ, k ∈ N. Without loss of generality we suppose that 3 n1 < m1 < n2 < m2 < . . . with m2k − n2k tending to infinity. By (2) we obtain that kT m2k xk 4kkT kn2k−1 kT n2k xk kT m2k−n2kuk ≥ > 4kkT kn2k−1 kT n2k xk − δ k−1 −Xj,k Xj=0 1 4 j > k − 4−k, > k − ∞ Xj=k+1 kT m2k−n2k+n2 j xk 4 jkT kn2 j−1 kT n2 j xk M 4 jkT kn2 j−1 kT n2 j xk − ∞ Xj=k+1 kT km2k−n2k 4 jkT kn2 j−1 for all k ∈ N. Thus lim supnkT nuk = ∞ and u is an irregular vector. Finally, to see (iii) implies (i), if u ∈ X is an irregular vector for T , then S := span{u} is a scrambled set for T by a direct application of the definitions. Prajitura studied in [27] some properties of operators having irregular vectors. By [27] and Theorem 5, we obtain certain properties of Li-Yorke chaotic operators. Corollary 6. Let T : X → X be a Li-Yorke chaotic operator. The following assertions hold: 1. σ(T ) ∩ ∂D , ∅. 2. T n is Li-Yorke chaotic for every n ∈ N. 3. T is not compact. 4. T is not normal. Definition 7. An operator T : X → X satisfies the Li-Yorke Chaos Criterion (LYCC) if there exist an increasing sequence of integers (nk)k and a subset X0 ⊂ X such that (a) lim k→∞ (b) sup n T nk x = 0, x ∈ X0, kT nY k = ∞, where Y := span(X0) and T nY denotes the restriction operator of T n to Y. One might expect that the LYCC is a sufficient condition for Li-Yorke chaos. It turns out that it is in fact a characterization of this phenomenon. Theorem 8. Let T : X → X be an operator. The following are equivalent: (i) T is Li-Yorke chaotic. (ii) T satisfies the Li-Yorke Chaos Criterion. Proof. Suppose T is Li-Yorke chaotic, by Theorem 5 we find an irregular vector x ∈ X. By setting X0 = {x}, we verify the LYCC easily. Conversely, suppose T satisfies the LYCC, let (nk)k, and X0 ⊂ X be the respective sequence of integers and subset of vectors satisfying conditions (a) and (b) of Definition 7. If there is x ∈ X0 such that x is an irregular vector, then we are done. Otherwise we observe that every u ∈ span(X0) satisfies limk T nku = 0 and supnkT nuk < ∞. Passing to subsequences of (nk)k and (mk)k, if necessary, by property (b) and since supnkT nY k = ∞, we can obtain a sequence (u j) j of normalized vectors in span(X0) such that (a)' kT nk u jk < 1 j , j = 1, . . . , k, k ∈ N, and 4 (b)' kT m ju jk > 3 jM j−1, j > 1 where Mk := sup{kT nuik ; i = 1, . . . , k, n ≥ 0} < ∞, k ∈ N. Without loss of generality, we may suppose that m1 < n1 < m2 < n2 < . . . . We select any infinite subset I ⊂ N such that, for each j ∈ N, if i ∈ I with i > j, then 2i > 2 jkT kn j . The vector u := Xi∈I 1 2i ui is well-defined since the series is convergent. We will see that u is an irregular vector for T . On one hand, kT m j uk ≥ > 1 2 j kT m ju jk − Xi∈I, i, j 3 jM j−1 − Xi∈I, i< j 2 j 1 2i kT m j uik M j−1 2i − Xi∈I, i> j 1 2i kT m juik > 3 j 2 j − 1! M j−1 −Xi≥ j 1 2i −→ j→∞, j∈I +∞. On the other hand, kT n juk ≤ Xi∈I, i≤ j j−1 2i + Xi∈I, i> j 1 2i kT n juik < 1 j + 1 2 j−1 , j ∈ I, which shows that u is an irregular vector. 3. The strong criterion for distributional chaos and spectral properties The following criterion for distributional chaos was introduced in [17]. Since, for several reasons that will be clarified soon, this criterion is somehow very restrictive, we will call it the "strong" criterion. In this section we will study the spectral properties of operators that satisfy this criterion. Definition 9. An operator T : X → X satisfies the Strong Distributional Chaos Criterion (SDCC), if there is a constant r > 1 such that for any m ∈ N, there exists xm ∈ X \ {0} sat- isfying kT k xmk = 0, (i) lim k→∞ (ii) kT ixmk ≥ rikxmk for i = 1, 2, . . . , m. Such r is said to be a SDCC-constant for the operator T . Theorem 10 ([17, Theorem 3.3]). Let T : X → X be an operator. If T satisfies the Strong Distributional Chaos Criterion, then T is distributionally chaotic. By D we mean the open unit disc, its boundary is ∂D and the complement of the closed unit disc will be written as D c . We denote by r(T ) the spectral radius of T . 5 Proposition 11. Let T : X → X be an operator. The following properties hold: (a) If there exists r > 1 such that for all m ∈ N there exists xm ∈ X \ {0} with kT ixmk ≥ rikxmk, for all i = 1, . . . , m, (3) then r(T ) ≥ r. (b) [13, Lemma 6.4 (b)] If σ(T ) ⊂ D c and limk→∞ kT k xk = 0, then x = 0. Proof. (a) Assume that r(T ) < r. Let ε > 0 be such that r(T ) + ε < r. Then there exists m ∈ N such that for all n ≥ m we have that kT nk1/n ≤ r(T ) + ε < r. Moreover, by (3) for m + 1 there exists x ∈ X \ {0} such that kT ixk ≥ rikxk for i = 1, . . . , m + 1. Then kT m+1 x kxk k ≥ rm+1, so kT m+1k1/m+1 ≥ r, which is a contradiction. Corollary 12. If T : X → X satisfies the Strong Distributional Chaos Criterion with SDCC- constant r, then r(T ) ≥ r. Theorem 13. If T satisfies the Strong Distributional Chaos Criterion with SDCC-constant r, then the following properties hold: (a) For any r0 with 1 ≤ r0 ≤ r we have σ(T ) ∩ ∂D(0, r0) , ∅. (b) There are not disjoint closed subsets F1 and F2 of C such that F1 ⊂ D(0, r) and F2 ⊂ D c with F1 ∪ F2 = σ(T ). Proof. (a) Assume that there exists r0 ≤ r such that σ(T ) ∩ ∂D(0, r0) = ∅. Take the spectral c decomposition σ1 = D(0, r0) ∩ σ(T ) and σ2 = D(0, r0) ∩ σ(T ). Let Xi and Ti with i = 1, 2 be given by the above spectral decomposition such that X = X1⊕X2, T = T1⊕T2 with σi = σ(Ti), i = 1, 2. Let m ∈ N. Then there exists xm = x1 mk → 0 as k tends to infinity and kT ixmk ≥ rikxmk for i = 1, . . . , m. Using the same argument as in [13, Lemma 6.4 (b)] we have that kx2 m = 0. Hence xm = x1 mk ≥ rikxmk for i = 1, . . . , m. Henceforth, r(T1) ≥ r but this is a contradiction since σ(T1) ⊂ D(0, r0). mk, which tends to 0 as k → ∞, that is, x2 m ∈ X\{0} such that kT k xmk = kT k m and kT k xmk = kT k mk ≤ kT k mk+kT k m⊕x2 2 x2 1 x1 2 x2 1 x1 c (b) Let F1 ⊂ D(0, r) and F2 ⊂ D be closed subsets of σ(T ) such that F1 ∩ F2 = ∅ and F1 ∪ F2 = σ(T ). Let Xi and Ti with i = 1, 2 be given by this spectral decomposition, i.e., X = X1 ⊕ X2, T = T1 ⊕ T2 with Fi = σ(Ti), i = 1, 2. Let m ∈ N. Then there exists xm = x1 m ∈ X \ {0} such that kT k xmk = kT k mk → 0 as k tends to infinity and kT ixmk ≥ rikxmk for i = 1, . . . , m. By part (b) of Proposition 11 we have that x2 m = 0. Then by part (a) of Proposition 11 we obtain that r(T1) ≥ r which is a contradiction since σ(T1) = σ1 ⊂ D(0, r). mk + kT k m ⊕ x2 1 x1 2 x2 An operator S : X → X is called strictly singular if for every infinite dimensional subspace M of X, the restriction of S to M is not a homeomorphism. An example of strictly singular operator is the compact operators. Notice that a small compact perturbation of the unit operator could be distributionally chaotic [18, Proposition 3.5]. Also, in [18, Proposition 3.3] is proven that any compact perturbation of scalar operator can not satisfy the SDCC. In the next result we improve that property, that is, if S is a strictly singular operator, then S + λI does not satisfy the SDCC. Corollary 14. If S : X → X is a strictly singular operator and λ ∈ C, then λI + S does not satisfy the SDCC. 6 Proof. It is well know that the spectrum of strictly singular operator is at most a countable sequence of eigenvalues that converges to zero as its only limit point. Hence the result is a consequence of part (b) of Theorem 13. A hereditarily indecomposable Banach space is an infinite dimensional space such that no subspace can be written as a topological sum of two infinite dimensional subspaces. W.T. Gowers and B. Maurey constructed the first example of a hereditarily indecomposable space [15]. In particular, they proved that every operator T defined on a hereditarily indecomposable space X can be written as T = λI + S , where λ ∈ C and S is a strictly singular operator. The class of hereditarily indecomposable spaces appeared for the first time in relation with hypercyclicity and chaos in [9], where it was proved that some complex separable Banach spaces admit no Devaney chaotic operator. Corollary 15. There are no operators satisfying the SDCC on any hereditarily indecomposable Banach space. 4. Distributionally chaotic operators Our purpose in this section is to study distributional chaos in connection with the existence of distributionally irregular vectors, and to give useful criteria for distributional chaos. First of all, we obtain the analogous of one of the implications in Theorem 5. Proposition 16. If T : X → X admits a distributionally irregular vector x ∈ X, then T is distributionally chaotic. Proof. If x ∈ X is a distributionally irregular vector for T , then there exist increasing sequences of integers A = (nk)k and B = (mk)k such that dens(A) = dens(B) = 1, limkkT nk xk = 0 and limkkT mk xk = ∞. Let S := span{x}. If y, z ∈ S , y , z, then y − z = αx with α , 0. Therefore, kT nk y − T nkzk = α kT nk xk kT mky − T mkzk = α kT mk xk k→∞ −→ 0, and k→∞ −→ ∞, which yields that S is a distributionally ε-scrambled set for T , for every ε > 0. Problem 1. Does every distributionally chaotic operator T : X → X admit a distributionally irregular vector? The following result from [17] yields a sufficient condition for distributional chaos, less re- strictive than the SDCC. Theorem 17 ([17]). Let T : X → X be an operator. If for any sequence of positive numbers (Cm)m increasing to ∞, there exists (xm)m in X \ {0} satisfying T nxm = 0, and (a) lim n→∞ (b) there is a sequence of positive integers (Nm)m increasing to ∞, such that lim m→∞ 1 Nm card{0 ≤ i < Nm ; kT ixmk ≥ Cmkxmk} = 1, 7 then T is distributionally chaotic. We introduce a variation of the above criterion by allowing to have different sequences of vectors in part (a) and (b) of the above theorem. Definition 18. An operator T : X → X satisfies the Distributional Chaos Criterion (DCC) if there exist sequences (xm)m and (ym)m in X \ {0} with ym ∈ span{xk ; k ∈ N} satisfying T nx = 0, x ∈ X0, and (a) lim n→∞ (b) there is a sequence of positive integers (Nm)m increasing to ∞, such that lim m→∞ 1 Nm card{0 ≤ i < Nm ; kT iymk ≥ mkymk} = 1 . Observe that our criterion has, a priori, weaker requirements than the criterion of Cao, Cui and Hou in Theorem 17. We will show that they are actually equivalent. Theorem 19. Let T : X → X be an operator. The following properties are equivalent: (i) T satisfies the hypothesis of Theorem 17. (ii) T satisfies the Distributional Chaos Criterion. Proof. We only need to prove that (ii) implies (i). By (ii) there exists (xm)m ⊂ span(X0) \ {0} that satisfies the condition (a) of Theorem 17 by linearity, and condition (b) by density. Under the assumptions of the DCC we can ensure the existence of distributionally irregular vectors. Proposition 20. If T : X → X satisfies the DCC, then T admits a distributionally irregular vector. Proof. By Theorem 19, passing to subsequences if necessary, we find a sequence (xm)m of nor- malized vectors in X and an increasing sequence (Nm)m of integers with Nm − Nm−1 tending to infinity such that 1 Nm 1 Nm card{0 ≤ i < Nm : kT ixmk > mkT kNm−1 } > 1 − 1 m , card{0 ≤ i < Nm : kT ixkk < 1 m } > 1 − 1 m2 , k = 1, . . . , m − 1, (4) (5) where N0 := 1. Since our hypothesis obviously imply that kT k > 1, the series x := Xk 1 kT kN2k−1 x2k is convergent, thus x ∈ X. We will show that x is a distributionally irregular vector for T . Indeed, kT ixk ≥ 1 kT kN2m−1 1 > 2m − kT ix2mk −Xk,m 2m Xk<m kT kN2k−1 1 1 kT kN2k−1 kT ix2kk 1 kT kN2k−1 −N2m −→ m→∞ ∞ −Xk>m 8 if i < N2m, kT ix2mk > 2mkT kN2m−1, and kT ix2kk < 1 2m , k < m. Conditions (4) and (5) above imply card{0 ≤ i < N2m ; kT ix2mk > 2mkT kN2m−1 , kT ix2kk < 1 2m , k < m} N2m > m − 1 m , and we obtain an increasing sequence of integers B = (mk)k such that dens(B) = 1 and limk→∞kT mk xk = ∞. On the other hand, kT ixk ≤ Xk 1 kT kN2k−1 kT ix2kk < 1 2m + 1 Xk≤m 1 kT kN2k−1 +Xk>m 1 kT kN2k−1−N2m 0, −→ m→∞ if i < N2m+1, and kT ix2kk < 1 2m+1 , k ≤ m. Condition (5) above implies that 1 N2m+1 card{0 ≤ i < N2m+1 ; kT ix2kk < 1 2m + 1 , k = 1, . . . , m} > 1 − 1 m , which gives an increasing sequence A = (nk)k of integers such that dens(A) = 1 and limkkT nk xk = 0, concluding the result. Observe that Propositions 16 and 20 provide an alternative proof of Theorem 17. Remark 21. Not every distributionally chaotic operator satisfies the Distributional Chaos Crite- rion. Indeed, let us consider a weighted forward shift Fw : ℓ2 → ℓ2, defined as (x1, x2, . . . ) 7→ (0, w1x1, w2x2, . . . ), where the sequence of weights w = (wk)k consists, alternatively, of suffi- ciently large blocks of 2's and blocks of 1/2's such that the vector e1 = (1, 0, . . . ) is a distribu- tionally irregular vector. On the other hand, if limk Fk wx = 0 then x = 0. If we impose that the orbits converge to 0 on a dense subset, then the DCC can be character- ized in terms of the existence of distributionally irregular vectors. Corollary 22. Let T : X → X be an operator such that there exists a dense set D with lim n→∞ 0, for all x ∈ D. The following properties are equivalent: T nx = (i) T satisfies the Distributional Chaos Criterion. (ii) T admits a distributionally irregular vector. (iii) There exist a sequence (ym)m in X \{0} and a sequence of positive integers (Nm)m increasing to ∞, such that lim m→∞ 1 Nm card{0 ≤ i < Nm ; kT iymk ≥ mkymk} = 1. We can get stronger results under the assumptions above, which depend on the existence of large scrambled sets consisting of (distributionally) irregular vectors. Definition 23. A linear manifold Y ⊂ X is a (distributionally) irregular manifold for T : X → X if every non-zero vector y ∈ Y \ {0} is a (distributionally) irregular vector for T . Certainly, a (distributionally) irregular manifold is a (distributionally) scrambled set. Theorem 24. Let T : X → X be an operator. If there exists a dense subset X0 ⊂ X such that limn→∞ T nx = 0, for each x ∈ X0 and T admits a distributionally irregular vector, then T is admits a dense distributionally irregular manifold. 9 Proof. We consider a dense sequence (yn)n in X0. By Proposition 20 there is a distributionally irregular vector x ∈ X for T which can be written as a series x = ∞ Xk=1 xk with the properties specified in the proof. We select a countable collection (γm)m = ((γm,k)k)m of sequences of 0's and 1's such that each sequence γm contains an infinite number of 1's and γm,k = 1 =⇒ γn,k = 0, ∀m , n, ∀k ∈ N. We set the sequence of vectors um = Pk γm,k xk, m ∈ N, which, by following the proof of Propo- sition 20, are distributionally irregular vectors since each γm contains an infinite number of 1's. Define now zm = ym + um, m ∈ N. 1 m Since (yn)n is dense in X and the um's are uniformly bounded, we get that the sequence (zm)m is dense in X. We set Y = span{zm ; m ∈ N}, which is a dense subspace of X. If y ∈ Y \ {0}, then we can write y = y0 +Xk ρk xk, where y0 ∈ X0 and the sequence of scalars ρ = (ρk)k only takes a finite number of values, being non-zero an infinite number of ρk's. This means, by following again the proof of Proposition 20, that v := Xk ρk xk is a distributionally irregular vector for T . Since y = y0 + v and limk T ky0 = 0, we get that y is also a distributionally irregular vector for T , and Y is therefore a distributionally scrambled set for T . An adaptation of the above argument, taking into account the proof Theorem 8, yields a useful sufficient condition for dense Li-Yorke chaos. Theorem 25. Let T : X → X be an operator. If supnkT nk = ∞ and there is a dense subset X0 ⊂ X such that limk→∞ T nx = 0 for each x ∈ X0, then T admits a dense irregular manifold. A class of operators for which Theorems 24 and 25 are particularly interesting and easy to apply is the class of operators with dense generalized kernel. Corollary 26. Let T : X → X be an operator whose generalized kernelSn ker T n is dense in X. (i) If T has a distributionally irregular vector, then T admits a dense distributionally irregular manifold. (ii) If supnkT nk = ∞, then T has a dense irregular manifold. To apply Corollary 26 to an important class of operators, we will consider weighted backward shifts Bw : X → X, (x0, x1, . . . ) 7→ (w1x1, w2x2, . . . ) on X = ℓp, 1 ≤ p < ∞, or X = c0, where 10 w = (wi)i is a bounded sequence of non-zero weights, and the (unweighted) backward shift B : X → X, (x0, x1, . . . ) 7→ (x1, x2, . . . ) on the weighted spaces X = ℓp(v) :=  x = (xi)i ; kxk :=  Xi vi xip X = c0(v) := {x = (xi)i ; lim i 1/p < ∞ i vixi = 0, kxk := sup vi xi}, , 1 ≤ p < ∞, or where v = (vi)i is a sequence of strictly positive weights such that supi vi/vi+1 < ∞ in order to have a well-defined backward shift bounded operator. Proposition 27. The following assertions hold: (i) B is Li-Yorke chaotic if and only if Mv := sup{vn/vm; n ∈ N, m > n} = ∞. In this case, B admits a dense irregular manifold. case, Bw admits a dense irregular manifold. (ii) Bw is Li-Yorke chaotic if and only if Mw := supnQm k=n wk ; n ∈ N, m > no = ∞. In this Proof. Suppose that B : ℓp(v) → ℓp(v) is Li-Yorke chaotic, and let x = (xi)i be an irregular vector for B. Since ∞ kBnxkp = Xk=0 vk xk+np = ∞ Xk=0 vk vk+n! vk+n xk+np ≤ Mvkxkp, then Mv < ∞ implies that supnkBnxk < ∞, which is a contradiction. Conversely, if Mv = ∞, we find increasing sequences of integers (nk)k, (mk)k, nk < mk, k ∈ N, such that limk(mk − nk) = ∞ and vnk /vmk > 3k, k ∈ N, by the definition of Mv and the condition on v for continuity of B. We define the vector x = (xi)i ∈ ℓp(v) by xi = (2kvmk )−1/p if i = mk, and xi = 0 otherwise. We have kBmk−nk xk ≥ vnk (cid:12)(cid:12)(cid:12) xmk(cid:12)(cid:12)(cid:12) p = vnk vmk! 1 2!k 2k > 3 −→ k→∞ +∞. This shows that B is Li-Yorke chaotic, and it admits a dense linear manifold Y ⊂ X such that Y is a scrambled set for B by Corollary 26. The case X = c0 is analogous. (ii) is a consequence of (i) if we proceed by conjugation with a suitable diagonal operator (see, e.g., [21]). Another consequence of Theorem 24 is the following easy sufficient condition for dense distributional chaos. Corollary 28. If an operator T : X → X satisfies that 1. there exist an increasing sequence of integers B = (mk)k with dens(B) = 1, y ∈ X satisfying limk→∞kT mk yk = ∞, and 2. a dense subset X0 ⊂ X such that limn→∞ T nx = 0, for each x ∈ X0, then T admits a dense distributionally irregular manifold. A this point we want to recall recent result of Muller and Vrsovsky. Theorem 29 ([24]). Let Tn : X → X, n ∈ N, be a sequence of operators. Suppose that one of the following conditions is satisfied: 11 n=1 1 kTnk < ∞; (i) either P∞ (ii) or X is a complex Hilbert space and P∞ 1 kTnk2 < ∞. Then there exists a point y ∈ X such that limn→∞kTnyk = ∞. n=1 Corollary 28 and Theorem 29 combine nicely to obtain a powerful sufficient condition for dense distributional chaos. Corollary 30. Let T : X → X be an operator such that there exist a dense subset X0 ⊂ X with limn→∞ T nx = 0, for each x ∈ X0, and an increasing sequence of integers B = (mk)k with dens(B) = 1 satisfying that k=1 1 kT mk k < ∞; (i) either P∞ (ii) or X is a complex Hilbert space and P∞ Then T has a dense distributionally irregular manifold. 1 kT mk k2 < ∞. k=1 As consequence of Corollary 30, we obtain the following property. Corollary 31. Let T : X → X be an operator. If there exist a dense set X0 such that limn→∞ T nx = 0, for each x ∈ X0 and r(T ) > 1, then T admits a dense distributionally irregular manifold. Definition 32 ([10]). Let H be a complex separable Hilbert space. For Ω a connected open subset of C and n a positive integer, let Bn(Ω) denote the set of operators T in H which satisfy (a) Ω ⊂ σ(T ), (b) rank(T − wI) = H for w ∈ Ω, (c) W kerw∈Ω(T − wI) = H, (d) dim ker(T − wI) = n for w ∈ Ω. Corollary 33. Let T ∈ Bn(Ω). If Ω ∩ T , ∅, then T has a dense distributionally irregular manifold. Proof. Let U be a connected open subset of Ω ∩ D. Then X0 := spanSw∈U ker(T − wI) is dense in H and satisfies that limn→∞ kT nxk = 0, for all x ∈ X0 and r(T ) > 1. Hence the conclusion follows by an application of Corollary 31. 5. Existence of distributionally chaotic operators The study of general conditions under which a space X admits operators with certain wild be- haviour has attracted the interest of many researchers. Ansari [1] and Bernal-Gonz´alez [6] inde- pendently proved that any separable infinite dimensional Banach space X supports a hypercyclic operator. This result was extended to Fr´echet spaces by Bonet and Peris [8]. A generalization for polynomials is given in [23]. Other existence results for related properties and for C0-semigroups can be found, for instance, in [4, 7, 5, 29, 30]. In this section we want to establish the existence of distributionally chaotic operators on arbitrary infinite dimensional and separable Banach spaces. The following Lemma of Shkarin [29] and its consequences will be key to obtain our existence result. 12 Lemma 34 ([29]). Let X be a separable infinite dimensional Banach space, let (bk)k be a se- quence of numbers in [3, +∞[ such that bk → ∞ as k → ∞ and (Nk)k be a strictly increasing sequence of positive integers such that N0 = 0 and Nk+1 − Nk ≥ 2 for each k ∈ N. Then there exists a biorthogonal sequence ((yk, fk))k in X × X∗ such that (B1) kykk = 1 for each k ∈ N; (B2) span{yk : k ∈ N} is dense in X; (B3) k fnk k ≤ bk for each k ∈ N; (B4) k f jk ≤ 3 if j ∈ N \ {Nk : k ∈ N}; (B5) for any k ∈ Z and any c j ∈ K with Nk + 1 ≤ j ≤ Nk+1 − 1 1 2 k Nk+1 −1 Xj=Nk+1 Nk+1 −1 Xj=Nk+1 c jy jk ≤  c j2 1 2 ≤ 2k Nk+1−1 Xj=Nk+1 c jy jk In Section 3 we showed that there are no operators satisfying the Strong Distributional Chaos Criterion on certain Banach spaces. In the following result we give a positive answer for distri- butionally chaotic operators. Theorem 35. In every infinite dimensional separable Banach space there exists a hypercyclic and distributionally chaotic operator which admits a dense distributionally irregular manifold. Proof. By [29, Lemma 2.5], given a sequence (wn)n of positive weights in ℓ2 there exists T : X → X satisfying that T y0 = 0 and T yn+1 = wnyn with kynk = 1 where (yn)n is given by Lemma 34. We need to compute k(I + T )iy jk, that is k(I + T )iy jk = k i Xk=0 i = ky j + i k!T ky jk 1!T y j + · · · + i i!T iy jk = ky j + iw j−1y j−1 + · · · + w j−1 · · · w j−iy j−ik If j = Nk+1 − 1 and j − i ≥ Nk + 1, that is, i ≤ Nk+1 − Nk − 2, then k(I + T )iy jk ≥ 1 2 iw j−1 by condition (B5) in Lemma 34. Taking wk = k− 2 for i = 3k([((k + 2)!) 2 3 ] + 1), · · · , (k + 2)! − (k + 1)! − 2 3 , N0 := 0 and Nk = (k + 1)! + 1 for k ≥ 1, then k(I + T )iy jk ≥ 1 2 iw j−1 ≥ 1 2 3k([((k + 2)!) 2 3 ] + 1)) ((k + 2)! − 1) 2 3 > k + 1, for j = (k + 2)! = Nk+1 − 1. Thus, for um = yNm+1 −1, m ∈ N, we have 1 Nm card{0 ≤ i < Nm : k(I + T )iumk ≥ mkumk = m} ≥ (m + 1)! − m! − 2 − (3(m − 1)([((m + 1)!) 2 3 ] + 1) (m + 1)! + 1 13 1. −→ m→∞ Moreover, I + T satisfies the Kitai Criterion (therefore, it is hypercyclic) [29, Theorem 1.1 and Corollary 1.3], thus there exists a dense sequence (xk)k in X such that (I + T )nxk → 0 as n → ∞. Hence the operator I + T satisfies the conditions of Corollary 22 and Theorem 24, and we conclude that I + T has a dense distributionally irregular manifold. 6. Acknowledgements The second author is supported in part by MEC and FEDER, Project MTM2008-05891. The third and fourth authors are supported in part by MEC and FEDER, Projects MTM2007-64222, MTM2010-14909 and by Generalitat Velenciana, Project PROMETEO/2008/101. References [1] S. Ansari, Existence of hypercyclic operators on topological vector spaces, J. Func. Anal. 148 (1997), 384 -- 390. [2] F. Bayart and ´E. Matheron, Dynamics of linear operators, Cambridge University Press, Vol. 179, 2009. [3] B. Beauzamy, Introduction to Operator Theory and Invariant Subspaces, North-Holland, Amsterdam, 1988. [4] T. Berm´udez, A. Bonilla and A. Martin´on, On the existence of chaotic and hypercyclic semigroups on Banach spaces, Proc. Amer. Math. Soc. 131 (2003), 2435 -- 2441. [5] T. Berm´udez, A. Bonilla, J. A. Conejero and A. Peris, Hypercyclic, topologically mixing and chaotic semigroups on Banach spaces, Studia Math. 170 (2005), 57 -- 75. [6] L. Bernal-Gonz´alez, On hypercyclic operators on Banach spaces, Proc. Amer. Math. Soc. 127 (1999), 1003 -- 1010. [7] L. Bernal-Gonz´alez and K.-G. Grosse-Erdmann, Existence and nonexistence of hypercyclic semigroups, Proc. Amer. Math. Soc. 135 (2007), 755 -- 766. [8] J. Bonet, A. Peris, Hypercyclic operators on non-normable Fr´echet spaces, J. Funct. Anal. 159 (1998), 587 -- 595. [9] J. Bonet, F. Mart´ınez-Gim´enez and A. Peris, A Banach space which admits no chaotic operator, Bull. London Math. Soc. 33 (2001), 196 -- 198. [10] M. J. Cowen and R. G. Douglas, Complex geometry and operator theory, Acta Math., 141 (1) (1978), 187 -- 261. [11] R. L. Devaney, An introduction to chaotic dynamical systems, 2nd Edition, Addison-Wesley Studies in Nonlinear- ity, Redwood City, CA, 1989. [12] J. Duan, X.-C. Fu, P.-D. Liu and A. Manning, A linear chaotic quantum harmonic oscillator, Appl. Math. Lett. 12 (1999), 15 -- 9. [13] N.S. Feldman, T.L. Miller and V.G. Miller, Hypercyclic and supercyclic cohyponormal operators, Acta Sci. Math.(Szeged), 68 (2002), 303 -- 328. [14] X.-C. Fu and J. Duan, Infinite-dimensional linear dynamical systems with chaoticity, J. Nonlinear Sci. 9 (1999), 197 -- 211. [15] W. T. Gowers and B. Maurey, The unconditional basic sequence problem, J. Amer. Math. Soc., 6 (1993), 851 -- 874. [16] K. G. Grosse-Erdmann and A. Peris, Linear Chaos, Universitext, Springer, to appear. [17] B. Hou, P. Cui and Y. Cao, Chaos for Cowen-Douglas operators, Proc. Amer. Math. Soc, 138 (2010), 929 -- 936. [18] B. Hou, G. Tian and L. Shi, Some dynamical properties for linear operators, preprint (arXiv:0903.4558v1). [19] B. Hou, L. Shi, G. Tian and S. Zuh, Approximation of chaotic operators, J. Oper. Theory, to appear. [20] T.Y. Li and J.A. Yorke, Period three implies chaos, Amer. Math. Monthly 82 (1975), 985 -- 992. [21] F. Mart´ınez-Gim´enez and A. Peris, Chaos for backward shift operators, Int. J. of Bifurcation and Chaos 12 (2002), 1703 -- 1715. [22] F. Mart´ınez-Gim´enez, P. Oprocha and A. Peris, Distributional chaos for backward shifts, J. Math. Anal. Appl., 351 (2009), 607 -- 615. [23] F. Mart´ınez-Gim´enez, A. Peris, Existence of hypercyclic polynomials on complex Fr´echet spaces, Top. Appl. 156 (2009), 3007 -- 3010. [24] V. Muller and J. Vrsovsky, Orbits of linear operators tending to infinity, Rocky Mountain J. Math. 39 (2009), 219 -- 230. [25] P. Oprocha, A quantum harmonic oscillator and strong chaos, J. Phys. A, 39 (2006), 14559 -- 14565. [26] P. Oprocha, Distributional chaos revisited, Trans. Amer. Math. Soc., 361 (2009), 4901 -- 4925. [27] G.T. Prajitura, Irregular vectors of Hilbert space operators, J. Math. Anal. Appl. 354 (2009), 689 -- 697. [28] B. Schweizer and J. Sm´ıtal, Measures of chaos and a spectral decomposition of dynamical systems on the interval, Trans. Amer. Math. Soc. 344 (1994), 737 -- 754. [29] S. Shkarin, The Kitai criterion and backward shifts, Proc Amer. Math. Soc. 136 (2008), 1659 -- 1670. [30] S. Shkarin, A short proof of existence of disjoint hypercyclic operators, J. Math. Anal. Appl., 367 (2010), 713 -- 715. 14
1508.06102
2
1508
2016-01-15T12:51:56
An L^1-estimate for certain spectral multipliers associated with the Ornstein--Uhlenbeck operator
[ "math.FA" ]
We study a class of spectral multipliers \phi(L) for the Ornstein--Uhlenbeck operator L arising from the Gaussian measure on R^n and find a sufficient condition for integrability of \phi(L)f in terms of the admissible conical square function and a maximal function.
math.FA
math
An L1-estimate for certain spectral multipliers associated with the Ornstein -- Uhlenbeck operator Mikko Kemppainen Abstract. We study a class of spectral multipliers φ(L) for the Ornstein -- Uhlenbeck operator L arising from the Gaussian measure on Rn and find a sufficient condition for integrability of φ(L)f in terms of the admissible conical square function and a maximal function. On the Euclidean space Rn, the Ornstein -- Uhlenbeck operator 1. Introduction is associated with the Gaussian measure by the Dirichlet form(cid:90) (Lf )g dγ = Rn ∇f · ∇g dγ, f, g ∈ C∞ 0 (Rn). It has discrete spectrum σ(L) = {0, 1, 2, . . .} on L2(γ), and an orthonormal basis of eigen- functions is given by Hermite polynomials hβ, β ∈ Nn, so that Lhβ = βhβ. Moreover, it generates a (positive) diffusion semigroup on L2(γ) which can be expressed as L = − 1 2 ∆ + x · ∇ dγ(x) = π−n/2e−x2 dx (cid:90) 1 2 Rn Mt(x, y)f (y) dγ(y), (cid:90) 1 − e−2tx − y2(cid:17) (cid:16) − e−t Rn e−tLf (x) = by means of the Mehler kernel (see [12]) Mt(x, y) = (1 − e−2t)n/2 exp 1 t > 0, (cid:16) e−t (cid:17) 1 + e−t (x2 + y2) . exp The semigroup is contractive on Lp(γ) for each 1 ≤ p ≤ ∞, and acts conservatively so that e−tL1 = 1. Therefore, E. M. Stein's theory [13] is applicable in studying the boundedness of spectral multipliers φ(L) defined as φ(L)hβ = φ(β)hβ for β ∈ Nn. More precisely, [13, Corollary 3, p. 121] guarantees Lp(γ)-boundedness with 1 < p < ∞, for any spectral multiplier of 'Laplace transform type', i.e. of the form (cid:90) ∞ φ(λ) = λ Φ(t)e−tλ dt, λ ≥ 0, where Φ : (0,∞) → C is bounded. In particular, the imaginary powers Liτ , τ ∈ R, arising from Φ(t) = t−iτ /Γ(1 − iτ ) are bounded on Lp(γ). The general theory was improved by 0 2010 Mathematics Subject Classification. 42B25 (Primary); 42B15, 42B30 (Secondary). Key words and phrases. conical square function, admissibility function, Mehler kernel, Gaussian measure. 1 AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 2 M. Cowling [3] who showed that, for a given p ∈ (1,∞), φ(L) is bounded on Lp(γ) as soon as φ extends analytically to a sector of angle greater than π1/p− 1/2. (See also the more recent development [2].) The L1-theory in the Gaussian setting is quite problematic. Although finite linear combinations of Hermite polynomials are dense in L1(γ), the spectral projections onto their eigenspaces are not L1(γ)-bounded. Moreover, tLe−tL is bounded (uniformly) on Lp(γ) only when 1 < p < ∞ (see [7, Chapter 5]). A Gaussian weak (1, 1)-type estimate for spectral multipliers of Laplace transform type was established by J. García-Cuerva et al. [5]. Moreover, in [4] they showed that requiring analyticity of φ in a sector of angle smaller than arcsin1 − 2/p will not alone suffice for boundedness of φ(L) on Lp(γ). Observing that arcsin1 − 2/p → π/2 as p → 1 is in line with the fact that the spectrum of L on L1(γ) is the (closed) right half-plane. Furthermore, L1(γ)-boundedness of dilation invariant spectral multipliers for L was characterised in [6, Theorem 3.5 (ii)]. The main obstruction in developing a metric theory of Hardy spaces in the Gaussian setting arises from the fact that the rapidly decaying measure γ is non-doubling, that is, for every t > 0 γ(B(x, 2t)) γ(B(x, t)) −→ ∞, as x → ∞. G. Mauceri and S. Meda overcame this problem in [10] and developed an atomic theory for a Gaussian Hardy space which relies of the fact that the Gaussian measure behaves well locally with respect to the admissibility function m(x) = min(1,x−1), x ∈ Rn. Indeed, γ is doubling on families of 'admissible' Euclidean balls Bα = {B(x, t) : 0 < t ≤ αm(x)}, α ≥ 1, in the sense that for all λ ≥ 2 we have (1) γ(λB) ≤ e4λ2α2 γ(B), B ∈ Bα. Other natural objects that are suitable for defining Hardy spaces, namely maximal functions and square functions, were studied in the Gaussian setting by J. Maas, J. van Neerven, and P. Portal. In [8] they considered (a version1 of) the admissible conical square function Sf (x) = 0 γ(B(x, t)) B(x,t) t2Le−t2Lf (y)2 dγ(y) dt t x ∈ Rn, , and showed that it is controlled by a non-tangential semigroup maximal function. The converse inequality was presented in [11] along with a proof that the Riesz transform satisfies (cid:107)∇L−1/2f(cid:107)1 (cid:46) (cid:107)Sf(cid:107)1 + (cid:107)f(cid:107)1. The benefit of conical objects (as opposed to vertical ones) is the applicability of tent space theory, which in the Gaussian setting was initiated in [9] and further developed by A. Amenta and the author in [1]. The aim of this paper is to examine the decomposition method presented in [11] and to see what kind of L1-estimates one can obtain for spectral multipliers φ(L)f in terms of the admissible conical square function Sf and other relevant objects. The hope is that these considerations will help in developing a fully satisfactory theory of Gaussian Hardy spaces. (cid:16)(cid:90) 2m(x) (cid:90) 1 (cid:17)1/2 (cid:90) ∞ Theorem. Let φ(λ) = 0 1They have t∇e−t2L instead of t2Le−t2L. Φ(t)(tλ)2e−tλ dt t , λ ≥ 0, AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 3 where Φ : (0,∞) → C is twice continuously differentiable and satisfies (Φ(t) + tΦ(cid:48)(t) + t2Φ(cid:48)(cid:48)(t)) + sup 0<t<∞ Then, for all f ∈ L1(γ), we have (Φ(cid:48)(t) + tΦ(cid:48)(cid:48)(t)) dt < ∞. (cid:90) ∞ 1 where (cid:107)φ(L)f(cid:107)1 (cid:46) (cid:107)Sf(cid:107)1 + (cid:107)f(cid:107)1 + (cid:107)(1 + log+ · ) M f(cid:107)1, M f (x) = sup εm(x)2<t≤1 e−tLf (x), x ∈ Rn, and ε > 0 does not depend on f. Remarks. Several remarks are in order: (1) The term (cid:107)(1 + log+ · ) M f(cid:107)1 is highly undesirable for two reasons. Firstly, the maximal operator M is of a non-admissible kind in the sense that it is not restricted to times t (cid:46) m(·). Secondly, the weight factor (1+log+ ·), which arises from the admissibility function m, seems problematic. However, it is difficult to see how the appearance of the term could be avoided. Notice, nevertheless, that (cid:107)(1 + log+ · ) M f(cid:107)1 is finite at least if f ∈ Lp(γ) with 1 < p < ∞. (2) The operators in the theorem above are special kind of Laplace type multipliers; φ(λ) = Φ(t)(tλ)2e−tλ dt t = λ (Φ(t) + tΦ(cid:48)(t))e−tλ dt, λ ≥ 0, and therefore bounded on Lp(γ) when 1 < p < ∞. Note that if, in addition, we had (cid:90) 1 0 (Φ(cid:48)(t) + tΦ(cid:48)(cid:48)(t)) dt < ∞, then φ(L) would be bounded even on L1(γ). Indeed, using integration by parts we have φ(L)f = −Φ(0)f + (2Φ(cid:48)(t) + tΦ(cid:48)(cid:48)(t))e−tLf dt so that (cid:107)e−tLf(cid:107)1 ≤ (cid:107)f(cid:107)1 implies (cid:107)φ(L)f(cid:107)1 (cid:46)(cid:16)Φ(0) + (Φ(cid:48)(t) + tΦ(cid:48)(cid:48)(t)) dt (cid:17)(cid:107)f(cid:107)1. 0 (3) An example of a multiplier satisfying the conditions of the theorem is the localized imaginary power arising from Φ(t) = tiτ χ(t), where τ ∈ R and χ is a smooth cutoff with, say, 1(0,1] ≤ χ ≤ 1(0,2]. Observe that for 0 < t ≤ 1 we have Φ(cid:48)(t) (cid:104) t−1 and Φ(cid:48)(cid:48)(t) (cid:104) t−2 so that(cid:90) 1 (Φ(cid:48)(t) + tΦ(cid:48)(cid:48)(t)) dt = ∞. 0 Acknowledgements. The research has been supported by the Academy of Finland via the Centre of Excellence in Analysis and Dynamics Research (project No. 271983). The author wishes to thank Alex Amenta and Jonas Teuwen for enlightening discussions. (cid:90) ∞ 0 (cid:90) ∞ 0 (cid:90) ∞ (cid:90) ∞ 0 AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 4 2. Proof of the theorem (cid:40) (cid:101)m(x) = Strategy. The proof of the theorem follows the decomposition method from [11]. Let us begin by introducing a discretized version of the admissibility function x < 1, 1, 2−k, 2k−1 ≤ x < 2k, k ≥ 1, and write (cid:101)Bα for the associated family of admissible balls. From (cid:101)m ≤ m ≤ 2(cid:101)m it follows that (cid:101)Bα ⊂ Bα ⊂ (cid:101)B2α. This discretization is relevant for Proposition 7. (cid:101)m(y)} for which We define the Gaussian tent space adapted to this new admissibility function as the space t1(γ) of functions u on the admissible region D = {(y, t) ∈ Rn × (0,∞) : 0 < t < (cid:107)u(cid:107)t1(γ) = u(y, t)2 dγ(y) dt tγ(B(y, t)) dγ(x) < ∞. (cid:90) (cid:16)(cid:90)(cid:90) Rn Γ(x) (cid:17)1/2 Here Γ(x) = {(y, t) ∈ D : y − x < t} is an admissible cone at x ∈ Rn. The main theorem of [1] guarantees that every u ∈ t1(γ) admits a decomposition into 'atoms' ak so that Recall that atom is a function a on D associated with a ball B ∈ (cid:101)B5 for which supp a ⊂ k u = λk (cid:104) (cid:107)u(cid:107)t1(γ). (cid:88) λkak, with (cid:88) (cid:16)(cid:90) rB k (cid:107)a(·, t)(cid:107)2 2 dt t 0 (cid:17)1/2 ≤ γ(B)−1/2. B × (0, rB) and Let then φ and Φ be as in Theorem and let f be a polynomial. For any δ, δ(cid:48) > 0 and φ(L)f = cδ,δ(cid:48) Φ((δ(cid:48) + δ)t2)(t2L)2e−(δ(cid:48)+δ)t2Lf For such a function, (cid:107)a(cid:107)t1(γ) (cid:46) 1. (cid:90) ∞ κ ≥ 1 we can decompose φ(L)f into three parts as follows: (cid:16)(cid:90) (cid:101)m(·)/κ (cid:90) (cid:101)m(·)/κ (cid:90) ∞ (cid:101)m(·)/κ (cid:101)Φ(t2)t2Le−δ(cid:48)t2Lu(·, t) (cid:101)Φ(t2)t2Le−δ(cid:48)t2L(1Dc(·, t)t2Le−δt2Lf ) (cid:101)Φ(t2)(t2L)2e−(δ(cid:48)+δ)t2Lf where u(·, t) = 1D(·, t)t2Le−δt2Lf and (cid:101)Φ(t) = Φ((δ(cid:48) + δ)t). =: cδ,δ(cid:48)(π1u + π2f + π3f ), = cδ,δ(cid:48) (cid:17) dt t + + dt t dt t 0 0 0 dt t Now (cid:107)φ(L)f(cid:107)1 ≤ cδ,δ(cid:48)((cid:107)π1u(cid:107)1 + (cid:107)π2f(cid:107)1 + (cid:107)π3f(cid:107)1), and the proof consists of estimating these three terms separately for sufficiently small δ > δ(cid:48) > 0 and large enough κ ≥ 1. Analysis of the three parts. Proposition 2 deals with (cid:90) (cid:101)m(·)/κ π1u = 0 (cid:101)Φ(t2)t2Le−δ(cid:48)t2Lu(·, t) dt t and relies on the following L2-L2 -off diagonal estimate (cf. [11, Proposition 4.2] and [14]). AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 5 Lemma 1. There exists a constant cod > 0 such that for j = 0, 1 we have (cid:16) − d(E, E(cid:48))2 (cid:17) codt , t > 0, (cid:107)1E(cid:48)(tL)je−tL1E(cid:107)2→2 (cid:46) exp whenever E, E(cid:48) ⊂ Rn. Proposition 2. Let κ ≥ 1 and 0 < δ ≤ 1. For sufficiently small δ(cid:48) > 0 we have (cid:107)π1u(cid:107)1 (cid:46) (cid:107)u(cid:107)t1(γ). Moreover, the function u(·, t) = 1D(·, t)t2Le−δt2Lf satisfies (cid:107)u(cid:107)t1(γ) (cid:46) (cid:107)Sf(cid:107)1. atom a associated with a ball B ∈ (cid:101)B5. Let us consider the annuli Ck(B) = 2k+1B \ 2kB Proof. By the atomic decomposition, it suffices to show that (cid:107)π1a(cid:107)1 (cid:46) 1 for any for k ≥ 1, and C0(B) = 2B. By Hölder's inequality we have (cid:107)π1a(cid:107)1 = (2) We estimate the norms on the right hand side of (2) by pairing with a g ∈ L2(γ) and relying on the assumption that Φ is bounded: (cid:101)Φ(t2)t2Le−δ(cid:48)t2La(·, t) (cid:13)(cid:13)(cid:13)2 . dt t ≤ 0 0 k=0 dt t (cid:101)Φ(t2)t2Le−δ(cid:48)t2La(·, t) (cid:13)(cid:13)(cid:13)1 (cid:90) rB∧2−k−1/κ (cid:13)(cid:13)(cid:13)(cid:90) (cid:101)m(·)/κ γ(2k+1B)1/2(cid:13)(cid:13)(cid:13)1Ck(B) ∞(cid:88) (cid:90) rB∧2−k−1/κ (cid:12)(cid:12)(cid:12)(cid:90) (cid:101)Φ(t2)t2Le−δ(cid:48)t2La(·, t) (cid:12)(cid:12)(cid:12)(cid:90) rB∧2−k−1/κ (cid:90) a(·, t)(cid:101)Φ(t2)t2Le−δ(cid:48)t2Lg dγ (cid:17)1/2(cid:16)(cid:90) rB (cid:46)(cid:16)(cid:90) rB (cid:16) (cid:88) (cid:90) rB (cid:17)1/2 (cid:107)a(·, t)(cid:107)2 B dt t dt t Rn = 0 0 2 0 0 g dγ (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) dt t (cid:107)1Bt2Le−δ(cid:48)t2Lg(cid:107)2 2 (cid:17)1/2 . dt t (cid:107)1Bt2Le−δ(cid:48)t2Lg(cid:107)2 2 dt t (cid:104)g, hβ(cid:105)2 (t2β)2e−2δ(cid:48)t2β dt t 0 = β∈Nn (cid:46) (cid:107)g(cid:107)2. (cid:17)1/2 (cid:16)(cid:90) rB 0 Now, for g supported in C0(B) = 2B we have When k ≥ 1 we have d(Ck(B), B) ≥ (2k − 1)rB ≥ 2k−1rB and so, by Lemma 1, it follows that for 0 < t ≤ rB, (cid:107)1Bt2Le−δ(cid:48)t2L1Ck(B)(cid:107)2→2 (cid:46) exp Hence, for g supported in Ck(B), k ≥ 1, we have dt t We have therefore shown that, for k ≥ 0, (cid:107)1Bt2Le−δ(cid:48)t2Lg(cid:107)2 0 (cid:17)1/2 . rB 2 B codδ(cid:48) codδ(cid:48)t2 (cid:17)(cid:16) t (cid:16) − 4k−2 (cid:17) (cid:46) exp (cid:16) − 4k−1r2 (cid:17)(cid:107)g(cid:107)2. (cid:16) − 4k−2 (cid:17)1/2 (cid:46) exp (cid:13)(cid:13)(cid:13)2 (cid:16) − 4k−2 (cid:17) (cid:16) ∞(cid:88) (cid:46) exp codδ(cid:48) codδ(cid:48) dt t γ(B)−1/2 (cid:46) exp γ(B)−1/2. 50 · 4k+1 − 4k−2 codδ(cid:48) k=0 (cid:17) (cid:46) 1 (cid:101)Φ(t2)t2Le−δ(cid:48)t2La(·, t) (cid:17) (cid:16) − 4k−2 codδ(cid:48) (cid:16)(cid:90) rB (cid:90) rB (cid:13)(cid:13)(cid:13)1Ck(B) ∞(cid:88) 0 k=0 According to the doubling inequality (1), we have γ(2k+1B)1/2 (cid:46) e2·4k+1·25γ(B)1/2 and therefore (cid:107)π1a(cid:107)1 (cid:46) γ(2k+1B)1/2 exp as soon as δ(cid:48) < 1/(3200cod). This proves the first claim. 6 For the second claim, let u(·, t) = 1D(·, t)t2Le−δt2Lf. We perform a change of variable, AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS δt2 = s2, i.e. t = s/ δ so that (y, t) ∈ Γ(x) ⇔ y − x < t < (cid:101)m(y) √ ⇔ y − x < s/ ⇔ (y, s) ∈ Γ(cid:48) √ √ where D(cid:48) := {(y, s) ∈ Rn×(0,∞) : s < tent space on D(cid:48) (see [1, Corollary 3.5]) guarantees that t2Le−δt2Lf (y)2 dγ(y) dt tγ(B(y, t)) (cid:107)u(cid:107)t1(γ) = 1/ δ Rn δ < (cid:101)m(y) δ(cid:101)m(y)}. Now, change of aperture in the Gaussian √ (x) := {(y, s) ∈ D(cid:48) : y − x < s/ δ}, (cid:17)1/2 dγ(x) (cid:17)1/2 dγ(x) δ−1s2Le−s2Lf (y)2 s2Le−s2Lf (y)2 dγ(y) ds sγ(B(y, s)) √ dγ(y) ds sγ(B(y, s/ δ)) (cid:17)1/2 dγ(x). √ (cid:90) (cid:90) (cid:90) = (cid:46) Rn Rn 1/ (x) Γ(cid:48) Γ(x) √ δ (cid:16)(cid:90)(cid:90) (cid:16)(cid:90)(cid:90) (cid:16)(cid:90)(cid:90) Γ(cid:48)(x) ⊂ Γ(x) ⊂ (cid:91) (cid:90) 2m(x) Γ(cid:48)(x) We then observe (see [9, Lemma 2.3]) that for any x, y ∈ Rn, y − x < s < m(y) implies s < 2m(x), and therefore B(x, s) × {s}. 0<s<2m(x) Moreover, γ(B(y, s)) (cid:104) γ(B(x, s)) when y − x < s < δ(cid:101)m(y), and hence (cid:90)(cid:90) for every x ∈ Rn, which shows that (cid:107)u(cid:107)t1(γ) (cid:46) (cid:107)Sf(cid:107)1 as required. ds s (cid:3) For π2 and π3 (more precisely, for Proposition 5 and Lemma 6) we need the following s2Le−s2Lf (y)2 dγ(y) ds sγ(B(y, s)) s2Le−s2Lf (y)2 dγ(y) γ(B(x, s)) Γ(cid:48)(x) (cid:90) B(x,s) (cid:46) 1 0 two lemmas concerning pointwise kernel estimates. Lemma 3. Let j = 0, 1. For all x, y ∈ Rn we have the pointwise kernel estimate tj∂j t Mt(x, y) (cid:46) t−n/2 exp 8t As a consequence, for all 0 < t ≤ 1 we have (cid:107)1E(cid:48)(tL)je−tL1E(cid:107)1→∞ (cid:46) t−n/2 exp (cid:16) − x − y2 (cid:17) , (cid:17) (cid:16) − d(E, E(cid:48))2 (cid:16)x2 + y2 (cid:17) exp 2 8t exp sup x∈E y∈E(cid:48) 0 < t ≤ 1. (cid:16)x2 + y2 (cid:17) , 2 whenever E, E(cid:48) ⊂ Rn. Proof. For 0 < t ≤ 1 we have the elementary estimates 1 1 − e−2t (cid:104) 1 t , 1 4t and the case j = 0 follows immediately: , 2t 1 8 ≤ e−t 1 − e−2t ≤ 1 (cid:16) − e−t 1 − e−2tx − y2(cid:17) (cid:17) (cid:16)x2 + y2 (cid:17) exp exp 2 ≤ e−t 1 + e−t ≤ 1 (cid:17) (cid:16) e−t 1 + e−t (x2 + y2) (cid:17) e−t . (1 − e−2t)2 − (x2 + y2) (1 + e−t)2 Mt(x, y). 1 (1 − e−2t)n/2 exp (cid:16) − x − y2 Mt(x, y) = (cid:46) t−n/2 exp For j = 1 we calculate: 4t (cid:16) − ne−2t 1 − e−2t + x − y2 e−t(1 + e−2t) 2 ∂tMt(x, y) = AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 7 Using the previous case j = 0 we then see that x − y2 t∂tMt(x, y) (cid:46)(cid:16) (cid:90) E (cid:46) t−n/2 exp 1 + t (cid:46) t−n/2 exp Mt(x, y) (cid:16)x2 + y2 + x2 + y2(cid:17) (cid:17) (cid:16) − x − y2 (cid:16)x2 + y2 (cid:17) (cid:16)x2 + y2 exp exp 8t 2 2 (cid:17)f (y) dγ(y) (cid:17)(cid:90) (cid:17) . exp sup y∈E 2 E (cid:16) − x − y2 (cid:16) − d(E, E(cid:48))2 (cid:17) 8t 8t The consequence is also immediate: for any x ∈ E(cid:48) we have (tL)je−tLf (x) (cid:46) t−n/2 exp f (y) dγ(y). (cid:3) Lemma 4. For α large enough there exists a constant c > 0 such that for all x, y ∈ Rn and all 0 < t ≤ 1 we have Mt/α(x, y) (cid:46) exp and, consequently, (cid:16) (cid:17) (cid:16) − x − y2 (cid:16) ct (cid:17) exp αt min(x2,y2) Mt(x, y), t∂tMt/α(x, y) (cid:46) exp αt min(x2,y2) Mt(x, y). Proof. An alternative way to express the Mehler kernel is (see [12]) (cid:17) (cid:16) − e−tx − y2 1 − e−2t The first claim now follows because for all x, y ∈ Rn and all 0 < t ≤ 1 we have x − y2 (cid:46) max(e−tx − y2,x − e−ty2). In order to see this, let us assume, with no loss of generality, that x ≤ y, and show that x − y2 (cid:46) e−tx − y2. Then x − y2 ≤ e(e−tx2 − 2e−tx · y + e−ty2) = e(e−tx2 − (1 − e−t)y2 − 2e−tx · y + y2), where because x ≤ y. Indeed, e−tx2 − (1 − e−t)y2 ≤ e−2tx2, e−tx2 − (1 − e−t)x2 − e−2tx2 = (2e−t − 1 − e−2t)x2, where 2e−t − 1 − e−2t ≤ 0 for all t > 0. By [11, Lemma 3.4] for α large enough we have for all x, y ∈ Rn and all 0 < t ≤ 1 that exp Therefore Mt(x, y) = (1 − e−2t)n/2 exp 1 (cid:16) − e−t/αx − y2 1 − e−2t/α Mt/α(x, y) (cid:46) exp exp (cid:17) e−tx − y2 1 − e−2t (cid:16) − 2 (cid:17) ≤ exp (cid:16) − e−tx − y2 (cid:16) − max(e−tx − y2,x − e−ty2) (cid:16) t2 min(x2,y2) (cid:17) 1 − e−2t/α 1 − e−2t (cid:17) exp . 1 − e−2t (cid:17) exp where, by symmetry, the first exponential factor can be replaced by ey2 (cid:17) (cid:16) t2 min(x2,y2) . 1 − e−2t/α (cid:17) . Mt(x, y), 8 (cid:3) The second claim now follows from the first one since AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS t∂tMt/α(x, y) (cid:46)(cid:16) + x2 + y2(cid:17) (cid:17) x − y2 t (cid:16) 1 + (cid:46) exp αt min(x2,y2) Mt(x, y). Mt/α(x, y) Here the first inequality is obtained as in the proof of Lemma 3 (case j = 1). Let us then consider π2f = (cid:90) (cid:101)m(·)/κ 0 (cid:101)Φ(t2)t2Le−δ(cid:48)t2L(1Dc(·, t)t2Le−δt2Lf ) dt t . We then decompose π2f (using boundedness of Φ) as follows: Proposition 5. Let κ ≥ 4. For sufficiently small δ > δ(cid:48) > 0 we have (cid:107)π2f(cid:107)1 (cid:46) (cid:107)f(cid:107)1. Proof. We begin by observing that if t ≤ (cid:101)m(x)/4 and t > 2−k−1 for some k ≥ 2, then x < 2k−2. Moreover, if t ≥ (cid:101)m(y) and t ≤ 2−k, then y ≥ 2k−1. (cid:13)(cid:13)(cid:13)(cid:90) (cid:101)m(·)/κ (cid:101)Φ(t2)t2Le−δ(cid:48)t2L(1Dc(·, t)t2Le−δt2Lf ) (cid:90) 2−k ∞(cid:88) (cid:107)1B(0,2k−2)t2Le−δ(cid:48)t2L(1Rn\B(0,2k−1)t2Le−δt2Lf )(cid:107)1 (cid:90) 2−k ∞(cid:88) ∞(cid:88) (cid:107)π2f(cid:107)1 = (cid:107)1B(0,2k−2)t2Le−δ(cid:48)t2L(1Ck+l−1t2Le−δt2Lf )(cid:107)1 (cid:13)(cid:13)(cid:13)1 2−k−1 dt t dt t (3) ≤ (cid:46) k=2 0 dt t , k=2 l=1 2−k−1 where Ck+l−1 := B(0, 2k+l−1) \ B(0, 2k+l−2). First, by Lemma 4, we choose a δ > 0 such that for all 0 < t ≤ 1 we have Then, since the distance between B(0, 2k−2) and Ck+l−1 is at least 2k+l−3, we have, by Lemma 3, for 2−k−1 < t ≤ 2−k that t2Le−δt2Lf (x) (cid:46) exp Hence, for 2−k−1 < t ≤ 2−k we have (cid:107)1Ck+l−1t2Le−δt2Lf(cid:107)1 (cid:46) exp (cid:16) t2x2 (cid:16) 4−k · 4k+l−1 δ δ δ x ∈ Rn. (cid:17)e−tLf (x), (cid:17)(cid:107)e−tLf(cid:107)1 (cid:46) exp (cid:16) 4l−1 (cid:17)(cid:107)f(cid:107)1. (cid:17) (cid:16) 4k−2 + 4k+l−1 (cid:16) − 4k+l−3 (cid:16) − 42k+l−5 + 4k+l−1(cid:17) (cid:17)(cid:107)f(cid:107)1 4l−1 (cid:17)(cid:17)(cid:107)f(cid:107)1 + 4k+l−1 + δ δ(cid:48) − 4−2 − 4−2 8δ(cid:48)t2 exp δ(cid:48) 2 δ (cid:17) (cid:107)1B(0,2k−2)t2Le−δ(cid:48)t2L(1Ck+l−1t2Le−δt2Lf )(cid:107)1 (cid:46) 2kn exp (cid:16) − 42k+l−5 (cid:16) − 4k+l+1(cid:16) 4k−6 δ(cid:48) = 2kn exp (cid:46) exp(−4k+l)(cid:107)f(cid:107)1, (cid:107)1B(0,2k−2)t2Le−δ(cid:48)t2L1Ck+l−1(cid:107)1→1 (cid:46) t−n exp (cid:46) 2kn exp . Combining the two estimates we see that for 2−k−1 < t ≤ 2−k we have where in the last step we chose δ(cid:48) < δ small enough. The right-hand side of (3) is therefore dominated by ∞(cid:88) ∞(cid:88) k=2 l=1 (cid:90) 2−k 2−k−1 exp(−4k+l)(cid:107)f(cid:107)1 (cid:46) (cid:107)f(cid:107)1. dt t AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 9 (cid:3) Lemma 6. For any α > 0 we have Moreover, for α large enough we have (cid:107)(e−tLf )t=(cid:101)m(·)2/α(cid:107)1 (cid:46) (cid:107)f(cid:107)1. (cid:107)(tLe−tLf )t=(cid:101)m(·)2/α(cid:107)1 (cid:46) (cid:107)f(cid:107)1. Proof. Write C0 = B(0, 1) and Ck = B(0, 2k) \ B(0, 2k−1) for k ≥ 1. Moreover, let 0 = B(0, 2), C∗ 1 = B(0, 4), and C∗ C∗ We first show that for any α > 0, Denote ε = 1/α for notational convenience. For x ∈ Ck we have (cid:101)m(x)2 = 4−k and hence (cid:107)(e−tLf )t=(cid:101)m(·)2/α(cid:107)1 (cid:46) (cid:107)f(cid:107)1. k = B(0, 2k+1) \ B(0, 2k−2) for k ≥ 2. (4) (cid:107)(e−tLf )t=ε(cid:101)m(·)2(cid:107)1 = ∞(cid:88) (cid:107)1Ck e−ε4−kL(1C∗ f and 1Rn\C∗ k k f )(cid:107)1 ≤ f, and first estimate ∞(cid:88) (cid:107)1Ck e−ε4−kLf(cid:107)1. ∞(cid:88) k=0 (cid:107)1C∗ f(cid:107)1 (cid:46) (cid:107)f(cid:107)1. k We split f into 1C∗ k Fixing an integer N for which 8ε ≤ 4N, we use the trivial estimate for k = 0, 1, . . . , N + 3: k=0 k=0 For k ≥ N + 4 we have the decomposition Observing that d(Ck, B(0, 2k−2)) = 2k−2 we obtain, by Lemma 3, (cid:107)1Ck e−ε4−kL1B(0,2k−2)(cid:107)1→1 (cid:46) 2kn exp Furthermore, since d(Ck, Ck+l) = 2k+l−2, Lemma 3 implies that (cid:107)1Ck e−ε4−kL1Ck+l(cid:107)1→1 (cid:46) 2kn exp (cid:107)1Ck e−ε4−kL(1Rn\C∗ f )(cid:107)1 ≤ (cid:107)f(cid:107)1. k Rn \ C∗ k = B(0, 2k−2) ∪ Ck+l. ∞(cid:91) (cid:16) − 4k−2 l=2 (cid:17) (cid:16) 4k + 4k−2 (cid:17) 2 8ε4−k exp ≤ 2kn exp(−42k−2−N + 4k) (cid:46) exp(−4k). (cid:16) − 4k+l−2 8ε4−k (cid:17) (cid:16) 4k + 4k+l (cid:17) exp 2 ≤ 2kn exp(−42k+l−2−N + 4k+l) (cid:46) exp(−4k+l). (cid:16)(cid:107)1Ck e−ε4−kL(1B(0,2k−2)f )(cid:107)1 (cid:17) ∞(cid:88) (cid:16) (cid:107)1Ck e−ε4−kL(1Ck+lf )(cid:107)1 ∞(cid:88) exp(−4k)(cid:107)f(cid:107)1 + l=2 (cid:17) exp(−4k+l)(cid:107)f(cid:107)1 l=2 ∞(cid:88) k=N +4 + ∞(cid:88) (cid:46) k=N +4 (cid:46) (cid:107)f(cid:107)1. Therefore, ∞(cid:88) k=N +4 (cid:107)1Ck e−ε4−kL(1Rn\C∗ k f )(cid:107)1 = AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 10 We have now proven (4) which includes the first case from the statement of the lemma. such that for all x, y ∈ Rn The second case follows by using Lemma 4, which guarantees that there exists an α > 0 (cid:12)(cid:12)(cid:12)(t∂tMt(x, y))t=(cid:101)m(x)2/α (cid:12)(cid:12)(cid:12) (cid:46) exp (cid:16) α(cid:101)m(x)2 α x2(cid:17) M(cid:101)m(x)2(x, y) (cid:46) M(cid:101)m(x)2(x, y). Then (cid:107)(tLe−tLf )t=(cid:101)m(·)2/α(cid:107)1 (cid:46) (cid:107)(e−tLf )t=(cid:101)m(·)2(cid:107)1 (cid:46) (cid:107)f(cid:107)1. (cid:3) Finally, we turn to π3f = (cid:90) ∞ (cid:101)m(·)/κ (cid:101)Φ(t2)(t2L)2e−(δ(cid:48)+δ)t2Lf dt t . = c = c Repeating for the last term we get Proof. Integrating by parts we obtain Proposition 7. Let 0 < δ, δ(cid:48) ≤ 1/2. For κ large enough we have (cid:107)π3f(cid:107)1 (cid:46) (cid:107)f(cid:107)1 + (cid:107)(1 + log+ · ) M f(cid:107)1, where M f (x) = supεm(x)2<t≤1 e−tLf (x) and ε > 0 does not depend on f. ((cid:101)Φ(t) + t(cid:101)Φ(cid:48)(t))∂te−(δ(cid:48)+δ)tLf dt. + c(cid:48)(cid:90) ∞ (cid:101)m(·)2/κ2 + c(cid:48)(cid:90) ∞ (cid:101)m(·)2/κ2 (2(cid:101)Φ(cid:48)(t) + t(cid:101)Φ(cid:48)(cid:48)(t))e−(δ(cid:48)+δ)tLf dt. Now, having assumed that sup0<t<∞(Φ(t) + tΦ(cid:48)(t)) < ∞, we may use Lemma 6 to choose κ large enough so that (cid:90) ∞ (cid:101)Φ(t2)(t2L)2e−(δ(cid:48)+δ)t2Lf (cid:90) ∞ dt (cid:101)m(·)/κ (cid:101)Φ(t)t∂2 t t e−(δ(cid:48)+δ)tLf dt (cid:105)∞ (cid:104)(cid:101)Φ(t)t∂te−(δ(cid:48)+δ)tLf (cid:101)m(·)2/κ2 t=(cid:101)m(·)2/κ2 (cid:90) ∞ ((cid:101)Φ(t) + t(cid:101)Φ(cid:48)(t))∂te−(δ(cid:48)+δ)tLf dt (cid:105)∞ (cid:104) (cid:101)m(·)2/κ2 ((cid:101)Φ(t) + t(cid:101)Φ(cid:48)(t))e−(δ(cid:48)+δ)tLf t=(cid:101)m(·)2/κ2 (cid:13)(cid:13)(cid:13)(cid:104)(cid:101)Φ(t)t∂te−(δ(cid:48)+δ)tLf (cid:13)(cid:13)(cid:13)1 (cid:105)∞ (cid:46) (cid:107)(tLe−(δ(cid:48)+δ)tLf )t=(cid:101)m(·)2/κ2(cid:107)1 (cid:46) (cid:107)f(cid:107)1 t=(cid:101)m(·)2/κ2 and (cid:13)(cid:13)(cid:13)(cid:104) (cid:13)(cid:13)(cid:13)1 (cid:105)∞ ((cid:101)Φ(t) + t(cid:101)Φ(cid:48)(t))e−(δ(cid:48)+δ)tLf (cid:46) (cid:107)(e−(δ(cid:48)+δ)tLf )t=(cid:101)m(·)2/κ2(cid:107)1 (cid:46) (cid:107)f(cid:107)1. t=(cid:101)m(·)2/κ2 (cid:90) ∞ (cid:13)(cid:13)(cid:13)(cid:90) ∞ (cid:13)(cid:13)(cid:13)1 (2(cid:101)Φ(cid:48)(t) + t(cid:101)Φ(cid:48)(cid:48)(t))e−(δ(cid:48)+δ)tLf dt ((cid:101)Φ(cid:48)(t) + t(cid:101)Φ(cid:48)(cid:48)(t))(cid:107)e−(δ(cid:48)+δ)tLf(cid:107)1 dt (cid:46) (cid:107)f(cid:107)1. (cid:90) 1(cid:101)m(·)2/κ2 (cid:12)(cid:12)(cid:12)(cid:90) 1(cid:101)m(·)2/κ2 (cid:12)(cid:12)(cid:12) (cid:46) ((cid:101)Φ(cid:48)(t) + t(cid:101)Φ(cid:48)(cid:48)(t))e−(δ(cid:48)+δ)tLf dt (2(cid:101)Φ(cid:48)(t) + t(cid:101)Φ(cid:48)(cid:48)(t))e−(δ(cid:48)+δ)tLf dt Finally, having assumed that sup0<t<∞(tΦ(cid:48)(t) + t2Φ(cid:48)(cid:48)(t)) < ∞, we get Moreover, (cid:46) 1 = c 1 (cid:90) 1(cid:101)m(·)2/κ2 sup e−tLf (cid:46) (cid:46) (1 + log+ · ) M f, εm(·)2<t≤1 dt t (cid:3) where ε > 0 is chosen small enough depending on δ, δ(cid:48) and κ. AN L1-ESTIMATE FOR SPECTRAL MULTIPLIERS 11 References [1] A. Amenta and M. Kemppainen. Non-uniformly local tent spaces. Publ. Mat., 59(1):245 -- 270, 2015. [2] A. Carbonaro and O. Dragičević. Functional calculus for generators of symmetric contraction semi- groups. 2013. arXiv:1308.1338. [3] M. G. Cowling. Harmonic analysis on semigroups. Ann. of Math. (2), 117(2):267 -- 283, 1983. [4] J. García-Cuerva, G. Mauceri, S. Meda, P. Sjögren, and J. L. Torrea. Functional calculus for the Ornstein-Uhlenbeck operator. J. Funct. Anal., 183(2):413 -- 450, 2001. [5] J. García-Cuerva, G. Mauceri, P. Sjögren, and J. L. Torrea. Spectral multipliers for the Ornstein- Uhlenbeck semigroup. J. Anal. Math., 78:281 -- 305, 1999. [6] W. Hebisch, G. Mauceri, and S. Meda. Holomorphy of spectral multipliers of the Ornstein-Uhlenbeck operator. J. Funct. Anal., 210(1):101 -- 124, 2004. [7] S. Janson. Gaussian Hilbert spaces, volume 129 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1997. [8] J. Maas, J. van Neerven, and P. Portal. Conical square functions and non-tangential maximal functions with respect to the Gaussian measure. Publ. Mat., 55(2):313 -- 341, 2011. [9] J. Maas, J. van Neerven, and P. Portal. Whitney coverings and the tent spaces T 1,q(γ) for the Gaussian [10] G. Mauceri and S. Meda. BMO and H 1 for the Ornstein-Uhlenbeck operator. J. Funct. Anal., measure. Ark. Mat., 50(2):379 -- 395, 2012. 252(1):278 -- 313, 2007. [11] P. Portal. Maximal and quadratic Gaussian Hardy spaces. Rev. Mat. Iberoam., 30(1):79 -- 108, 2014. [12] P. Sjögren. Operators associated with the Hermite semigroup -- a survey. In Proceedings of the confer- ence dedicated to Professor Miguel de Guzmán (El Escorial, 1996), volume 3, pages 813 -- 823, 1997. [13] E. M. Stein. Topics in harmonic analysis related to the Littlewood-Paley theory. Annals of Mathematics Studies, No. 63. Princeton University Press, Princeton, N.J., 1970. [14] J. van Neerven and P. Portal. Finite speed of propagation and off-diagonal bounds for Ornstein -- Uhlenbeck operators in infinite dimensions. 2015. arXiv:1507.02082. Department of Mathematics and Statistics, University of Helsinki, Gustaf Hällströmin katu 2b, FI-00014 Helsinki, Finland E-mail address: [email protected]
1606.06158
1
1606
2016-06-20T14:59:06
New formulas for the spectral radius via Aluthge transform
[ "math.FA", "math.SP" ]
In this paper we give several expressions of spectral radius of a bounded operator on a Hilbert space, in terms of iterates of Aluthge transformation, numerical radius and the asymptotic behavior of the powers of this operator. Also we obtain several characterizations of normaloid operators.
math.FA
math
NEW FORMULAS FOR THE SPECTRAL RADIUS VIA ALUTHGE TRANSFORM FADIL CHABBABI AND MOSTAFA MBEKHTA Abstract. In this paper we give several expressions of spectral radius of a bounded operator on a Hilbert space, in terms of iterates of Aluthge transformation, numerical radius and the asymptotic behavior of the powers of this operator. Also we obtain several characterizations of normaloid operators. 1. Introduction Let H be complex Hilbert spaces and B(H) be the Banach space of all bounded linear operators from H into it self. For T ∈ B(H), the spectrum of T is denoted by σ(T ) and r(T ) its spectral radius. We denote also by W(T ) and w(T ) the numerical range and the numerical radius of T . As usually, for T ∈ B(H) we denote the module of T by T = (T ∗T )1/2 and we shall always write, without further mention, T = UT to be the unique polar decomposition of T , where U is the appropriate partial isometry satisfying N(U) = N(T ). The Aluthge transform introduced in [1] as ∆(T ) = T 1 2 UT 1 2 , T ∈ B(H), to extend some properties of hyponormal operators. Later, in [8], Okubo introduced a more general notion called λ−Aluthge transform which has also been studied in detail. For λ ∈ [0, 1], the λ-Aluthge transform is defined by, ∆λ(T ) = T λUT 1−λ, T ∈ B(H). Notice that ∆0(T ) = UT = T , and ∆1(T ) = T U which is known as Duggal's trans- form. It has since been studied in many different contexts and considered by a number of authors (see for instance, [1, 2, 3, 7, 6, 5] and some of the references there). The interest of the Aluthge transform lies in the fact that it respects many properties of the original operator. For example, (see [5, Theorems 1.3, 1.5]) (1.1) σ(∆λ(T )) = σ(T ), for every T ∈ B(H), Another important property is that Lat(T ), the lattice of T -invariant subspaces of H, is nontrivial if and only if Lat(∆(T )) is nontrivial (see [5, Theorem 1.15]). 2000 Mathematics Subject Classification. 47A13, 47A30, 47B37. Key words and phrases. spectral radius, polar decomposition, λ-Aluthge transform, normaloid operator This work was supported in part by the Labex CEMPI (ANR-11-LABX-0007-01). 1 2 FADIL CHABBABI AND MOSTAFA MBEKHTA Moreover, Yamazaki ([12]) (see also, [11, 10]), established the following interesting formula for the spectral radius (1.2) lim n→∞ k∆n λ(T )k = r(T ) where ∆n λ the n-th iterate of ∆λ, i.e ∆n+1 λ (T ) = ∆λ(∆n λ(T )), ∆0 λ(T ) = T . In this paper we give several expressions of the spectral radius of an operator. Firstly in terms of the Aluthge transformation (section 2), and secondly, in section 3, we give several expressions of spectral radius, based on numerical radius and Aluthge transformation. Also, We infer several characterizations of normaloid operators (i.e. r(T ) = kT k). 2. formulas of spectral radius via Aluthge transform In this section, we use the of Rota's Theorem, order to obtain new formulas of spectral radius via Aluthge transformation Theorem 2.1. For every operator T ∈ B(H), we have r(T ) = inf{k∆λ(XT X−1)k, X ∈ B(H) invertible } = inf{k∆λ(eAT e−A)k, A ∈ B(H) self adjoint }. Proof. For every invertible operator X ∈ B(H), by (1.1), we have σ(∆λ(XT X−1)) = σ(XT X−1) = σ(T ). It follows that r(T ) = r(∆λ(XT X−1)) ≤ k∆λ(XT X−1)k for every invertible operator X ∈ B(H). Hence r(T ) ≤ inf{k∆λ(XT X−1)k; X ∈ B(H) invertible } ≤ inf{k∆λ(exp(A)T exp(−A))k; A ∈ B(H) self adjoint }, In the other hand, for ε > 0, we have T r(cid:0) r(T ) + ε (cid:1) = r(T ) r(T ) + ε < 1. T r(T ) + ε is similar to an contraction. Thus there From Rota's Theorem [9, Theorem 2], exists an invertible operator Xε ∈ B(H) such that ε )k ≤ kXεT X−1 (2.1) Now, let Xε = UεXε be the polar decomposition of Xε. Clearly Uε is a unitary oper- ator, and Xε is invertible. Therefore there exist α > 0 such that σ(Xε) ⊆ [α, +∞[. Consequently Aε = ln(Xε) exits and it is self adjoint, we also have ε k ≤ r(T ) + ε. k∆λ(XεT X−1 Xε = eAε and Xε−1 = e−Aε. NEW FORMULAS FOR THE SPECTRAL RADIUS VIA ALUTHGE TRANSFORM 3 Thus Hence k∆λ(XεT X−1 ε )k = k∆λ(UeAεT e−AεU ∗)k = kU∆λ(eAεT e−Aε)U ∗k = k∆λ(eAεT e−Aε)k. k∆λ(eAεT e−Aε)k ≤ kXεT X−1 ε k ≤ r(T ) + ε. It follows that for all ε > 0, r(T ) ≤ inf{k∆λ(XT X−1)k, X ∈ B(H) invertible } ≤ inf{k∆λ(eAT e−A)k, A ∈ B(H) self adjoint } ≤ k∆λ(eAεT e−Aε)k ≤ kXεT X−1 ε k ≤ r(T ) + ε. Finally, since ε > 0 is arbitrary, we obtain r(T ) = inf{k∆λ(XT X−1)k, X ∈ B(H) invertible } = inf{k∆λ(eAT e−A)k, A ∈ B(H) self adjoint }. Theroforte the proof of Theorem is complete. (cid:3) As immediate consequence of the Theorem 2.1, we obtain the following corollary which gives a formula of the spectral radius based on n-th iterate of ∆λ. Corollary 2.1. If T ∈ B(H), then for every n ≥ 0, r(T ) = inf{k∆n = inf{k∆n λ(XT X−1)k, X ∈ B(H) invertible } λ(eAT e−A)k, A ∈ B(H) self adjoint }. Proof. First, note that k∆λ(T )k ≤ kT k, consequently we have k∆n (2.2) Now, clearly σ(∆n operator X ∈ B(H) we have λ(T )k ≤ k∆n−1 λ(T )) = σ(T ), for all n ∈ N. It follows that, for every invertible (T )k ≤ ... ≤ k∆λ(T )k ≤ kT k, ∀n ∈ N∗. λ λ(XT X−1)) r(T ) = r(∆n λ(XT X−1)k ≤ k∆n ≤ k∆λ(XT X−1)k. Therefore λ(XT X−1)k, X ∈ B(H) invertible } r(T ) ≤ inf{k∆n ≤ inf{k∆n λ(eAT e−A)k, A ∈ B(H) self adjoint } ≤ inf{k∆λ(eAT e−A)k, A ∈ B(H) self adjoint } = r(T ). (cid:3) 4 FADIL CHABBABI AND MOSTAFA MBEKHTA An operator T is said to be normaloid if r(T ) = kT k. As immediate consequence of the Corollary 2.1, we obtain the following corollary which is a characterization of normaloid operators via λ-Aluthge transformation : Corollary 2.2. If T ∈ B(H), then the following assertions are equivalent (i) T is normaloid; (ii) kT k ≤ k∆λ(XT X−1)k, for all invertible X ∈ B(H); (iii) kT k ≤ k∆n λ(XT X−1)k, for all invertible X ∈ B(H) and for all natural number n. As immediate consequence of the Corollary 2.2, we obtain a new characterization of normaloid operators Corollary 2.3. If T ∈ B(H), then the following assertions are equivalent (i) T is normaloid; (ii) kT k ≤ kXT X−1k, for all invertible X ∈ B(H); Theorem 2.2. Let T ∈ B(H). Then for each natural number n, we have r(T ) = lim k = lim k k∆n λ(T k)k1/k k∆λ(T k)k1/k. Proof. Note that, (2.3) r(T ) = r(∆n λ(T )) ≤ k∆n λ(T )k ≤ k∆λ(T )k ≤ kT k ∀n ∈ N∗. Let k ∈ N be arbitrary, we have r(T )k = r(T k) = r(∆n λ(T k)) ≤ k∆n λ(T k)k ≤ k∆λ(T k)k ≤ kT kk ∀n ∈ N. Hence Therefore r(T ) ≤ k∆n λ(T k)k1/k ≤ k∆λ(T k)k1/k ≤ kT kk1/k. r(T ) ≤ lim k k∆n λ(T k)k1/k ≤ lim k k∆λ(T k)k1/k ≤ lim k kT kk1/k = r(T ). Which completes the proof. (cid:3) As immediate consequence of Theorem 2.2, we obtain the following corollary which is a new characterization of normaloid operators Corollary 2.4. If T ∈ B(H), then the following assertions are equivalent (i) T is normaloid; (ii) kT kk = k∆λ(T k)k, for all natural number k; (iii) kT kk = k∆n λ(T k)k, for every natural number k, n NEW FORMULAS FOR THE SPECTRAL RADIUS VIA ALUTHGE TRANSFORM 5 3. Spectral radius via numerical radius and Aluthge transform For T ∈ B(T ), we denote the numerical range and numerical radius of T by W(T ) and w(T ), respectively. W(T ) = {< T x, x >; kxk = 1} and w(T ) = sup{λ; λ ∈ W(T )}. In the following theorem, we obtain a new expression of the spectral radius by means of the numerical radius and Aluthge transform. Theorem 3.1. For every operator T ∈ B(H) and for each natural number n, we have r(T ) = inf{w(∆n = inf{w(∆n λ(XT X−1)), X ∈ B(H) invertible } λ(eAT e−A)), A ∈ B(H) self adjoint }. Proof. It is well known that r(T ) ≤ w(T ) ≤ kT k. Thus, for all X ∈ B(H), invertible and for each natural number n, we have r(T ) = r(∆n λ(XT X−1)) ≤ w(∆n λ(XT X−1)) ≤ k∆n λ(XT X−1)k It follows that r(T ) ≤ inf{w(∆n ≤ inf{w(∆n ≤ inf{k∆n = r(T ) λ(XT X−1)); X ∈ B(H) invertible } λ(exp(A)T exp(−A))); A ∈ B(H) self adjoint }, λ(exp(A)T exp(−A))k; A ∈ B(H) self adjoint } ( by Corollary 2.1). Hence we obtain the desired equalities. (cid:3) For a bounded linear operator S , we will write Re(S ) = 1 2(S + S ∗), the real part of S . And we denote by W(S ) the closure of the numerical range of S . Then we have de following result Theorem 3.2. For every operator T ∈ B(H), there exists θ ∈ R such that for all natural number n, r(T ) = inf{w(Re(∆n = inf{kRe(∆n λ(exp(iθ)XT X−1))), X ∈ B(H) invertible } λ(exp(iθ)XT X−1))k, X ∈ B(H) invertible }. Proof. First assume that r(T ) ∈ σ(T ). Then for all invertible X ∈ B(H), we have r(T ) ∈ Re(σ(T )) = Re(σ(∆n λ(XT X−1))). Thus r(T ) ∈ Re(σ(∆n λ(XT X−1))) ⊆ Re(W(∆n λ(XT X−1))) = W(Re(∆n λ(XT X−1))) which implies r(T ) ≤ w(Re(∆n ≤ kRe(∆n ≤ k∆n λ(XT X−1))) λ(XT X−1))k λ(XT X−1)k. 6 FADIL CHABBABI AND MOSTAFA MBEKHTA Since the last inequalities are satisfied for all X ∈ B(H) invertible, we obtain r(T ) ≤ inf{w(Re(∆n ≤ inf{kRe(∆n ≤ inf{k∆n = r(T ) (by Corollary 2.1). λ(XT X−1))) X ∈ B(H) invertible } λ(XT X−1))k X ∈ B(H) invertible } λ(XT X−1)k X ∈ B(H) invertible } We have shown that if r(T ) ∈ σ(T ) then r(T ) = inf{w(Re(∆n = inf{kRe(∆n λ(XT X−1))) X ∈ B(H) invertible } λ(XT X−1))k X ∈ B(H) invertible } Now, if T is an arbitrary operator, then there exists z ∈ σ(T ) such that, z = r(T ). Put θ = − arg(z). Then r(T ) = z exp(iθ) ∈ σ(exp(iθ)T ). Hence by the first part of de proof, we conclude that r(T ) = r(exp(iθ)T ) = inf{w(Re(∆n = inf{kRe(∆n λ(exp(iθ)XT X−1))), X ∈ B(H) invertible } λ(exp(iθ)XT X−1))k, X ∈ B(H) invertible }. This completes the proof of the theorem. (cid:3) As immediate consequence of Theorem 3.2, we obtain the following corollary which is a characterization of normaloid operators Corollary 3.1. If T ∈ B(H), and a natural number n, then the following assertions are equivalent (i) T is normaloid; (ii) there exists θ ∈ R such that, for all X ∈ B(H) invertible kT k ≤ w(Re(∆n λ(exp(iθ)XT X−1))); (iii) there exists θ ∈ R such that, for all X ∈ B(H) invertible kT k ≤ kRe(∆n λ(exp(iθ)XT X−1))k. We end this paper by the following theorem which gives a new formula of the spec- tral radius of T , in terms of the asymptotic behavior of powers and the numerical radius of T Theorem 3.3. For every operator T ∈ B(H) and for each natural number n, we have r(T ) = lim k w(∆n λ(T k))1/k. Proof. For each natural number n and k, we have λ(T k)) ≤ w(∆n r(T )k = r(T k) = r(∆n λ(T k)) ≤ k∆n λ(T k)k Hence r(T ) ≤ w(∆n λ(T k))1/k ≤ k∆n λ(T k)k1/k. NEW FORMULAS FOR THE SPECTRAL RADIUS VIA ALUTHGE TRANSFORM 7 By Theorem 2.2, we deduce that r(T ) = lim k w(∆n λ(T k))1/k. Which completes the proof. (cid:3) References [1] A. Aluthge, On p-hyponormal operators for 0 < p < 1, Integral Equations Operator Theory 13 (1990), 307-315. [2] T. Ando and T. Yamazaki, The iterated Aluthge transforms of a 2-by-2 matrix converge, Linear Algebra Appl. 375 (2003), 299-309 [3] AJ. Antezana, P. Massey and D. Stojanoff, λ-Aluthge transforms and Schatten ideals, Linear Algebra Appl, 405 (2005), 177-199. [4] T. Furuta, Invitation to linear operators, Taylor Francis, London 2001. [5] I. Jung, E. Ko, and C. Pearcy , Aluthge transform of operators, Integral Equations Operator Theory 37 (2000), 437-448. [6] I. Jung, E. Ko, C. Pearcy , Spectral pictures of Aluthge transforms of operators, Integral Equations Operator Theory 40 (2001), 52-60. [7] I. Jung, E. Ko, C. Pearcy , The iterated Aluthge transform of an operator, Integral Equations Operator Theory 45 (2003), 375-387. [8] K. Okubo, On weakly unitarily invariant norm and the Aluthge transformation, Linear Algebra Appl. 371 (2003), 369 -- 375. [9] G. Rota, On models for linear operators, comm. Pure Appl. Math. 13 (1960), 496 -- 472. [10] T. Tam, λ- Aluthge iteration and spectral radius, Integral Equations Operator Theory 60 (2008), 591-596. [11] D.Wang, Heinz and McIntosh inequalities, Aluthge tranformation and the spectral radius, Math. Inequal. Appl. 6 (2003), 121-124. [12] T. Yamazaki, An expression of the spectral radius via Aluthge tranformation, Proc. Amer. Math. Soc. 130 (2002), 1131-1137. Universit´e Lille1, UFR de Math´ematiques, Laboratoire CNRS-UMR 8524 P. Painlev´e, 59655 Vil- leneuve d'Ascq Cedex, France E-mail address: [email protected] E-mail address: [email protected]
1602.05496
2
1602
2016-04-28T15:35:13
A Gr\"uss type operator inequality
[ "math.FA" ]
In [P. Renaud, "A matrix formulation of Gr\"uss inequality", Linear Algebra Appl. 335 (2001), 95--100] it was proved an operator inequality involving the usual trace functional. In this article, we give a refinement of such result and we answer positively the Renaud's open problem.
math.FA
math
A GR USS TYPE OPERATOR INEQUALITY T. BOTTAZZI1 AND C. CONDE1,2 Abstract. In [P. Renaud, A matrix formulation of Gruss inequality, Linear Algebra Appl. 335 (2001), 95 -- 100] it was proved an operator inequality involving the usual trace functional. In this article, we give a refinement of such result and we answer positively the Renaud's open problem. 1. Introduction In 1935, Gruss [6] obtained the following inequality: if f, g are integrable real functions on [a, b] and there exist real constant α, β, γ, δ such that α ≤ f (x) ≤ β, γ ≤ g(x) ≤ δ for all x ∈ [a, b] then 1 b − aZ b a f (x)g(x)dx − 1 (b − a)2 Z b a f (x)dxZ b a 1 4 (β − α)(δ − γ), (1.1) (cid:12)(cid:12)(cid:12)(cid:12) g(x)dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ and the inequality is sharp, in the sense that the constant 1 4 cannot be replaced by a smaller one. This inequality has been investigated, applied and generalized by many mathemati- cians in different areas of mathematics, such as inner product spaces, quadrature formulae, finite Fourier transforms, linear functionals, etc. Along this work H denotes a (complex, separable) Hilbert space with inner product h·, ·i. Let (B(H), k · k) be the C ∗-algebra of all bounded linear operators acting on (H, h·, ·i) with the uniform norm. We denote by Id the identity operator, and for any A ∈ B(H) we consider A∗ its adjoint and A = (A∗A) 2 the absolute value of A. By B(H)+ we denote the cone of positive operators of B(H), i.e. B(H)+ := {T ∈ B(H) : hT h, hi ≥ 0 ∀h ∈ H}. In the case when dim H = n, we identify B(H) with the full matrix algebra Mn of all n × n matrices with entries in the complex field C. For each T ∈ B(H), we denote its spectrum 1 by σ(T ), that is, σ(T ) = {λ ∈ C : T − λId is not invertible} and a complex number λ ∈ C is said to be in the approximate point spectrum of the operator T , and we denote by σap(T ), if there is a sequence {xn} of unit vectors satisfying (T − λ)xn → 0. 2010 Mathematics Subject Classification. Primary: 39B05, 47A12, 47A30; Secondary: 39B42, 47B10. Key words and phrases. Gruss inequality; variance; trace inequality, distance formula. 1 2 T. BOTTAZZI, C. CONDE For each operator T we consider r(T ) = sup{λ : λ ∈ σ(T )} spectral radius of T, W (T ) = {hT h, hi : khk = 1} numerical range of T and w(T ) = sup{λ : λ ∈ W (T )} numerical radius of T. Recall that for all T ∈ B(H), r(T ) ≤ w(T ) ≤ kT k ≤ 2w(T ), σ(T ) ⊆ W (T ) and by the Toeplitz-Hausdorff's Theorem W (T ) is convex. Renaud [10] gave a bounded linear operator analogue of Gruss inequality by replacing integrable functions by operators and the integration by a trace function as follows: let A, T ∈ B(H), suppose that W (A) and W (T ) are contained in disks of radii RA and RT , respectively. Then for any positive trace class operator P with tr(P ) = 1 holds tr(P AT ) − tr(P A)tr(P T ) ≤ 4RART , (1.2) and if A and T are normal (i.e. T ∗T = T T ∗), the constant 4 can be replaced by 1. We can see can easily see that if A = αId or T = βId with α, β ∈ C then the left hand side is equal to zero. In the same article, Renaud proposed the following open problem: to characterise k(A, T ), where tr(P AT ) − tr(P A)tr(P T ) ≤ k(A, T )RART , (1.3) with 1 ≤ k(A, T ) ≤ 4. In particular, whether it depends on A and T separately, i.e. whether we can write k(A, T ) = h(A)h(T ), where h(A), h(B) are suitably defined constants. In this paper we give a positive answer to the open problem proposed by Renaud and we obtain an explicit formula for k(A, T ) = h(A)h(T ). Also, we generalize the inequality (1.2) for normal to transloid operators. 2. Preliminaries Let us begin with the notation and the necessary definitions. The set of compact operators in H is denoted by B0(H). If T ∈ B0(H) we denote by {sn(T )} the sequence of singular values of T , i.e., the eigenvalues of T (decreasingly ordered). The notion of unitary invariant norms can be defined also for operators on Hilbert spaces. A norm . that satisfies the invariance property UXV = X. If dim R(T ) = 1, then T = s1(T )g(e1) = g(e1)kT k. By convention, we assume that g(e1) = 1. If x, y ∈ H, then we denote x ⊗ y the rank one operator defined on H by (x ⊗ y)(z) = hz, yix then kx ⊗ yk = kxkkyk = x ⊗ y. A GR USS TYPE OPERATOR INEQUALITY 3 The most known examples of unitary invariant norms are the Schatten p-norms For 1 ≤ p < ∞, let and kT kp p =Xn sn(T )p = tr Xp , Bp(H) = {T ∈ H : kT kp < ∞}, called the p−Schattenclass of B(H). That is the subset of compact operators with singular values in lp. The positive operators with trace 1 are called density operator (or states) and we denote this set by S(H). The ideal B2(H) is called the Hilbert-Schmidt class and it is a Hilbert space with the inner product hS, T i2 = tr(ST ∗). On the theory of norm ideals and their associated unitarily invariant norms, a reference for this subject is [5]. An operator A ∈ B(H) is called normaloid if r(A) = kAk = ω(A). If A−µId is normaloid for all µ ∈ C, then A is called transloid. Finally, for A, T ∈ B(H) and P ∈ S(H) we introduce the following notation VP (A, T ) = tr(P AT ) − tr(P A)tr(P T ). In the particular case T = A∗ we get the variance of A respect to P . More precisely, Audenaert in [1] consider the following notion, given A, P ∈ Mn, P ≥ 0, tr(P ) = 1 the variance of A respect to the matrix P VP (A) = tr(A2P ) − tr(AP )2 = VP (A, A∗), Note that VP (A − λId) = VP (A). Futhermore, he showed that if A ∈ Mn then max{tr(A2P ) − tr(AP )2 : P ∈ M+ n , tr(P ) = 1} = dist(A, CId)2, (2.1) and the maximization over P on the left hand side can be restricted to density matrices of rank 1. 3. Distance formulas and Renaud's inequality Let A and T linear bounded operators acting in H, the vector-function A − λT is known as the pencil generated by A and T . Evidently there is at least one complex number λ0 such that kA − λ0T k = inf λ∈C kA − λT k. The number λ0 is unique if 0 /∈ σap(T ) (or equivalently if inf{kT xk : kxk = 1} > 0). Different authors, following [11], called to this unique number as center of mass of A respect 4 T. BOTTAZZI, C. CONDE to T and we denote by c(A, T ) and when T = Id we write c(A). Following Paul, for A, T ∈ B(H) such that 0 /∈ σap(T ) we consider MT (A) = sup kxk=1(cid:20)kAxk2 − hAx, T xi2 hT x, T xi (cid:21)1/2 = sup kxk=1(cid:13)(cid:13)(cid:13)(cid:13)Ax − in [8], he proved that MT (A) = dist(A, CT ). The unique minimizer is characterized by the following conditions: there exists a sequence of unit vectores {xn} such that hAx, T xi hT x, T xi T x(cid:13)(cid:13)(cid:13)(cid:13) , (3.1) k(A − λ0T )xnk → kA − λ0T k and h(A − λ0T )xn, xni → 0. In [4], Gevorgyan proved that c(A, T ) = lim n→∞ hAyn, T yni hT yn, T yni , (3.2) where {yn} is a sequence of unit vectores which approximate the supremum in (3.1). In the particular case that T = Id and A is a Hermitian operator then it is easy to see that min λ∈C kA − λIdk = λmax(A) − λmin(A) 2 , (3.3) where λmax(A) (resp. λmax(A)) denotes the maximum (resp. mnimum) eigenvalue of A. Observe that the minumumis attained at c(A) = λmax(A) + λmin(A) 2 . We recall other formulas that express the distance from A to the one-dimensional sub- space CT . Then dist(A, CT ) = sup{hAx, yi : kxk = kyk = 1, hT x, yi = 0}, (3.4) if A, T ∈ B(H) and 0 /∈ σap(T ). In the particular case, where T = Id we get dist(A, CId) = 1 2 sup{kAX − XAk : X ∈ B(H), kXk = 1} = sup{k(Id − Q)AQk : Q is a rank one projection} = sup{k(Id − Q)AQk : Q is a projection}. (3.5) In the following statement we present a new proof of the relation between the variance of A respect to P and the distance from A to the unidimensional subspace CId. A GR USS TYPE OPERATOR INEQUALITY 5 Proposition 3.1. Let A ∈ B(H) and P ∈ S(H) then tr(A2P ) − tr(AP )2 = kAP 1/2k2 2 − hAP 1/2, P 1/2i22 = kAP 1/2 − hAP 1/2, P 1/2i2P 1/2k2 2 = min λ∈C kAP 1/2 − λP 1/2k2 2 ≤ min λ∈C kA − λIdk. Proof. These inequalities are simple consequences from following general statement for any Hilbert space H: let x, y ∈ H with y 6= 0 then kx − λyk2 = inf λ∈C kxk2kyk2 − hx, yi2 kyk2 . (cid:3) The following statement is an extension of the Audenaert's formula to infinite dimension. Remark 3.2 (Audenaert's formula for infinite dimensional spaces). We exhibit that the equality (2.1) holds in infinite dimensional context, that is for A ∈ B(H) holds sup{[tr(A2P ) − tr(AP )2]1/2 : P ∈ S(H)} = dist(A, CId). (3.6) First, we obtain this equality from a Prasanna's result in [9]. Indeed, note that dist(A, CId)2 = sup kxk=1 kAxk2 − hAx, xi2 ≤ sup{tr(A2P ) − tr(AP )2 : P ∈ S(H)} ≤ dist(A, CId)2. On the other hand, another way to prove (3.6) is to reduce the problem to finite dimension and use the classical Audenaert's formula. Now we give the idea of this proof. For the sake of clarity, we denote m := min λ∈C kA − λIdk and M := sup{[tr(A2P ) − tr(AP )2]1/2 : P ∈ S(H)}. By Proposition 3.1 we have that M ≤ m. Suppose by contradiction that M < m then there exists ǫ > 0 such that M < kA − λIdk − ǫ, (3.7) 6 T. BOTTAZZI, C. CONDE for any λ ∈ C. By the equality (3.2), we have that c(A) ∈ W (A) and then c(A) ≤ w(A). As any closed ball in the complex plane is a compact set, we can find λ1, ..., λm ∈ H such that B(0, ω(A)) ⊆ ∪m j=1{λ ∈ C : λ − λj < ǫ 2 }. Now, we choose unit vectors h1, ..., hm ∈ H with the following property: k(A − λjId)hjk > kA − λjIdk − ǫ 2 . Let H′ = gen{h1, ..., hm, Ah1, ..., Ahm} and n = dim H′. Applying (2.1) to the compressions of A and Id respectively, we get dist(A′, CIdn) = max{[tr(A′2P ′) − tr(A′P ′)2]1/2 : P ′ ∈ M+ n , tr(P ′) = 1} = M ′. (3.8) One easily verifies that if λ ∈ B(0, ω(A)) there exists j ∈ {1, ..., m} such that kA′ − λIdnk > kA′ − λjIdnk − ǫ 2 = k(A − λjId)hjk − ≥ k(A′ − λjIdn)hjk − ǫ 2 > kA − λjIdk − ǫ. ǫ 2 Thus, combining (3.7) and (3.9) we get kA′ − λIdnk > M ≥ M ′, min λ∈C and we have here a contradiction with (3.8), therefore m = M. The next result gives and upper bound for VP (A, T ). (3.9) (3.10) Proposition 3.3. Let A, T ∈ B(H) and P ∈ S(H). Then, for any λ, µ ∈ C holds VP (A, T ) ≤ (A − λId, A − λId)1/2 2,P (T ∗ − ¯µId, T ∗ − ¯µId)1/2 2,P − GId(A − λId, T ∗ − ¯µId) ≤ kA − λIdk kT − µIdk − tr (P (A − λId)) tr (P (T − µId)) , (3.11) with GId(A − λId, T ∗ − ¯µId) = tr (P (A − λId)) tr (P (T − µId)). Therefore, VP (A, T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId). Proof. Define the following semi-inner product for X, Y ∈ B(H) and P ∈ S(H): (3.12) (X, Y )2,P =(cid:10)P 1/2X, P 1/2Y(cid:11)2 . Following the proof given by Dragomir in [[3],Theorem 2], holds for any E ∈ B(H) such that (E, E)2,P = 1 (X, Y )2,P − (X, E)2,P (E, Y )2,P ≤ (X, X)1/2 = (X, X)1/2 2,P (Y, Y )1/2 2,P (Y, Y )1/2 2,P − (X, E)2,P (E, Y )2,P . 2,P − GE(X, Y ). (3.13) A GR USS TYPE OPERATOR INEQUALITY 7 Since (Id, Id)2,P = 1, then VP (A, T ) = VP (A − λId, T − µId) = (A − λId, (T − µId)∗)2,P − (A − λId, Id)2,P (Id, (T − µId)∗)2,P ≤ (A − λId, A − λId)1/2 2,P (T ∗ − ¯µId, T ∗ − ¯µId)1/2 2,P − GId(A − λId, T ∗ − ¯µId) = tr(cid:0)P (A − λId)∗2(cid:1)1/2 tr(cid:0)P T − µId2(cid:1)1/2 − GId(A − λId, T ∗ − ¯µId) ≤ (cid:13)(cid:13)(A − λId)∗2(cid:13)(cid:13)1/2(cid:13)(cid:13)T − µId2(cid:13)(cid:13)1/2 − tr (P (A − λId)) tr (P (T − µId)) = kA − λIdk kT − µIdk − tr (P (A − λId)) tr (P (T − µId)) . (3.14) Therefore, sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId). (cid:3) Remark 3.4. If we define VP : B(H) × B(H) → C, VP (A, T ) := tr(P AT ) − tr(P A)tr(P T ). Then VP is a bilinear function and by (3.12) a continuous mapping with kVP k ≤ 1. Now, we give a new proof and a refinement of (1.2). Proposition 3.5. Let A, T ∈ B(H) and we suppose that W (A), W (T ) are contained in closed disk D(λ0, RA), D(µ0, RT ) respectively. Then for any P ∈ S(H) tr(P AT ) − tr(P A)tr(P T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId) ≤ kA − λ0IdkkT − µ0Idk ≤ 4w(A − λ0Id)w(T − µ0Id) ≤ 4RART . In particular, if A and T are normal operators, we have tr(P AT ) − tr(P A)tr(P T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId) = rArT , (3.15) (3.16) where rS denotes the radius of the unique smallest disc containing σ(S) for any S ∈ B(H). Proof. The inequalities are consequence of (3.12). In the last inequality we use that W (A − λ0Id) ⊂ D(0, RA) and W (T − µ0Id) ⊂ D(0, RT ) respectively. 8 T. BOTTAZZI, C. CONDE On the other hand, Bjorck and Thom´ee [2] have shown that for a normal operator A dist(A, CId) = sup kxk=1 (kAxk2 − hAx, xi2)1/2 = rA, and this completes the proof. (3.17) (cid:3) Remark 3.6. From (3.16), if we consider A is a positive invertible operator, T = A−1 and P = x ⊗ x with x ∈ H with kxk = 1, then tr(P AT ) − tr(P A)tr(P T ) = 1 − hAx, xihA−1x, xi ≤ dist(A, CId)dist(A−1, CId) = rArA−1, i.e. we obtain the Kantorovich inequality for an operator A acting on an infinite dimensional Hilbert space H with 0 < m ≤ A ≤ M. In 1972, Istratescu ([7]) generalized the equality (3.17) to the transloid class operators, then we have the following statement: Proposition 3.7. Let A, T ∈ B(H) with A and T transloid operators then tr(P AT ) − tr(P A)tr(P T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId) = rArT . Proof. It follows from the same arguments in the proof of inequality (3.16). (3.18) (cid:3) The previous proposition generalizes the Renaud's result for normal operators, since the classes of transloid and normal operators are related by the inclusion as follows normal ⊆ quasinormal ⊆ subnormal ⊆ hyponormal ⊆ transloid, where at least the first inclusion is proper. In the following statement we obtain a parametric refinement of (1.2). Theorem 3.8. Let A, T ∈ B(H) with A, T /∈ CId and suppose that W (A), W (T ) are contained in the closed disk D(λ0, RA) and D(µ0, RT ) respectively. Thus for any P ∈ S(H) A GR USS TYPE OPERATOR INEQUALITY 9 tr(P AT ) − tr(P A)tr(P T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(eP T ) ≤ dist(A, CId)dist(T, CId) ≤ hλ(A)hµ(T )ω(A − λ0Id)ω(T − µ0Id) ≤ hλ(A)hµ(T )RART , (3.19) we get where hλ(A) = 2(1 − λ) + λ kA − c(A)Idk w(A − λ0Id) , hµ(T ) = 2(1 − µ) + µ kT −c(T )Idk w(T −µ0Id) and 1 ≤ hλ(A)hµ(T ) ≤ 4, for any λ, µ ∈ [0, 1]. Proof. Let λ ∈ [0, 1]. Then, kA − c(A)Idk ≤ λkA − c(A)Idk + (1 − λ)kA − λ0Idk ≤ λkA − c(A)Idk + 2(1 − λ)w(A − λ0Id) = w(A − λ0Id)(cid:18)2(1 − λ) + λ kA − c(A)Idk w(A − λ0Id) (cid:19) = w(A − λ0Id)hλ(A), where 1 ≤ hλ(A) ≤ 2 since kA − c(A)Idk ≤ kA − λ0Idk ≤ 2w(A − λ0Id). This inequality completes the proof. (cid:3) Note that the previous result gives a positive answer at the Renuad's open question (1.3). Corollary 3.9. Under the same notation as in Theorem 3.8, if A − λ0Id and T − µ0Id are normaloid operators then, for any λ, µ ∈ [0, 1] tr(P AT ) − tr(P A)tr(P T ) ≤ sup eP ∈S(H) tr(eP AT ) − tr(eP A)tr(ePT ) ≤ dist(A, CId)dist(T, CId) ≤ (2 − λ)(2 − µ)ω(A − λ0Id)ω(T − µ0Id) ≤ (2 − λ)(2 − µ)RART . (3.20) References [1] K. Audenaert, Variance bounds, with an application to norm bounds for commutators, Linear Algebra Appl. 432 (2010), no. 5, 1126 -- 1143. [2] G. Bjorck and V. Thom´ee, A property of bounded normal operators in Hilbert space, Ark. Mat. 4 (1963), 551 -- 555. 10 T. BOTTAZZI, C. CONDE [3] S. Dragomir, Some refinements of Schwarz inequality, Suppozionul de Matematica ¸si Aplica t¸ii, Poly- technical Institute Timi¸soara, Romania, 1-2, (1985), 13-16. [4] L. Gevorgyan, On minimal norm of a linear operator pencil, Dokl. Nats. Akad. Nauk Armen. 110 (2010), no. 2, 97 -- 104. [5] I. Gohberg; M. Krein, Introduction to the Theory of Linear Nonselfadjoint Operators. Translated from the Russian by A. Feinstein. Translations of Mathematical Monographs 18, American Mathematical Society, Providence, R.I. 1969. [6] G. Gruss, Uber 1 (b−a)2 R b a f (x)dxR b das Maximum des absoluten Betrages von a g(x)dx, Math. Z., 39 (1935), 215 -- 226. 1 b−aR b a f (x)g(x)dx − [7] V. Istratescu, O n a class of normaloid operators, Math. Z. 124 (1972), 199 -- 202. [8] K. Paul, Translatable radii of an operator in the direction of another operator, Sci. Math. 2 (1999), no. 1, 119 -- 122. [9] S. Prasanna, The norm of a derivation and the Bjorck-Thome-Istrat¸escu theorem, Math. Japon. 26 (1981), no. 5, 585 -- 588. [10] P. Renaud, A matrix formulation of Gruss inequality, Linear Algebra Appl. 335 (2001), 95 -- 100. [11] J. Stampfli, The norm of a derivation, Pacific J. Math. 33 (1970), 737 -- 747. 1Instituto Argentino de Matem´atica "Alberto P. Calder´on", Saavedra 15 3 piso, (C1083ACA) Buenos Aires, Argentina 2Instituto de Ciencias, Universidad Nacional de Gral. Sarmiento, J. M. Gutierrez 1150, (B1613GSX) Los Polvorines, Argentina E-mail address: [email protected] E-mail address: [email protected]
1808.06496
1
1808
2018-08-20T15:04:16
Frames for the solution of operator equations in Hilbert spaces with fixed dual pairing
[ "math.FA" ]
For the solution of operator equations, Stevenson introduced a definition of frames, where a Hilbert space and its dual are {\em not} identified. This means that the Riesz isomorphism is not used as an identification, which, for example, does not make sense for the Sobolev spaces $H_0^1(\Omega)$ and $H^{-1}(\Omega)$. In this article, we are going to revisit the concept of Stevenson frames and introduce it for Banach spaces. This is equivalent to $\ell^2$-Banach frames. It is known that, if such a system exists, by defining a new inner product and using the Riesz isomorphism, the Banach space is isomorphic to a Hilbert space. In this article, we deal with the contrasting setting, where $\mathcal H$ and $\mathcal H'$ are not identified, and equivalent norms are distinguished, and show that in this setting the investigation of $\ell^2$-Banach frames make sense.
math.FA
math
Frames for the solution of operator equations in Hilbert spaces with fixed dual pairing Peter Balazs⋆ and Helmut Harbrecht† August 21, 2018 ⋆Austrian Academy of Reichsratsstrasse stitute, [email protected], https://orcid.org/0000-0003-4939-0831; 17, Sciences, 1010 In- Pe- https://www.kfs.oeaw.ac.at/balazs, Acoustics Vienna, Research Austria, †University of Basel, Department of Mathematics and Computer Sci- ence, Spiegelgasse 1, 4051 Basel, Switzerland, [email protected], cm.dmi.unibas.ch, https://orcid.org/0000-0003-0093-5706 Abstract For the solution of operator equations, Stevenson introduced a def- inition of frames, where a Hilbert space and its dual are not identified. This means that the Riesz isomorphism is not used as an identifica- tion, which, for example, does not make sense for the Sobolev spaces 0 (Ω) and H −1(Ω). In this article, we are going to revisit the concept H 1 of Stevenson frames and introduce it for Banach spaces. This is equiv- alent to ℓ2-Banach frames. It is known that, if such a system exists, by defining a new inner product and using the Riesz isomorphism, the Banach space is isomorphic to a Hilbert space. In this article, we deal with the contrasting setting, where H and H′ are not identified, and equivalent norms are distinguished, and show that in this setting the investigation of ℓ2-Banach frames make sense. Keywords: Frames, Banach frames, Stevenson frames, matrix repre- sentation, discretization of operators, invertibility. 1 Introduction The standard definition of frames found first in the paper by Duffin and Schaefer [32] is the following: kfkH ≈ khf, ψkiHkℓ2 for all f ∈ H. (1) 1 Here, x ≈ y means that there are constants 0 < A ≤ B < ∞ such that A · x ≤ y ≤ B · x. This concept led to a lot of theoretical work, see e.g. [1, 14, 24, 50, 51], but has been used also extensively in signal processing [16], quantum mechanics [34], acoustics [10] and various other fields. Frames can be used also to represent operators. For the numerical solution of operator equations, the (Petrov-) Galerkin scheme [52] is used, where operators are represented by hOψk, φlik,l∈K, called the stiffness or system matrix. The collection Ψ = (ψk)k∈K consists of the ansatz functions, the collection Φ = (φk)k∈K are the test functions. If Ψ and Φ live in the same space, this is called Galerkin scheme, otherwise it is called Petrov-Galerkin scheme. In finite and boundary element approaches, bases were used [30, 33], but also frames have been applied, e.g. in [37, 40, 41, 48, 53]. Recently, such operator representations got also a more theoretical treatment [6, 9, 12]. In numerical applications, it is often advantageous to have self-adjoint ma- trices, e.g. for Krylov subspace methods, which necessitates to use the same sequence for the discretization at both sides, i.e. investigating hOψk, ψlik,l∈K. Note that this matrix is self-adjoint if O is, and semi-positive if O is. Pos- itivity is in general not preserved, only if a system without redundancy is used, i.e. a Riesz sequence. Partial differential operators are typically oper- ators of the form O : H → H′, while boundary integral operators might also be smoothing operators which map in accordance with O : H′ → H. One possible solution is to work with Gelfand triples i.e, H ⊂ H0 ⊂ H′. This is explicitly done for the concept of Gelfand frames [28]. Another possibility is the following, introduced by Stevenson in [53] and used e.g. in [41]: A collection Ψ = (ψk)k∈K ⊂ H is called a (Stevenson) frame for H, if for all f ∈ H′. (2) kfkH′ ≈(cid:13)(cid:13)(cid:13)hf, ψkiH′,H(cid:13)(cid:13)(cid:13)ℓ2 Note the difference to the definition (1) by Duffin and Schaefer, which is significant only if the Riesz isomorphism is not employed. Here, the Gelfand triple is only implicitly used and, if the fully general setting is used, the density of the spaces is not required. Clearly, the definitions (1) and (2) are equivalent by the Riesz isomor- phism. On the other hand, if the isomorphism H ∼= H′ is not considered, but another one is utilized, for example, considering the triple H ⊂ H0 ⊂ H′, then the Riesz isomorphism is usually used as an identification on the pivot space H0 ∼= H′ 0 , and therefore H and H′ cannot be considered to be equal. In this article, we consider the original definition by Stevenson and re- investigate in full detail all the derivation to ensure that the Riesz identifi- 2 cation does not 'creep in' again. On a more theoretical level, let us consider Banach frames [22, 25, 38]. Thus, we consider a Banach space X, a sequence space Xd, and a sequence Ψ ⊂ X ′. This is a Xd-frame if kfkX ≈(cid:13)(cid:13)(cid:13)hf, ψkiX,X ′(cid:13)(cid:13)(cid:13)Xd for all f ∈ X. It is called a Banach frame if a reconstruction operator exists, i.e. there exists R : Xd → X with R (ψk(f )) = f for all f ∈ X. In this setting, ℓ2-frames were not considered to be interesting as they are isomorphic to Hilbert frames, see e.g. [55, Proposition 3.10]: Let Ψ be a ℓ2-frame for X. Then, X can be equipped with an inner product hf, giX = hCΨf, CΨgiℓ2, becoming a Hilbert space, and Ψ is a (Hilbert) frame for X. The proof uses the Riesz isomorphism H ∼= H′ in the last line. But if a context is considered, where this isomorphism cannot be applied, like for example a Gelfand triple setting, suddenly the concept of ℓ2-frames might become non-trivial again, and the concept of Stevenson frames is different to a frame. In this article, we investigate this approach. The rest of this article is structured as follows. In Section 2, we moti- vate Gelfand triples H′ ⊂ H0 ⊂ H by a simple example arising from the variational formulation of second order elliptic partial differential equations. Section 3 then provides the main ingredients we need, especially it introduces the different notions of frames for solving operator equations. By an illustra- tive example, we show that Stevenson frames seem to offer the most flexible concept for the discretization of operator equations. Finally, in Section 4, we generalize Stevenson frames to Banach spaces and discuss the consequences. 2 Motivation: Solving Operator Equations Let O : H → H′ and define the bilinear form a : H × H → R by a(u, v) = hOu, vi. Assume that a satisfies the following properties: 1. Let a be bounded, i.e. there is a constant CS, such that a(u, v) ≤ CS · kukH kvkH . This is equivalent to O being bounded. 2. Let a be elliptic, i.e. there exists a constant CE such that a(u, u) ≥ CE · kuk2 H . 3 Both conditions are equivalent to O being bounded, boundedly invertible, and positive, see e.g. [18, 20]. The general goal is to find the solution u ∈ H such that a(u, v) = ℓ(v) for all v ∈ H. (3) This is the weak formulation of the operator equation Ou = b, setting ℓ(v) = hb, viH′,H for u ∈ H and b ∈ H′. In numerical approximation schemes, to get an approximate solution, finite dimensional subspaces V ⊂ H are considered and the solution uV ∈ V such that a(uV , v) = ℓ(v) for all v ∈ V is calculated. The error between the continuous solution u ∈ H and the approximate solution uV ∈ V is orthogonal to the space V , which is known as the Galerkin orthogonality: a(u − uV , v) = 0 for all v ∈ V . Note that, in difference to e.g. a Gelfand triple approach, the norms on V and H are the same in the setting above. Instead, the Gelfand triple setting would be H ⊂ H0 ⊂ H′ with k·kH0 ≤ c k·kH. We shall illustrate the setting also by a practical example from the theory of partial differential equations. To that end, assume that Ω is a bounded domain in Rd and let H0 := L2(Ω) be the space of all square-integrable functions v : Ω → R. As space H ⊂ H0 we consider the Sobolov space H 1 0 (Ω) which consists of all functions in L2(Ω) whose first order week derivatives are also square-integrable and which are zero at the boundary ∂Ω. Thus, the variational formulation of the Poisson equation reads −∆u = f in Ω, u = 0 on ∂Ω 0 (Ω), (4) a : H 1 0 (Ω) × H 1 0 (Ω) → R, seek u ∈ H 1 compare [18] for example. The bilinear form 0 (Ω) such that a(u, v) = ℓ(v) for all v ∈ H 1 a(u, v) =ZΩ ∇u∇vx. ℓ(v) =ZΩ is continuous provided that f ∈ H −1(Ω) = (cid:0)H 1 0 (Ω) → R, is continuous and elliptic due to Friedrichs' inequality, cf. [18], and the linear form ℓ : H 1 f vx. product in the pivot space L2(Ω) is continuously extended onto the duality pairing H −1(Ω) × H 1 0 (Ω) ⊂ L2(Ω) ⊂ H −1(Ω). 0 (Ω). Hence, the underlying Gelfant triple is H 1 0 (Ω)(cid:1)′. Hereby, the inner 4 3 Main Definitions and Notations 3.1 Dual Pairs Let X, Y be vector spaces and a(x, y) a bilinear functional on X × Y . Then (X, Y ) is called a dual pair [56], if 1. ∀x ∈ X\{0} ∃y ∈ Y s.t. a(x, y) 6= 0, 2. ∀y ∈ Y \{0} ∃x ∈ X s.t. a(x, y) 6= 0. In short, the notation a(x, y) = hx, yia = hx, yi is used. A classical example is a Banach space X and its dual space X ′. But looking at other dual pairs allows to have an explicit form for the dual elements [42]. Note that often an isomorphism is considered as an identity. For example, by using the Riesz mapping H ∼= H′, the dual space H′ is often identified with H. If two or more isomorphisms are involved, this identification, of course, can only be considered for one of those isomorphisms. For example, if we consider two Hilbert spaces H1 ⊂ H2, the Riesz isomorphism can be considered only for one of them to be an identification, see also Section 3.3.2. 3.2 Gelfand Triples Let X be a Banach space and H a Hilbert space. Then, the triple (X,H, X ′) is called a Banach Gelfand triple [26], if X ⊂ H ⊂ X ′, where X is dense in H, and H is w⋆-dense in X ′. The prototype of such a triple is (ℓ1, ℓ2, ℓ∞) in case of sequence spaces. Note that, even if we consider the spaces all being Hilbert spaces -- such a sequence is also called rigged Hilbert spaces [2] -- the Riesz isomorphism, in general, is not just the composition of the inclusion with its adjoint. This depends on the chosen concrete dual pairing. As another example, consider the triple H 1 0 (Ω) ⊂ L2(Ω) ⊂ H −1(Ω), which has been presented in the practical example for the Poisson equation in Sec- tion 2. 3.3 Frames A sequence Ψ = (ψk)k∈K in a separable Hilbert space H is a frame for H, if there exist positive constants AΨ and BΨ (called lower and upper frame bound, respectively) that satisfy AΨkfk2 ≤Xk∈K hf, ψki2 ≤ BΨkfk2 for all f ∈ H. (5) 5 An upper (resp. lower) semi-frame is a complete system that only satisfies the upper (resp. lower) frame inequality, see [3, 4]. A frame where the two bounds can be chosen to be equal, i.e. AΨ = BΨ, is called tight. We will denote the corresponding sequences in H by Ψ = (ψk)k∈K and Φ = (φk)k∈K in the following, where we consider general discrete index sets K ⊂ Rd. A sequence that is a frame for its closed linear span is called a frame sequence. By CΨ : H → ℓ2 we denote the analysis operator defined by (CΨf )k = hf, ψki. The adjoint of CΨ is the synthesis operator DΨ(ck) = Pk ckψk. The frame operator SΨ = DΨCΨ can be written as SΨf = Pk hf, ψki ψk. It is positive and invertible. Note that those 'frame-related' operators can be defined as possibly unbounded operators for any sequence in the Hilbert space [13]. By using the canonical dual frame ( ψk), i.e., ψk = S−1 Ψ ψk for all k, we get a reconstruction formula: f =Xk hf, ψki ψk =Xk Df, ψkE ψk for all f ∈ H. The Gramian matrix GΨ is defined by (GΨ)k,l = hψl, ψki, also called the mass matrix. This matrix defines an operator on ℓ2 by matrix multiplication, corresponding to GΨ = CΨDΨ. Similarily, we can define the cross-Gramian matrix (GΨ,Φ)k,l = hφl, ψki between two different frames Φ and Ψ. Clearly, GΨ,Φc =Xl clφl, ψk+ = CΨDΦc. If, for the sequence Ψ, there exist constants AΨ, BΨ > 0 such that the inequalities (GΨ,Φ)k,l cl =*Xl ckψk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xk∈K 2 H AΨ kck2 ≤ BΨ kck2 2 are fulfilled, Ψ is called a Riesz sequence. If Ψ is complete, it is called a Riesz basis. 3.3.1 Banach Frames The concept of frames can be extended to Banach spaces [22, 25, 38]: Let X be a Banach space and Xd be a Banach space of scalar sequences. A sequence (ψk) in the dual X ⋆ is called an Xd-frame for the Banach space X (1 < p < ∞), if there exist constants AΨ, BΨ > 0 such that AΨkfkX ≤ kψk(f )kXd ≤ BΨkfkX for all f ∈ X. 6 An Xd-frame is called a Banach frame with respect to a sequence space Xd, if there exists a bounded reconstruction operator R : Xd → X, such that R (ψk(f )) = f for all f ∈ X. In our setting, we use p-frames, that is Xd = ℓp for 1 ≤ p ≤ ∞, especially, we use Xd = ℓ2. A family (gk)k∈K ⊂ X is called a q-Riesz sequence (1 ≤ q ≤ ∞) for X, if there exist constants AΨ, BΨ > 0 such that AΨ Xk∈K q dkq! 1 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xk∈K dkgk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ≤ BΨ Xk∈K q dkq! 1 (6) for all finite scalar sequence (dk). The family is called a q-Riesz basis if it fulfills (6) and span{gk : k ∈ K} = X. [25]. Any q-Riesz basis for X ⋆ is a p-frame for X, where 1 q = 1, compare p + 1 3.3.2 Gelfand Frames A frame for H is called a Gelfand frame [28] for the Gelfand triple (X,H, X ′) if there exists a Gelfand triple of sequence spaces (Xd, ℓ2, X ′ d), such that the synthesis operator DΨ : Xd → X and the analysis operator C eΨ : X → Xd are bounded. As a result, see [28, 45], this means that Ψ is a Banach frame for Xd and eΨ a Banach frame for X ′ d. B is an isomorphism. B = diag(cid:16) 1 wk(cid:17), then Ψ =(cid:0) 1 In many approaches, see e.g. [28], it is assumed for the implementation that there exists an isomorphism DB : Xd → ℓ2. Should Xd be non-reflexive, then it is also assumed that D⋆ If DB is a diagonal operator, i.e. DB = diag (wk) and D−1 real weights in [57]. It is easy to see also for complex weights when using a weighted frame viewpoint [7, 54]. These cases cover the weighted spaces ℓ2 w. The above setting can be generalized as follows: We define, similar to ψk(cid:1) is a Hilbert frame for X and(cid:0)wk ψk(cid:1) is a Hilbert frame for X ′. This is shown for X := (cid:10)DBC eΨf, DBC eΨg(cid:11)ℓ2. It is obviously [55], the sesqui-linear form hf, gio bounded and coercive, and, in particular, kfkX o :=phf, fiX o is equivalent to kfkX . Therefore, (X,kfkX o) is a Hilbert space which is isomorphic to (X,kfkX). Now let ξl := DΨD−1 B δl, where δl is the standard basis in ℓ2. This B)−1 δl is a Hilbert is a Hilbert space frame for X. Similarly, ηl := D eΨ (D⋆ space frame for X ′. As a consequence X and X ′ are Hilbert spaces, but X 6= X ′ and the inner products and the corresponding norms are changed, albeit equivalent to the original ones. wk 7 H′ In AΨ · kfk2 3.4 Stevenson Frames We consider the duality (H,H′) without using the Riesz isomorphism. particular, we use the duality with respect to a second Hilbert space H0. Definition 3.1 ([53]) A sequence Ψ = (ψk)k∈K ⊂ H is called a (Stevenson) frame for H if there exists constants 0 < AΨ ≤ BΨ < ∞ such that for all f ∈ H′. Different to the Gelfand frames setting, we do not assume density. H′ ≤(cid:13)(cid:13)(cid:13)hf, ψkiH′,H(cid:13)(cid:13)(cid:13)ℓ2 ≤ BΨ · kfk2 Typically, we consider Sobolev spaces and the L2-inner product, which we can consider as co-orbit spaces with the sequence spaces ℓ2 w varying w. Here, invertible operators between different spaces exist, see Section 3.3.2, and density is also given. In this article, we treat the most general setting. In [53], the author states "We adapted the definition of a frame given in [31, Section 3] by identifying H with its dual H′ via the Riesz mapping". Then, the following results are stated, also in [41], without proofs: The analysis operator CΨ : H′ → ℓ2, CΨ(f ) = (hf, ψki)k∈K is bounded by (7), as is its adjoint C ⋆ Ψ = DΨ is the Ψ : ℓ2 → H. It can be easily shown that C ⋆ synthesis operator with DΨc =Pk∈K ckψk. Especially, one has (7) ℓ2 = ran (CΨ) ⊕ ker (DΨ) . 1 BΨ Ψ and D eΨ = S−1 We have the reconstructions CΨS−1 S eΨ : H → H′. Define the frame operator SΨ = DΨCΨ. It is a mapping SΨ : H′ → . Here, C eΨ = Ψ and, therefore, and 1 AΨ Ψ DΨ. Furthermore, it holds S eΨ = S−1 H, which is boundedly invertible. We can show that the sequence eΨ = (cid:0)S−1 Ψ ψk(cid:1)k∈K is a (Stevenson) H-frame with bounds f = DΨC eΨh =Xk∈KDf, ψkEH,H′ h = D eΨCΨh =Xk∈K hh, ψkiH′,H ψk, and ψk, (8) (9) for all f ∈ H and h ∈ H′. The cross-Gramian matrix GΨ, eΨ = DΨC eΨ is the orthogonal projection on ran (CΨ) and coincides with G eΨ,Ψ. Therefore, ran (CΨ) = ran(cid:0)C eΨ(cid:1). In this article, we are revisiting those statements, make them slightly more general, in order to make sure that not using the Riesz isomorphism is possible. 8 3.5 An Illustrative Example Let Ω ⊂ Rn be a sufficiently smooth, bounded domain. We consider a multi- scale analysis, i.e., a dense, nested sequence of finite dimensional subspaces V0 ⊂ V1 ⊂ · · · ⊂ Vj ⊂ · · · ⊂ L2(Ω), consisting of piecewise polynomial ansatz functions Vj = span{ϕj,k : k ∈ ∆j}, such that dim Vj ∼ 2jn and L2(Ω) = [j∈N0 Vj, V0 = \j∈N0 Vj. One might think here of a multigrid decomposition of standard Lagrangian finite element spaces or of a sequence of spline spaces originating from dyadic subdivision. Trial spaces Vj which are used for the Galerkin method satisfy typically a direct or Jackson estimate. This means that kv − PjvkL2(Ω) ≤ CJ 2−jqkvkH q(Ω), (10) holds for all 0 ≤ q ≤ d uniformly in j. Here, Pj : L2(Ω) → Vj is the L2(Ω)- orthogonal projection onto the trial space Vj and H q(Ω) ⊂ L2(Ω), q ≥ 0, denotes the Sobolev space of order q. The upper bound d > 0 refers in general to the maximum order of the polynomials which can be represented in Vj, while the factor 2−j refers to the mesh size of Vj, i.e., the diameter of the finite elements, compare [18] for example. v ∈ H q(Ω), Besides the Jackson type estimate (10), there also holds the inverse or Bernstein estimate (cid:13)(cid:13)Pjv(cid:13)(cid:13)H q(Ω) ≤ CB2jq(cid:13)(cid:13)Pjv(cid:13)(cid:13)L2(Ω), for all 0 ≤ q < γ, where the upper bound v ∈ H q(Ω), (11) γ := sup(cid:8)t ∈ R : Vj ⊂ H t(Ω)(cid:9) > 0 refers to the regularity of the functions in the trial spaces Vj. There holds γ = d − 1/2 for trial functions based on cardinal B-splines, since they are globally C d−1-smooth, and γ = 3/2 for standard Lagrangian finite element shape functions, since they are only globally continuous. A crucial requirement is the uniform frame stability of the systems under consideration, i.e., the existence of constants AΦ, BΦ > 0 such that L2(Ω) for all f ∈ L2(Ω) AΦ(cid:13)(cid:13)Pjf(cid:13)(cid:13)2 L2(Ω) ≤ Xk∈∆j hf, ϕj,ki2 ≤ BΦ(cid:13)(cid:13)Pjf(cid:13)(cid:13)2 (12) 9 holds uniformly for all j. This stability is satisfied for example by Lagrangian finite element basis functions defined on a multigrid hierarchy resulting from uniform refinement of a given coarse grid, see [18] for example. It is also satisfied by B-splines defined on a dyadic subdivision of the domain under consideration. Having a multiscale analysis at hand, it can be used for telescoping a given function to account for the fact that Sobolev norms act different on different length scales. Namely, the interplay of (10) and (11) gives rise to the norm equivalence eH −q(Ω) ∼Xj∈N0 kfk2 2−2jq(cid:13)(cid:13)(Pj − Pj−1)f(cid:13)(cid:13)2 for all 0 ≤ q < γ, where P−1 := 0 and eH −q(Ω) :=(cid:0)H q(Ω)(cid:1)′ In accordance with [41], using (12), we can estimate to H q(Ω), see [29] for a proof. L2(Ω) (13) denotes the dual Xj∈N0Xk∈∆j L2(Ω) 2−2jq(cid:13)(cid:13)Pjf(cid:13)(cid:13)2 2−2jqhf, ϕj,ki2 ≈Xj∈N0 = Xj∈N0 =Xℓ∈N0(cid:13)(cid:13)(Pℓ − Pℓ−1)f(cid:13)(cid:13)2 2−2jq L2(Ω) 2−2jq. L2(Ω) jXℓ=0(cid:13)(cid:13)(Pℓ − Pℓ−1)f(cid:13)(cid:13)2 ∞Xj=ℓ 2−2ℓq(cid:13)(cid:13)(Pℓ − Pℓ−1)f(cid:13)(cid:13)2 L2(Ω). The latter sum converges provided that q > 0 and we arrive at Xj∈N0 Xk∈∆j 2−2jqhf, ϕj,ki2 ≈Xℓ∈N0 In view of the norm equivalence (13), we have thus proven that there exist constants AΦ, BΦ > 0 such that AΦkfk2 eH −q(Ω) ≤Xj∈N0Xk∈∆j 2−2jqhf, ϕj,ki2 ≤ BΦkfk2 eH −q(Ω) (14) for all 0 < q < γ. Therefore, in accordance with Definition 3.1, the collection Φ = {2−jqϕj,k : k ∈ ∆j, j ∈ N0} (15) defines a Stevenson frame for H = H q(Ω), where H′ = eH −q(Ω) with duality related to H0 = L2(Ω). Notice that this frame underlies the construction of 10 the so-called BPX preconditioner, see e.g. [19, 29, 47]. Especially, by remov- ing all basis functions which are associated with boundary nodes, one gets a Stevenson frame for H = H 1 0 (Ω), as required for the Galerkin discretization of elliptic partial differential equations, compare Section 2. We like to emphazise that the collection (15) does not define a Gelfand frame, since (14) does not hold in H0 = L2(Ω), i.e., for q = 0. Hence, the concept of Stevenson frames seems to be more flexible than the concept of Gelfand frames. 3.6 Operator Representation in Frame Coordinates For orthonormal sequences, it is well known that operators can be uniquely described by a matrix representation [36]. The same can be constructed with frames and their duals, see [6, 9]. Let Ψ = (ψk) be a frame in H1 with bounds AΨ, BΨ > 0, and let Φ = (φk) be a frame in H2 with AΦ, BΦ > 0. 1. Let O : H1 → H2 be a bounded, linear operator. Thus, the infinite matrix (cid:0)M(Φ,Ψ) (O)(cid:1)m,n = hOψn, φmi defines a bounded operator from ℓ2 to ℓ2 with kMkℓ2→ℓ2 ≤ √BΦ · BΨ · kOkH1→H2. As an operator ℓ2 → ℓ2, we have M(Φ,Ψ) (O) = CΦ ◦ O ◦ DΨ. 2. On the other hand, let M be an infinite matrix defining a bounded operator from ℓ2 to ℓ2, (Mc)i =Pk Mi,kck. Then, the operator O(Φ,Ψ) defined by (cid:0)O(Φ,Ψ) (M)(cid:1) h =Xk Xj Mk,j hh, ψji! φk for all h ∈ H1 is a bounded operator from H1 to H2 with and (cid:13)(cid:13)O(Φ,Ψ) (M)(cid:13)(cid:13)H1→H2 ≤pBΦ · BΨ kMkℓ2→ℓ2 O(Φ,Ψ)(M) = DΦ ◦ M ◦ CΨ =Xk Xj Mk,j · φk ⊗i ψj. 11 Please note that there is a classification of matrices that are bounded operators from ℓ2 to ℓ2 [27]. If we start out with frames, more properties can be proved [6]: Let Ψ = (ψk) be a frame in H1 with bounds AΨ, BΨ > 0, Φ = (φk) in H2 with AΦ, BΦ > 0. 1. It holds (cid:16)O(Φ,Ψ) ◦ M ( Φ, Ψ)(cid:17) = idB(H1,H2) =(cid:16)O( Φ, Ψ) ◦ M (Φ,Ψ)(cid:17) . Therefore, for all O ∈ B(H1,H2): O =Xk,j DO ψj, φkE φk ⊗i ψj. 2. M(Φ,Ψ) is injective and O(Φ,Ψ) is surjective. 3. If H1 = H2, then O(Ψ, Ψ)(idℓ2) = idH1. 4. Let Ξ = (ξk) be any frame in H3, and O : H3 → H2 and P : H1 → H3. Then, it holds M(Φ,Ψ) (O ◦ P ) =(cid:16)M(Φ,Ξ) (O) · M(Ξ,Ψ) (P )(cid:17) . Note that, in the Hilbert space of Hilbert-Schmidt operators, the tensor product Ψ⊗ Φ := {ψk ⊗ ψl}(k,l)∈K×K is a Bessel sequence / frame sequence / Riesz sequence, if the starting sequences Ψ and Φ are [5], with M(Φ,Ψ) being the analysis and O(Φ,Ψ) being the synthesis operator. This relation is even an equivalence [17]. For the invertibility, it can be shown [11, 12]: If and only if O is bijective, then M = M(Φ,Ψ)(O) is bijective as operator from ran (CΨ) onto ran (CΦ). In this case, one has M † = M( Ψ, Φ)(O−1) = G Ψ, Φ ◦ M(Φ,Ψ)(cid:0)O−1(cid:1) G Ψ, Φ = M(Ψ,Φ)(cid:0)S−1 If we have an operator equation Ou = b, we use Ψ O−1S−1 Φ (cid:1) . which implies Ou = b ⇐⇒Xk Du, ψkE Oψk = b, Xk Du, ψkE hOψk, ψli = hb, ψli 12 for all l ∈ K. Setting M = M(Ψ,Ψ)(O), ~u = C eΨu and ~b = CΨb, we thus have Ou = b ⇐⇒ M~u = ~b. Note that, for numerical computations, see e.g. [28, 53], the system of linear equations M~u = ~b is solved. Then, u = DΨ~u is the solution to Ou = b, avoiding the numerically expensive calculation of a dual frame [8, 44, 49]. If the frame is redundant, then uk can be different to (cid:10)u, ψk(cid:11). If a Tychonov regularization is used, we obtain uk =(cid:10)u, ψk(cid:11) by [39, Prop. 5.1.4]. 4 Stevenson Frames Revisited As some of the references dealing with Stevenson frames used an unlucky formulation, when stating if or if not the Riesz isomorphism is used, see e.g. [28, 53], the authors decided to check everything again, and pay particular attention to the avoidance of the Riesz isomorphism, i.e. to not use H ∼= H′. To not use the Riesz isomorphism in a treatment of Hilbert spaces is mind-boggling, so we decided to use Banach spaces, to be sure to avoid all (Note, however, that the Riesz isomorphism will be used on the pitfalls. sequence space ℓ2.) In particular, this is a generalization of the original definition. The used spaces are necessarily isomorphic to Hilbert spaces, but not Hilbert spaces per se. 4.1 Stevenson Banach Frames We start out with a generalized definition. (We will show that this is iso- morphic, but not identical to the original definition.) Definition 4.1 Let (X, X ′) be a dual pair of reflexive Banach spaces. Let Ψ = (ψk)k∈K ⊂ X. It is called a Stevenson Banach frame for X, if there exist bounds 0 < AΨ ≤ BΨ < ∞ such that AΨ kfk2 The analysis operator X ′ X ′ ≤(cid:13)(cid:13)(cid:13)hf, ψkiX ′,X(cid:13)(cid:13)(cid:13)ℓ2 ≤ BΨ kfk2 CΨ : X ′ → ℓ2, CΨ(f ) =(cid:16)hf, ψkiX ′,X(cid:17)k∈K for all f ∈ X ′. is bounded by √BΨ by definition. (Note that we use here the notation which is more common for Banach spaces [38].) As a consequence of the open mapping theorem, CΨ is one-to-one and has closed range. 13 For d = (dk) ∈ ℓ2(K) with finitely many non-zero entries, i.e. d ∈ c00, consider hCΨf, diℓ2 =Xk∈K hf, ψkiX ′,X dk =*f,Xk∈K dkψk+X ′,X . By using a standard density argument and the reflexivity, it can easily be Ψ = DΨ, where DΨ : ℓ2 → X is the synthesis operator with shown that C ⋆ unconditionally. Indeed, consider c ∈ ℓ2. Then, let K0 ⊂ K be a finite set, such that DΨc = Pk∈K ckψk. The bound of DΨ is also √BΨ. The sum converges ǫ √BΨ For another finite index set K1 ⊃ K0, we thus find ck2 < ǫ′ := . Xk6∈K0 ckψk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H ckψk − Xk6∈K1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xk6∈K = kDΨ (c − c · χK1)kH <pBΨǫ′ = ǫ. Hence, by e.g. [56, IV.5.1] and the fact that ℓ2 is a Hilbert space, we deduce ℓ2 = ran (CΨ) ⊕ ker (DΨ) . (16) We define the frame operator SΨ = DΨCΨ, which is a mapping SΨ : X ′ → X. In particular, the operator SΨ is self-adjoint. By definition of SΨ, it follows that hSΨf, giX,X ′ ≤ kSΨfkX kgkX ′ ≤ BΨ · kfkX ′ · kgkX ′ . (17) Hence, SΨ is bounded with bound BΨ. Furthermore, we have ΨCΨf, fiX,X ′ = hCΨf, CΨfiℓ2 = kCΨfk2 hSΨf, fiX,X ′ = hC ⋆ X ′ , (18) which implies that SΨ is one-to-one and positive. By [56, IV.5.1], this also means that S⋆ Ψ = SΨ has dense range. SΨ also has a bounded inverse since ℓ2 ≥ AΨ · kfk2 kSΨfkX = sup kgkX ′ =1 g∈X ′ hSΨf, giX,X ′ ≥(cid:28)SΨf, f kfkX ′(cid:29)X,X ′ ≥ AΨ · kfkX ′ . Therefore, it has closed range [35, Theorem XI.2.1]. Consequently, SΨ is onto and bijective with Thus, S−1 Ψ is also self-adjoint, and AΨ kfkX ′ ≤ kSΨfkX ≤ BΨ kfkX ′ . 1 BΨ kgkX ≤(cid:13)(cid:13)S−1 Ψ g(cid:13)(cid:13)X ′ ≤ 14 1 AΨ kgkX . (19) son Banach frame for X ′ with bounds 1 BΨ Theorem 4.1 The sequence eΨ =(cid:0) ψk(cid:1)k∈K :=(cid:0)S−1 Ψ ψk(cid:1)k∈K ⊂ X ′ is a Steven- operator coincides with the one of the primal frame, i.e. ran (CΨ) = ran(cid:0)C eΨ(cid:1). The related operators are C eΨ = CΨS−1 f ∈ X and g ∈ X ′, we have the reconstructions Ψ DΨ and S eΨ = S−1 . The range of its analysis Ψ , D eΨ = S−1 and 1 AΨ Ψ . For Proof: Ψ ψk ∈ X ′. Moreover, we have on the one hand hg, ψkiX ′,X ψk. 2 =Xk∈K(cid:12)(cid:12)(cid:12)(cid:10)S−1 Ψ f, ψk(cid:11)X ′,X(cid:12)(cid:12)(cid:12) X ′ ≥ AΨ Ψ kfk2 X ′ . B2 and on the other hand BΨ Ψ kfk2 A2 X 2 2 2 ψk Xk∈K(cid:12)(cid:12)(cid:12)(cid:12)Df, ψkEX,X ′(cid:12)(cid:12)(cid:12)(cid:12) X ′ ≤ It obviously holds S−1 f =Xk∈KDf, ψkEX,X ′ and g =Xk∈K =Xk∈K(cid:12)(cid:12)(cid:12)(cid:10)f, S−1 Ψ ψk(cid:11)X,X ′(cid:12)(cid:12)(cid:12) Ψ f(cid:13)(cid:13)2 ≤ BΨ(cid:13)(cid:13)S−1 Xk∈K(cid:12)(cid:12)(cid:12)(cid:12)Df, ψkEX,X ′(cid:12)(cid:12)(cid:12)(cid:12) ≥ AΨ(cid:13)(cid:13)S−1 Ψ f(cid:13)(cid:13)2 Ψ ψk(cid:11)X,X ′ =(cid:10)SΨg, S−1 S eΨf =Xk (cid:10)f, S−1 Ψ Xk (cid:10)S−1 Ψ ψk(cid:11)X,X ′ S−1 Ψ f, ψk(cid:11)X,X ′ ψk = S−1 = S−1 Ψ ψk we get Hence, eΨ is an X ′-frame. By employing the invertibility of SΨ for g = S−1 (cid:10)f, S−1 Ψ ψk(cid:11)X ′,X = hg, ψkiX ′,X . This implies ran (CΨ) = ran(cid:0)C eΨ(cid:1), where C eΨ = CΨS−1 Ψ ψk(cid:11)X,X ′ =(cid:10)g, SΨS−1 Ψ : X → X ′, because it holds Furthermore, S eΨ = S−1 Ψ and D eΨ = S−1 Ψ DΨ. Ψ f , Ψ SΨS−1 Ψ f = S−1 Ψ f for all f ∈ X. Finally, we have the reconstructions f = DΨCΨS−1 Ψ f = DΨC eΨf =Xk∈KDf, ψkEX,X ′ ψk for all f ∈ X and g = S−1 Ψ DΨCΨ = D eΨCΨg =Xk∈K hh, ψkiX ′,X ψk 15 Ψ ·,·(cid:11) defines for all g ∈ X ′. As a positive sesqui-linear form, the Cauchy-Schwarz inequality implies 1 AΨ kxk2 X , Thus, with x = SΨu, there holds we have the sharper upper bound. On the other hand, since(cid:10)S−1 Ψ y, y(cid:11)X ′,X . Ψ y, y(cid:11)X ′,X Ψ y, y(cid:11)X ′,X . Ψ x, x(cid:11)X ′,X ≤(cid:13)(cid:13)S−1 Ψ x(cid:13)(cid:13)X ′ kxkX ≤ (cid:10)S−1 (cid:12)(cid:12)(cid:12)(cid:10)S−1 Ψ x, y(cid:11)X ′,X(cid:12)(cid:12)(cid:12) ≤(cid:10)S−1 Ψ x, x(cid:11)X ′,X(cid:10)S−1 (cid:12)(cid:12)(cid:12)hu, yiX ′,X(cid:12)(cid:12)(cid:12) ≤ hu, SΨuiX ′,X(cid:10)S−1 (cid:12)(cid:12)(cid:12)hu, yiX ′,X(cid:12)(cid:12)(cid:12) ≤ BΨ(cid:10)S−1 and consequently X = sup kukX ′ =1 u∈X ′ kyk2 2 2 2 So, the sharper bounds 1 BΨ and 1 AΨ follow. ✷ The fact that ran (CΨ) = ran(cid:0)C eΨ(cid:1) is very different to the Gelfand frame setting, where the ranges ran(cid:0)CΨX ′(cid:1) 6= ran(cid:16)C eΨX(cid:17) even live in different sequence spaces. Theorem 4.2 The cross-Gramian matrix GΨ, eΨ = CΨD eΨ is the orthogonal projection on ran (CΨ) and coincides with G eΨ,Ψ. Proof: We have that the cross-Gramian matrix of a frame and its dual is a projection: (cid:16)GΨ, eΨ(cid:17)2 = CΨD eΨCΨD eΨ = CΨD eΨ. = GΨ, eΨ. = C eΨDΨ = G eΨ,Ψ. Next, it holds In addition, since (cid:16)GΨ, eΨ(cid:17)k,l =D ψl, ψkEX ′,X G⋆ Ψ, eΨ =(cid:0)CΨD eΨ(cid:1)⋆ =(cid:10)S−1 we conclude GΨ, eΨ = G eΨ,Ψ. Thus, GΨ, eΨ is self-adjoint. Ψ ψl, ψk(cid:11)X ′,X =(cid:10)ψl, S−1 Ψ ψk(cid:11)X,X ′ =(cid:16)G eΨ,Ψ(cid:17)k,l , ✷ 16 Theorem 4.3 The collection Ψ is a Stevenson Banach frame for X with bounds AΨ and BΨ if and only if 1 BΨ kfkX ≤ inf d∈ℓ2, DΨd=f kdkℓ2 ≤ 1 AΨ kfkX . In particular, for any f ∈ X with f = Pk∈K dkψk and d = (dk) ∈ ℓ2, we have kdkℓ2 ≥(cid:13)(cid:13)C eΨf(cid:13)(cid:13)ℓ2. Proof: Given dkψk ∈ X, we have the representation f =Xk∈K f =Xk∈KDf, ψkEX,X ′ (cid:18)dk −Df, ψkEX,X ′(cid:19) ∈ ker (DΨ) . By (16) and Theorem 4.1, there follows kdkℓ2 ≥(cid:13)(cid:13)C eΨf(cid:13)(cid:13)ℓ2. Hence, ψk. ✷ Consequently, a Stevenson frame is a Riesz basis for X if and only if DΨ is one-to-one. H Is X′ a Hilbert Space? 4.2 := hu, SΨviX ′,X. This is, trivially, a symmetric and positive Set hu, viX ′ bilinear form by above and, therefore, an inner product on X ′. Hence, X ′ is a pre-Hilbert space with this inner product. By (17) and (18), the corre- H ≤ sponding norm is equivalent to the original one as √AΨ kfkX ′ ≤ kfkX ′ √BΨ kfkX ′. Thus, (X ′,h·,·i) is a Hilbert space. Note that, in particular for numerics, it is sometimes not enough to con- sider equivalent norms. While well-posed problems stay well-posed for equiv- alent norms, this becomes important for concrete implementations, as things like condition numbers, constants in convergence rates, etc. are considered. From a frame theory perspective, switching to an equivalent norm can destroy or create tightness, in particular, the switch from one norm to the other changes the frame bound ratio BΨ . We refer the reader to e.g. weighted AΨ and controlled frames [7], which are under very mild conditions equivalent to classical Hilbert frames. Nonetheless, they have applications for example in 17 the implementation of wavelets on the sphere [15, 43], and nowadays become important for the scaling of frames [23, 46]. As a trivial example, look at Ψ := {e1, e1, e2, e2, e3, e3, . . .}, where E = {ei}i∈N is an orthonormal basis for H. Then, Ψ is a tight frame with AΨ = 2. Looking at the reweighted version Φ := {2e1, 2e1, e2/2, e2/2, 2e3, 2e3, . . .}, we loose tightness, since this frame has bounds A = 1 and B = 4. Note that there exists an invertible bounded operator that maps the single elements from Ψ into Φ, i.e they are equivalent sequences [21]. Also note that, if it does not make sense to assume that X ⊂ X ′, then Ψ cannot be a Hilbert space frame per se. This can only be true for Ψ′ := k) = (Iψk), where I is an isomorphism from X to X ′, for example, choosing (ψ′ I = S−1 Ψ . In this case, the frame bounds are preserved, but the roles of primal and dual frames interchange. This especially means that, if the frame bound ratio is important, dis- tinguishing ℓ2-Banach frames from Hilbert frames is necessary, especially if concrete examples for X and X ′ are used, where an identification is not possible, i.e. X 6= X ′. As such, Definition 4.1 is, of course, equivalent to the standard frame definition for Hilbert spaces, but the frame bound ratio changes. We like to remark that, by using the dual frame, one can also conclude that X itself is a Hilbert space. 4.3 Matrix Representation Let us also revisit the statements about the matrix representation of opera- tors [41, 53]. To this end, let Ψ be Stevenson Banach frame for X. Let us now consider an operator O : X → X ′ and define (As in Section 3.6, we could consider different sequences, and the arguments would still work, but following the argument in the introduction and for easy reading we will not.) For an invertible operator O, we have M( eΨ)(O−1)M(Ψ)(O) = C eΨO−1D eΨCΨODΨ = G eΨ,Ψ. (For the analogue result in the Hilbert frame case, see [11, 12].) Equivalently, M(Ψ)(O)M( eΨ)(O−1) = G eΨ,Ψ. 18 Then, M(Ψ) (O) = CΨODΨ, which implies (cid:0)M(Ψ) (O)(cid:1)m,n = hOψn, ψmiX ′,X . (cid:13)(cid:13)M(Ψ) (O)(cid:13)(cid:13)ℓ2→ℓ2 ≤ BΨ kOkX→X ′ . Therefore, as G eΨ,Ψ is the orthogonal projection on ran (CΨ) the operator M(Ψ)(O)ran(CΨ) is boundedly invertible, as Furthermore, (cid:13)(cid:13)M(Ψ)(O)(cid:13)(cid:13)ran(CΨ)→ran(CΨ) ≥ AΨ(cid:13)(cid:13)O−1(cid:13)(cid:13)−1 ker(cid:16)M( eΨ)(O)(cid:17) = ker (DΨ) . X ′→X . If O is symmetric, then M(Ψ) (O) is symmetric. If O is non-negative, so is M(Ψ) (O). In particular, we have now have settled all statements in [41, 53]. Acknowledgment This research was supported by the START project FLAME Y551-N13 of the Austrian Science Fund (FWF) and the DACH project BIOTOP I-1018- N25 of the Austrian Science Fund (FWF) and 200021E-142224 of the Swiss National Science Foundation (SNSF). The authors like to thank Stephan Dahlke, Wolfgang Kreuzer, and Diana Stoeva for fruitful discussions. References [1] Akram Aldroubi, Portraits of frames, Proc. Amer. Math. Soc. 123 (1995), no. 6, 1661 -- 1668. [2] Jean-Pierre Antoine, Quantum mechanics beyond Hilbert space, pp. 1 -- 33, Springer, Berlin -- Heidelberg, 1998. [3] Jean-Pierre Antoine and Peter Balazs, Frames and semi-frames, J. Phys. A: Math. Theor. 44 (2011), 205201. [4] , Frames, semi-frames, and Hilbert scales, Numer. Funct. Anal. Optim. 33 (2012), no. 7 -- 9, 736 -- 769. [5] Peter Balazs, Hilbert-Schmidt operators and frames -- classification, best approximation by multipliers and algorithms, Int. J. Wavelets Mul- tiresolution Inf. Process. 6 (2008), no. 2, 315 -- 330. [6] , Matrix-representation of operators using frames, Sampl. Theory Signal Image Process. 7 (2008), no. 1, 39 -- 54. 19 [7] Peter Balazs, Jean-Pierre Antoine, and Anna Grybos, Weighted and controlled frames: Mutual relationship and first numerical properties, Int. J. Wavelets Multiresolution Inf. Process. 8 (2010), no. 1, 109 -- 132. [8] Peter Balazs, Hans G. Feichtinger, Mario Hampejs, and Gunther Kracher, Double preconditioning for Gabor frames, IEEE Trans. Signal Process. 54 (2006), no. 12, 4597 -- 4610. [9] Peter Balazs and Karlheinz Grochenig, A guide to localized frames and applications to Galerkin-like representations of operators, Frames and Other Bases in Abstract and Function Spaces, vol. 1 (Isaac Pesenson, Hrushikesh Mhaskar, Azita Mayeli, Quoc T. Le Gia, and Ding-Xuan Zhou, eds.), Applied and Numerical Harmonic Analysis, Birkhauser, Basel, 2017, pp. 47 -- 79. [10] Peter Balazs, Nicki Holighaus, Thibaud Necciari, and Diana T. Sto- eva, Frame theory for signal processing in psychoacoustics, Excursions in Harmonic Analysis, vol. 5, (Radu Balan, John J. Benedetto, Wojciech Czaja, and Kasso Okoudjou, eds.), Applied and Numerical Harmonic Analysis, Birkhauser, Basel, pp. 225 -- 268. [11] Peter Balazs and Georg Rieckh, Redundant representation of operators, arXiv:1612.06130, 2016. [12] , Oversampling operators: Frame representation of operators, Analele Universitatii "Eftimie Murgu" 18 (2011), no. 2, 107 -- 114. [13] Peter Balazs, Diana T. Stoeva, and Jean-Pierre Antoine, Classification of general sequences by frame-related operators, Sampl. Theory Signal Image Process. 10 (2011), no. 2, 151 -- 170. [14] John J. Benedetto and William. Heller, Irregular sampling and the theory of frames, I, Note di Matematica 10 (1990), no. 1, 103 -- 125. [15] Iva Bogdanova, Pierre Vandergheynst, Jean-Pierre Antoine, Laurent Jacques, and Marcela Morvidone, Stereographic wavelet frames on the sphere, Appl. Comput. Harmon. Anal. 19 (2005), 223 -- 252. [16] Helmut Bolcskei, Franz Hlawatsch, and Hans Georg Feichtinger, Frame- theoretic analysis of oversampled filter banks, IEEE Trans. Signal Pro- cessing 46 (1998), no. 12, 3256 -- 3268. [17] Abdelkrim Bourouihiya, The tensor product of frames, Sampl. Theory Signal Image Process. 7 (2008), no. 1, 427 -- 438. 20 [18] Dietrich Braess, Finite Elements. Theory, Fast Solvers, and Applications in Solid Mechanics, second ed., Cambridge University Press, Cambridge, 2001. [19] James Bramble, Joseph Pasciak, and Jingjia Xu, Parallel multilevel pre- conditioners, Math. Comput. 55 (1990), 1 -- 22. [20] Susann C. Brenner and L. Ridgway Scott, The Mathematical Theory of Finite Element Methods, Texts in Applied Mathematics, vol. 15, Springer, New York, 1994. [21] Peter Casazza, The art of frame theory, Taiwanese J. Math. 4 (2000), no. 2, 129 -- 202. [22] Peter Casazza, Ole Christensen, and Diana T. Stoeva, Frame expansions in separable Banach spaces, J. Math. Anal. Appl. 307 (2005), no. 2, 710 -- 723. [23] Peter G. Casazza and Xuemei Chen, Frame scalings: A condition num- ber approach, Linear Algebra Appl. 523 (2017), 152 -- 168. [24] Ole Christensen, An Introduction to Frames and Riesz Bases, second ed., Birkhauser/Springer, Cham, 2016. [25] Ole Christensen and Diana T. Stoeva, p-frames in separable Banach spaces., Adv. Comput. Math. 18 (2003), no. 2 -- 4, 117 -- 126. [26] Elena Cordero, Hans G. Feichtinger, and Franz Luef, Banach Gelfand triples for Gabor analysis, Pseudo-differential Operators, Lecture Notes in Mathematics, vol. 1949, Springer, Berlin, 2008, pp. 1 -- 33. [27] Lawrence Crone, A characterization of matrix operator on ℓ2, Math. Z. 123 (1971), 315 -- 317. [28] Stephan Dahlke, Massimo Fornasier, and T. Raasch, Adaptive frame methods for elliptic operator equations, Adv. Comput. Math. 27 (2007), no. 1, 27 -- 63. [29] Wolfgang Dahmen, Wavelet and multiscale methods for operator equa- tions, Acta Numerica 6 (1997), 55 -- 228. [30] Wolfgang Dahmen and Reinhold Schneider, Composite wavelet basis for operator equations, Math. Comp. 68 (1999), 1533 -- 1567. 21 [31] Ingrid Daubechies, Ten Lectures on Wavelets, CBMS-NSF Regional Conference Series in Applied Mathematics, SIAM, Philadelphia, 1992. [32] Richard J. Duffin and Albert C. Schaeffer, A class of nonharmonic Fourier series, Trans. Amer. Math. Soc. 72 (1952), 341 -- 366. [33] Lothar Gaul, Martin Kogler, and Marcus Wagner, Boundary Element Methods for Engineers and Scientists, Springer, Berlin, 2003. [34] Jean-Pierre Gazeau, Coherent States in Quantum Physics, Wiley, Wein- heim, 2009. [35] Israel Gohberg, Seymour Goldberg, and Marinus A. Kaashoek, Classes of Linear Operators, vol. I, Operator Theory: Advances and Applica- tions, Birkhauser, Basel, 1990. [36] Israel Gohberg, Seymour Goldberg, and Marinus A. Kaashoek, Basic Classes of Linear Operators, Birkhauser, Basel, 2003. [37] Michael Griebel, Multilevelmethoden als Iterationsverfahren uber Erzeu- gendensystemen, B.G. Teubner, Stuttgart, 1994. [38] Karlheinz Grochenig, Describing functions: Atomic decompositions ver- sus frames, Monatsh. Math. 112 (1991), no. 3, 1 -- 41. [39] Karlheinz Grochenig, Foundations of Time-Frequency Analysis, Birkhauser, Boston, 2001. [40] Phillipp Grohs and Axel Obermeier, Optimal adaptive ridgelet schemes for linear transport equations, Appl. Comput. Harmon. Anal. 41 (2015), no. 3, 768 -- 814. [41] Helmut Harbrecht, Reinhold Schneider, and Christoph Schwab, Multi- level frames for sparse tensor product spaces, Numer. Math. 110 (2008), no. 2, 199 -- 220. [42] Harro Heuser, Funktionalanalysis -- Theorie und Anwendungen, fourth ed., B.G. Teubner, Stuttgart, 2006. [43] Laurent Jacques, Ondelettes, Rep`eres et Couronne Solaire, Ph.D. thesis, Univ. Cath. Louvain, Louvain-la-Neuve, 2004. [44] Augustus J.E.M. Janssen and Peter Søndergaard, Iterative algorithms to approximate canonical Gabor windows: Computational aspects, J. Fourier Anal. Appl. 13 (2007), no. 2, 211 -- 241. 22 [45] Jens Kappei, Adaptive Wavelet Frame Methods for Nonlinear Elliptic Problems, Logos-Verlag, Berlin, 2012. [46] Gitta Kutyniok, Kasso A. Okoudjou, Friedrich Philipp, and Elizabeth K. Tuley, Scalable frames, Linear Algebra Appl. 438 (2013), 2225 -- 2238. [47] Peter Osswald, On function spaces related to finite element approxima- tion theory, Z. Anal. Anwend. 9 (1990), 43 -- 66. [48] Peter Oswald, Multilevel Finite Element Approximation, B.G. Teubner, Stuttgart, 1994. [49] Nathanael Perraudin, Nicki Holighaus, Peter Søndergaard, and Pe- ter Balazs, Designing Gabor windows using convex optimization, Appl. Math. Comput., (2018), to appear. [50] Isaac Pesenson, Hrushikesh Mhaskar, Azita Mayeli, Quoc T. Le Gia, and Ding-Xuan Zhou, Frames and Other Bases in Abstract and Function Spaces, Applied and Numerical Harmonic Analysis, Birkhauser, Basel, 2017. [51] Stevan Pilipovi´c and Diana T. Stoeva, Fr´echet frames, general definition and expansion, Anal. Appl. 12 (2014), no. 2, 195 -- 208. [52] Stefan Sauter and Christoph Schwab, Boundary Element Methods, Springer Series in Computational Mathematics, Springer, Berlin -- Heidelberg, 2010. [53] Rob Stevenson, Adaptive solution of operator equations using wavelet frames, SIAM J. Numer. Anal. 41 (2003), no. 3, 1074 -- 1100. [54] Diana T. Stoeva and Peter Balazs, Weighted frames and frame multi- pliers, Annual of the University of Architecture, Civil Engineering and Geodesy XLIIILIV 2004 -- 2009 (Fasc. II Mathematics Mechanics), 2012, pp. 33 -- 42. [55] Diana T. Stoeva, Xd-frames in Banach spaces and their duals, Int. J. Pure Appl. Math. 52 (2009), no. 1, 1 -- 14. [56] Dirk Werner, Funktionalanalysis, Springer, Berlin, 1995. [57] Manuel Werner, Adaptive Wavelet Frame Domain Decomposition Meth- ods for Elliptic Operator Equations, Ph.D. thesis, Philipps-Universitat Marburg, 2009. 23
1510.05441
1
1510
2015-10-19T12:08:45
Unbounded composition operators via inductive limits: cosubnormal operators with matrix symbols. II
[ "math.FA" ]
The paper deals with unbounded composition operators with infinite matrix symbols acting in $L^2$-spaces with respect to the gaussian measure on $\mathbb{R}^\infty$. We introduce weak cohyponormality classes $\EuScript{S}_{n,r}^*$ of unbounded operators and provide criteria for the aforementioned composition operators to belong to $\EuScript{S}_{n,r}^*$. Our approach is based on inductive limits of operators.
math.FA
math
UNBOUNDED COMPOSITION OPERATORS VIA INDUCTIVE LIMITS: COSUBNORMAL OPERATORS WITH MATRIX SYMBOLS. II PIOTR BUDZY ´NSKI, PIOTR DYMEK, AND ARTUR P LANETA Abstract. The paper deals with unbounded composition operators with infinite matrix sym- bols acting in L2-spaces with respect to the gaussian measure on R∞. We introduce weak cohyponormality classes S∗ n,r of unbounded operators and provide criteria for the aforemen- n,r. Our approach is based on inductive limits of tioned composition operators to belong to S∗ operators. 1. Introduction Bounded composition operators in L2-spaces are classical object of investigation in operator theory (see the monograph [24]). Their unbounded counterparts have attracted attention quite recently but already proved to be source of interesting problems and results (see [4, 7, 8, 9, 11, 17]. Many of them are related to the subnormality, a subject widely recognized as difficult and important in operator theory (see the monograph [14] concerning bounded subnormal operators, and trilogy [26, 27, 28] concerning unbounded ones). There is no effective general criterion for subnormality of unbounded operators. As a conse- quence, the methods of verifying the subnormality of an operator depends on its properties. In general, the moment problem approach has been very successful (cf. [29, 13]), especially for oper- ators with a dense set of C∞-vectors. On the other hand, for unbounded composition operators the consistency condition approach (related to a problem of selecting appropriate probability mea- sures) is much better (cf. [9]). This calls for testing various methods when studying the subject. As shown recently, inductive limit techniques might also be helpful in this matter, e.g., in a case of weighted shifts on directed trees or composition operators (cf. [5, 6, 4]). In a recent paper [4] we have provided a criterion for cosubnormality of unbounded composition operators induced by finite matrix symbols and also a new proof of the criterion for subnormality of these operators given in [9]. Inductive limit method played a pivotal role in the proofs. A natural setting for generalization and new area of testing our methods is where finite matrix symbols are exchanged by infinite ones. Unbounded composition operators with such symbols have already been investigated in [11], where we have dealt with questions of their boundedness and dense definiteness. Motivated by these previous results and the criterion for subnormality of general unbounded operators due to Stochel and Szafraniec (see [27, Theorem 3]), we introduce in this paper classes S∗n,r of unbounded operators closely related to cosubnormal operators (they resemble, in a sense, weak hyponormality classes studied in the case of bounded operators, cf. [22, 20, 15, 19]) and investigate under what conditions composition operators with infinite matrix symbols belong to the classes. We use inductive limits to achieve our goal. This results in three criteria (see Theorem 5.1 and Propositions 5.2 and 5.9). The symbol of a composition operator in the first of the criteria has unspecified form whence in the second and the third one the symbol is induced by an infinite matrix. Using this type of matrices allowed us to formulate the criterion in a tractable form, which consequently enabled us to construct explicit examples. To the best of our knowledge none of the examples cannot be studied by other means than our criteria. The paper is organized as follows. We begin by introducing basic notation and defining the classes Sn,r and S∗n,r in Section 2. In Section 3 we provide necessary information about com- position operators in L2-spaces and their relatives -- weighted composition operators and partial 2010 Mathematics Subject Classification. Primary 47B33, 47B37; secondary 47A05, 28C20. Key words and phrases. Composition operator in L2-space, inductive limits of operators, subnormal operators. 1 2 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA composition operators. Then, in Section 4, we give a brief description of composition operators with finite and infinite matrix symbols in L2-spaces with respect to the gaussian measure. The last part of the paper, Section 5, is devoted to the criteria and examples. 2. Preliminaries Throughout the paper Z+, N, R and C stand for the set of nonnegative integers, positive integers, real numbers and complex numbers, respectively. For κ ∈ N, Iκ stands for the set {1, 2, . . . , κ}, the Cartesian product of κ-copies of R is denoted by Rκ, and R∞ denotes the Cartesian product of ℵ0-copies of R. If t, s ∈ N satisfy s > t, then by πs t and πt we denote the mappings πs t : Rs ∋ (x1, . . . , xs) 7→ (x1, . . . , xt) ∈ Rt and πt : R∞ ∋ (x1, x2, . . .) 7→ (x1, x2, . . . , xt) ∈ Rt. If {Xn}∞n=1 is a sequence of subsets of a set X such that Xk ⊆ Xk+1 for every k ∈ N and X = S∞n=1 Xn, then we write Xn ր X as n → ∞. The symmetric difference of sets A and B is denoted by A△B. For a topological space X, B(X) stands for the σ-algebra of all Borel subsets of X. If κ ∈ N and p = {pn}∞n=1 ⊆ (0, +∞), then ℓκ(p), or ℓκ(cid:0){pn}∞n=1(cid:1), stands for the weighted ℓκ-space (cid:8){xn}∞n=1 ∈ R∞ : P∞n=1 xnκpn < ∞(cid:9); ℓκ(N) denotes the space ℓκ(cid:0){1}∞n=1(cid:1). Let H be a (complex) Hilbert space and T be an operator in H (all operators are linear in this paper). By D(T ), T , and T ∗ we denote the domain, the closure, and the adjoint of T , respectively (if they exists). If T is closable and F is a subspace of D(T ) such that TF = T , then F is said to be a core of T . If T is densely defined, D(T ) ⊆ D(T ∗) and kT ∗fk 6 kT fk for all f ∈ D(T ), then T is called hyponormal. T is said to be subnormal if D(T ) is dense in H and there exist a complex Hilbert space K and a normal operator N in K (i.e., N is closed, densely defined and satisfies N∗N = N N∗) such that H is isometrically embedded in K and Sh = N h for all h ∈ D(S). p,q=0,...,n ⊂ C. Then na denotes the greatest na ∈ In ∪ {0} satisfying the following condition Let n ∈ Z+, m ∈ N and a = {ai,j p,q}i,j=1,...,m there exist i, j ∈ Im such that ai,j na,q + ai,j p,na > 0 for some p, q ∈ In. Clearly, ai,j p,q = 0 for all i, j ∈ Im and p, q ∈ In such that p > na and q > na. Definition 2.1. Let n, r ∈ Z+. We say that a densely defined operator T in a Hilbert space H belongs to the class Sn,r if and only if for every m ∈ N and every a = {ai,j (2.1) p,q}i,j=1,...,m p,q=0,...,n ⊂ C, p,qλp ¯λqzi ¯zj > 0, ai,j λ, z1, . . . , zm ∈ C, mXi,j=1 naXp,q=0 implies (2.2) rXk,l=0 naXp,q=0 mXi,j=1 i : i = 1, . . . , m, k = 0, . . . , r} ⊆ D(T na+r). In turn, we say that T p,qhT p+kf l ai,j i , T q+lf k j i > 0, for every finite sequence {f k belongs to S∗n,r if and only if T ∗ belongs to Sn,r. Remark 2.2. The case of n = r = 0 is of little interest to us, since every densely defined linear operator in H belongs to S0,0 (use the classical Schur lemma). Therefore, in the rest of the paper we tacitly assume that n + r > 1. The following is essentially contained in [27, Theorem 3] and [13, Theorem 29] (for the reader convenience we give a sketch of the proof). Proposition 2.3. Let S be an operator in a complex Hilbert space H. Then the following condi- tions are satisfied: (i) if S is subnormal, then S ∈ Sn,r for all n, r ∈ Z+, (ii) if S ∈ Sn,0 for every n ∈ Z+, then SD∞(S) is subnormal. Sketch of the proof. (i) Assume that S is a subnormal operator. Fix n, r ∈ Z+. For m ∈ N and a = {ai,j p,q=0,...,n ⊂ C such that (2.1) is satisfied, we define the polynomials pi,j of two complex p,q}i,j=1,...,m COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 3 variables λ and ¯λ by pi,j(λ, ¯λ) = naXp,q=0 p,qλp¯λq. ai,j i=1,...,m ⊆ D(Sna+r). Then i=1,...,m ⊆ D(N na+r). Since D(N∗k) = D(N k∗) for every k ∈ Z+, we have D(N na+r) ⊆ Let N be a normal extension of S in a Hilbert space K. Consider {f k {f k D(N∗r) and N∗r D(N na+r) ⊆ D(N na). This implies that i }k=0,...,r i }k=0,...,r gi = rXk=0 N∗kf k i ∈ D(N na) for i ∈ Im. Moreover, since hN∗kf, N∗lf′i = hN lf, N kf′i for all f, f′ ∈ D(cid:0)N max{k,l}(cid:1) and k, l ∈ Z+, we get naXp,q=0 = rXk,l=0 naXp,q=0 ai,j p,qhSp+kf l rXk,l=0 i , Sq+lf k j i = i , N q+lf k j i naXp,q=0 rXk,l=0 j i = ai,j p,qhN p+kf l naXp,q=0 ai,j p,qhN pN∗lf l i , N qN∗kf k ai,j p,qhN pgi, N qgji. Hence, proving (i) amounts to showing that mXi,j=1 naXp,q=0 ai,j p,qhN pgi, N qgji > 0. Let E be the spectral measure of N . For i, j ∈ {1, . . . m}, let hi,j be the Radon-Nikodym derivative mPi=1hE(·)gi, gii. of the complex measure hE(·)gi, gji with respect to the non-negative measure µ = Arguing as in the proof of [27, Theorem 3] we deduce that mXi,j=1 naXp,q=0 p,qhN pgi, N qgji = ai,j ZC mXi,j=1 pi,j(λ, ¯λ)hi,j (λ) dµ(λ) > 0. This completes the proof of (i). (ii) This follows directly from [13, Theorem 29]. (cid:3) It is worth noticing that every operator in Sn,1, n ∈ Z+, belongs to the class of hyponormal operators, provided its domain is invariant for the adjoint, in a sense. This follows from the fact that Sn+s,1 ⊆ Sn,1 for any s ∈ Z+ and the following. Proposition 2.4. Let T be densely defined operator such that T ∈ S0,1. If D(T ) ⊆ D(T ∗) and T ∗ D(T ) ⊆ D(T ), then T is hyponormal. Proof. Note that (2.1) is satisfied with m = 1, n = 0 and a1,1 0,0 = 1. Then, by (2.2) with r = 1, we have hf, fi + hg, T fi + hT f, gi +hT g, T gi > 0 for all f, g ∈ D(T ). Substituting f = −T ∗g into this inequality we get kT ∗gk 6 kT gk for every g ∈ D(T ), which completes the proof. (cid:3) 3. Composition operators Let (X,A, µ) be a σ-finite measure space. Let A be an A-measurable transformation of X. Define the measure µ◦ A−1 on A by setting µ◦ A−1(σ) = µ(A−1(σ)), σ ∈ A. If A is nonsingular, i.e., µ ◦ A−1 is absolutely continuous with respect to µ, then the operator CA : L2(µ) ⊇ D(CA) → L2(µ) given by D(CA) = {f ∈ L2(µ) : f ◦ A ∈ L2(µ)} and CAf = f ◦ A for f ∈ D(CA), is well defined and closed in L2(µ) (cf. [7, Proposition 1.5]), where L2(µ) = L2(X,A, µ) is the space of all A-measurable C-valued functions with RX f2 dµ < ∞. Call it a composition operator 4 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA induced by A; we say that A is the symbol of CA then. If the Radon-Nikodym derivative hA = dµ ◦ A−1 dµ belongs to L∞(µ), the space of all essentially bounded A-measurable C-valued functions on X, then CA is bounded on L2(µ) and kCAk = khAk1/2 L∞(µ). The reverse is also true. By the measure transport theorem we get It follows from [7, Proposition 3.2] that D(CA) = L2((1 + hA) dµ). (3.1) D(CA) = L2(µ) if and only if hA < ∞ a.e. [µ]. As shown below, in case of finite measure spaces dense definiteness of CA is automatic. Lemma 3.1. If (X,A, µ) is a finite measure space and A is a nonsingular transformation of X, then χσ ∈ D(Cn Proof. Since µ(X) < +∞ and χσ ◦ A = χA−1(σ), we deduce that χσ ∈ D(Cn n ∈ N. Therefore, D(Cn "moreover" part. A) for all σ ∈ A and A) is dense in L2(µ) for every n ∈ Z+. This and [7, Theorem 4.7] yield the A) for every σ ∈ A and n ∈ N. Moreover, D∞(CA) is dense in L2(µ). (cid:3) Now we recall some information concerning weighted composition operators. Let (X,A, ν) be a σ-finite measure space, A be a nonsingular A-measurable transformation of X and w be a A- measurable C-valued function on X such that the measure (w2 dν)◦ A−1 is absolutely continuous with respect to ν. A weighted composition operator WA,w : L2(ν) ⊇ D(WA,w) → L2(ν) is defined by D(WA,w) = {f ∈ L2(ν) : w · (f ◦ A) ∈ L2(ν)}, WA,wf = w · (f ◦ A), f ∈ D(WA,w). Any such operator WA,w is closed. Moreover, WA,w is densely defined if and only if hA·(EA(w2)◦ A−1) < ∞ a.e. [ν], where EA(·) denotes the conditional expectation operator with respect to σ- algebra A−1(A) (cf. [12, Lemma 6.1]). We refer the reader to [10] for more information on unbounded weighted composition operators and references. The adjoint of a composition operator induced by A-bimeasurable transformation turns out to be the weighted composition operator induced by the inverse of the symbol (cf. [7, Corollary 7.3] and [7, Remark 7.4]): (3.2) If A is an invertible transformation of X such that both A and A−1 are A-measurable and nonsingular, then C∗A = WA−1,hA . If µ is a finite measure, then the above implies immediately the following. Lemma 3.2. Let n ∈ N. Suppose (X,A, µ) is a finite measure space and A is an invertible transformation of X such that both A and A−1 are A-measurable and nonsingular. Then the following conditions are satisfied: n\k=1 D(cid:0)C∗Ak(cid:1), (i) D(cid:0)(C∗A)n(cid:1) = A(cid:1)∗ = C∗An , ⊆(cid:0)Cn (ii) (cid:0)C∗A(cid:1)n (iii) (cid:0)C∗A(cid:1)n =(cid:0)Cn A(cid:1)∗, whenever D(cid:0)C∗An(cid:1) = A ⊆ CAn (cf. [7, Inclusion (3.3)]), Lemma 3.1 implies that Cn Proof. Since Cn defined. This together with [7, Corollary 4.2.] yields D(cid:0)C∗Ak(cid:1). n\k=1 By [9, Lemma 15], we have A(cid:1)∗ = C∗An . hAn = hA · hA ◦ A−1 · hA ◦ A−2 ··· hA ◦ A−(n−1) A(cid:1)∗ =(cid:0)Cn (cid:0)Cn (3.3) (3.4) A and CAn are densely a.e. [µ], n ∈ N. COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 5 In view of (3.2) and (3.4), we obtain the equality (3.5) D(cid:0)(C∗A)n(cid:1) = n\k=1(cid:8)f ∈ L2(µ) : h Ak f ◦ A−k ∈ L2(µ)(cid:9) = n\k=1 D(cid:0)C∗Ak(cid:1). Thus (i) is satisfied. The fact that (B∗)n ⊆ (Bn)∗ for any operator B in a Hilbert space such that the adjoints exist imply that (3.6) (cid:0)C∗A(cid:1)n ⊆(cid:0)Cn A(cid:1)∗, which combined with (3.3) proves (ii). The equality in (iii) follows from (3.3), (3.5), (3.6) and the assumption on D(cid:0)C∗An(cid:1). That the inclusion (C∗A)n ⊆ C∗An in Lemma 3.2 can be proper is shown below. Example 3.3. Let X = [0, 1], A = B([0, 1]) and µ(σ) = Rσ 1√x dm1, where m1 is the Lebesgue measure on R. Let A(x) = 1 − x. Clearly, A and A−1 are A-measurable and nonsingular. Since A2(x) = x for every x ∈ X, we get C∗A2 = I, where I denotes the identity operator. On the other hand, by (3.5) we have (cid:3) D(cid:0)(C∗A)2(cid:1) =nf ∈ L2(µ) : ZX f (x)2q x 1−x dµ < ∞o 6= L2(µ). Now we show that certain families generated by characteristic functions form cores for n- tuples of weighted composition operators (this generalizes [4, Proposition 3.3]). Given operators i=1 D(Ti), we say that F is a (T1,...,Tn) := This proves that D(cid:0)(C∗A)2(cid:1) 6= D(C∗A2 ). T1, . . . , Tn, n ∈ N, in a Hilbert space H and a subspace F ⊆ Tn core for (T1, . . . , Tn) if F is dense in Tn kfk2 +Pn Proposition 3.4. Let (X,A, µ) be a σ-finite measure space and let B ⊆ A be a family of sets satisfying the following conditions i=1 D(Ti) with respect to the graph norm kfk2 i=1 kTifk2. (i) for all A, B ∈ B, A ∩ B ∈ B, (ii) A = σ(B), (iii) there exists {Xn}∞n=1 ⊆ B such that Xk ⊆ Xk+1 for every k ∈ N and µ(cid:0)X \S∞n=1 Xn(cid:1) = 0. Let n ∈ N. Suppose Ai : X → X, i ∈ In, is invertible, Ai and A−1 are A-measurable and non- singular. Let wi : X → C, i ∈ In, be A-measurable. If F := lin(cid:8)χσ : σ ∈ B(cid:9) ⊆Tn i=1 D(WAi,wi), then F is a core of (WA1,w1 , . . . , WAn,wn). Proof. Let hAi,wi, i ∈ In, denote the Radon-Nikodym derivative of the measure (cid:0)wi2 dµ(cid:1) ◦ A−1 with respect to the measure µ. By [4, Lemma 3.2] and [10, Proposition 10], hAi,wi < ∞ a.e. [µ] for every i ∈ In. Therefore, the measure (cid:0)1 + hA1,w1 + . . . + hAn,wn(cid:1) dµ is σ-finite. Combining [4, Lemma 3.2] and [10, Proposition 9], we see that F is a core of (WA1,w1, . . . , WAn,wn ) (see also the proof of [4, Proposition 3.3]). (cid:3) i i It is sometimes convenient to consider composition operators, or even weighted composition operators, induced by partial transformations of X, i.e., mappings defined not on the whole of X, but on a subset of X; such composition operators (resp. weighted composition operators) are occasionally called partial composition operators (resp. partial weighted composition operators). Suppose Y ∈ A. Let B : Y → X and w : Y → C be A-measurable having A-measurable extensions B : X → X and w : X → C. If the weighted composition operator W B, w is well-defined, then we define the operator WB,w : D(WB,w) → L2(µ) by WB,w = W B,χY w. Clearly, if Y = X, then the above definition agrees with the previous one given for "everywhere defined" transformations, which justifies the notation. The partial composition operator comes 6 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA out of it, when we consider w = χY . Let us note that the definition is independent of the choice of extensions (see [10, Proposition 7]). In particular, we have (3.7) If A : X → X is A-measurable and nonsingular, u : X → C is A-measurable, and A and u are their restriction to a full measure µ subset Y of X, then W A,u = WA,u. In view of (3.7), we see that Lemma 3.1 is still valid if a composition operator CA is induced by a (partial) nonsingular measurable transformation A : Y → X defined on a full measure µ subset Y of X. Moreover, if additionally A is injective, A(Y ) is a set of full measure µ and A−1 : A(Y ) → X is nonsingular and measurable, then we get also the claim of Lemma 3.2. That "partial" counterpart of Proposition 3.4 is true can be easily proved as well. By (3.2) and (3.7), , where A−1 and A if µ is finite, then CA is densely defined operator in L2(µ) and C∗A = W A−1,h A are any A-measurable extensions of A−1 and A, respectively, onto X. Note that there are transformations, which are invertible and measurable but not nonsingular. Example 3.5. Consider X = R∞, A = B(R∞) and µ = µG, where µG denotes gaussian measure n xn(cid:9)∞ on R∞ (see the next Section for details). Let A : R∞ → R∞ be given by A{xn}∞n=1 =(cid:8) 1 n=1. It is clear that A is invertible and measurable. However, A cannot be nonsingular since A−1 transforms ℓ2(N) into ℓ2(cid:0){1/n2}∞n=1(cid:1) and, as we know by [1, Lemma 11 and Theorem 1.3, page 92], µG(cid:16)ℓ2(N)(cid:17) = 0 while µG(cid:16)ℓ2(cid:0){1/n2}∞n=1(cid:1)(cid:17) = 1. 4. Composition operators with matrix symbols Let κ ∈ N. The κ-dimensional gaussian measure is the measure µG,κ given by dµG,κ = 1 (√2π)κ exp(cid:18)− x2 1 + . . . + x2 κ 2 (cid:19) dmκ, where mκ denotes the κ-dimensional Lebesgue measure on Rκ. For any linear transformation A of Rκ, the composition operator CA in L2(µG,κ) is well-defined if and only if A is invertible. If this is the case, then (cf. [25, equation (2.1)]) (4.1) Here, and later on, · stands1 for the Euclidean norm on Rn, n ∈ N. Combining (4.1) with (3.1) and [7, Proposition 6.2], we get: hA(x) = det A−1 · exp x2 − A−1x2 x ∈ Rκ. 2 , (4.2) Every invertible linear transformation A of Rκ induces densely defined and injective composition operator CA in L2(µG,κ). Cosubnormality of CA can be written in terms of the symbol A (cf. Theorem 3.8]). Theorem 4.1. Let A be an invertible linear transformation of Rκ. If A is normal in (Rκ, · ), then CA is cosubnormal. The reverse implication holds whenever CA is bounded on L2(µG,κ). [25, Theorem 2.5] and [4, The gaussian measure µG on R∞ is the (infinite) tensor product measure µG = µG,1 ⊗ µG,1 ⊗ µG,1 ⊗ . . . defined on B(R∞). Recall that B(R∞) is generated by cylindrical sets, i.e., sets of the form σ × R∞, with σ ∈ B(Rn) and n ∈ N; the family of all cylindrical sets will be denoted by Bc(R∞). For every n ∈ N, the space L2(µG,n) can be naturally embedded into L2(µG) via the isometry ∆ : L2(µG,n) ∋ f 7→ f ◦ πn ∈ L2(µG). For simplicity we suppress the explicit dependence on n in the notation. We will denote by symbol k · k any of the L2-norm on L2(µG,n), n ∈ N, or L2(µG). A function f ∈ L2(µG) is called a cylindrical function if f ∈ ∆(cid:16)L2(µG,n)(cid:17) for some n ∈ N. If f ∈ L2(µG) is a cylindrical function and f = ∆ef , with ef ∈ L2(µG,k) and k ∈ N, then Σl(f ), l > k, denotes the set {x ∈ Rk : ef (x) 6= 0} × Rl−k. 1For simplicity we use the same dimension-independent symbol for any of the Euclidean norms on R n, n ∈ N . COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 7 We are interested in properties of composition operators with symbols induced by infinite matrices. Such operators were investigated in [11, Corollary 5.1], where tractable criteria for dense definiteness in case of row-finite matrices were given. In this paper we need to relax our approach slightly and consider also non row-finite case. This can be done as follows. We take a matrix a = (aij )i,j∈N with real entries and we set Using the Cauchy condition we see that aij xj ∈ R for every i ∈ No. ∞Xj=1 Da =n{xn}∞n=1 ∈ R∞ : Da = \i∈N\l∈N [N∈N \m>n>Nn{xk}∞k=1 : (cid:12)(cid:12) mXj=n a2j xj, . . .(cid:19), A(x1, x2, . . .) =(cid:18) ∞Xj=1 ∞Xj=1 a1j xj , lo, aijxj(cid:12)(cid:12) < 1 {xn}∞n=1 ∈ D, which yields Da ∈ B(R∞). Now, if A : D → R∞, with D ∈ B(R∞) such that D ⊆ Da, satisfies then we say that A is induced by a, or that a induces A. Such an A is B(R∞)-measurable. Indeed, this follows from the fact that sets of the form Rκ−1 × (α, β) × R∞, with κ ∈ N and α, β ∈ R, generate B(R∞) and A−1(cid:0)Rκ−1 × (α, β) × R∞(cid:1) = [N∈N \m>Nn{xn}∞n=1 ∈ D : α <(cid:12)(cid:12) mXj=1 aκjxj(cid:12)(cid:12) < βo. The operator CA is defined according to the scheme of defining composition operators with partial symbols (see the previous Section). It is obvious that Da is a set of full measure µG for every row-finite matrix a. As shown below, if a is not row-finite, but entries in each row have appropriate asymptotics, then one can deduce a similar result. Lemma 4.2. Let a = (aij )i,j∈N be a real matrix. If there exist λ ∈ (0, 1) and C > 0 such that ai,j 6 Cλi−j for all i, j ∈ N, then Da is a set of full measure µG. Proof. By the Cauchy-Schwartz inequality ℓ2(cid:0){λn}∞n=1(cid:1) ⊆ Da. Since, by [1, Lemma 11 and Theorem 1.3, page 92], µG(cid:16)ℓ2(cid:0){λn}∞n=1(cid:1)(cid:17) = 1 we get the claim. Our focus will be on composition operators with block 3-diagonal (or, using another language, block 1-banded) matrix symbols. For this we need some extra terminology. Let s = {s(n)}∞n=1 be an increasing sequence of natural numbers. Let a = (aij )i,j∈N be an infinite matrix such that (4.3) with s(0) := 0. Then, for p, q ∈ N, we define matrices ap = as aij = aji = 0, (i, j) ∈ {s(n − 1) + 1, . . . , s(n)} × {s(n + 1) + 1, . . .}, n ∈ N, pq by (cid:3) p and apq = as app = (aij)s(p) i,j=s(p−1)+1, ap,p+1 = (aij )s(p),s(p+1) i=s(p−1)+1,j=s(p)+1, i,j=1, ap = (aij )s(p) ap+1,p = (aij )s(p+1),s(p) i=s(p)+1,j=s(p−1)+1, and apq = 0, p − q > 1. With this notation a is indeed a block 3-diagonal matrix, i.e., ··· ··· ··· ··· rank ap,p+1 ∈(cid:8)0, s(p) − s(p − 1)(cid:9), If additionally to (4.3), the matrix a satisfies 0 a23 a33 ··· a11 a21 0 ··· a12 a22 a32 ···   a =   . p ∈ N, 8 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA then we say that a belongs to the class F(s). 5. Criteria and examples In this section we propose criteria answering the question of when a composition operator CA in L2(µG) induced by an infinite matrix is of class S∗n,r. The first (see Theorem 5.1 below) is rather general in nature, while the second and third (see Propositions 5.2 and 5.9) are more concrete, enabling us to construct explicit examples. Theorem 5.1. Let n, r ∈ Z+. Let s = {s(k)}∞k=1 ⊆ N be an increasing sequence. Suppose that (i) D ∈ B(R∞) satisfies µG(D) = 1, (ii) A : D → A(D) ⊆ R∞ is a B(R∞)-measurable nonsingular and invertible transformation (iii) there exist κ ∈ N and {Ωl : l ∈ N} ⊆ B(Rs(κ)) such that Ωl ր Rs(κ) as l → ∞ and (iv) for every k ∈ N there exists a B(Rs(k))-measurable nonsingular and invertible transfor- is B(Rs(k))-measurable nonsingular, CAk is densely such that µG(cid:0)A(D)(cid:1) = 1, A−1 is B(R∞)-measurable and nonsingular, χΩk×R∞ h mation Ak of Rs(k) such that A−1 k defined and cosubnormal in L2(µG,s(k)), Ai ∈ L2(µG) for all k ∈ N and i ∈ In+r. (v) for all i ∈ In+r and k ∈ N, l→∞ ZRs(l) (5.1) (5.2) lim sup χΩk×Rs(l)−s(κ) h2 (vi) for every m ∈ N there exists ep ∈ N such that for all i ∈ In+r and σ ∈ B(Rm) we have (vii) for every m ∈ N, σ ∈ B(Rm) and i ∈ In+r we have dµG,s(l) 6 kχΩk×R∞ h p (σ × Rs(p)−m) × R∞(cid:17)(cid:19) = 0, Aik2, Ai l µG(cid:18)(cid:16)A−i(σ × R∞)(cid:17)△(cid:16)A−i µG(cid:18)(cid:16)Ai(σ × R∞)(cid:17)△(cid:16) ∞[k=1 \p>k p > ep. p(σ × Rs(p)−m) × R∞(cid:17)(cid:19) = 0. Ai (5.3) Then CA ∈ S∗n,r. Proof. We divide the proof into few steps. , χAi (5.4) h2 Ai l l (eω×Rs(l)−m)∩Σs(l)(f ) dµG,s(l)(cid:19)1/2 Step 1. For every m ∈ N there exists ep ∈ N such that for all i ∈ In+r and ω = eω × R∞, with eω ∈ B(Rη) for some η ∈ N, and every f = ∆ef ∈ D(CAi ), with ef ∈ L2(µG,m), we have l > ep. (cid:12)(cid:12)hCAi f, χωi(cid:12)(cid:12) 6 kfk(cid:18)ZRs(l) Take σ =eσ × R∞ with eσ ∈ B(Rk) and k > η. Then, by condition (vi), there exists ep ∈ N such that for all p > ep and i ∈ In+r we have hCAi χσ, χωi =ZR∞ = µG,s(p)(cid:16)A−i =ZRs(p) =ZRs(p) 6 kχσk(cid:18)ZRs(p) p (eσ × Rs(p)−k) ∩eω × Rs(p)−η(cid:17) dµG,s(p)(cid:19)1/2 p · χeω×Rs(p)−η dµG,s(p) (χσ ◦ Ai) · χω dµG χeσ×Rs(p)−k ◦ Ai χeσ×Rs(p)−k · χAi p(eω×Rs(p)−η ) · h χeσ×Rs(p)−k∩Ai p(eω×Rs(p)−η ) dµG,s(p) Ai p h2 Ai p , COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 9 which means that (5.4) holds with f = χσ. If f = ∆ef ∈ D(CAi ), where ef is a nonnegative step function, then arguing as above we see that hCAi f, χωi =ZRs(p) ∆ef χAi 6 kfk(cid:18)ZRs(p) p(eω×Rs(p)−η ) h Ai p dµG,s(p) χΣs(p)(f )∩Ai p(eω×Rs(p)−η ) dµG,s(p)(cid:19)1/2 , h2 Ai p In turn, using approximation by step functions, we deduce that (5.4) holds for every f ∈ D(CAi ) that is nonnegative and f = ∆ef with ef ∈ L2(µG,m). For every cylindrical C-valued function f ∈ D(CAi ) its module is also a cylindrical function, f ∈ D(CAi ) and Σs(p)(f ) = Σs(p)(f). Hence, by the inequality p > ep, i ∈ In+r. (cid:12)(cid:12)hCAi f, χωi(cid:12)(cid:12) 6 hCAif, χωi, we get the claim. Step 2. For all i ∈ In+r, m ∈ N and ω ∈ B(Rm), if (5.5) l (ω×Rs(l)−m) lim sup l→∞ ZRs(l) χAi h2 Ai l dµG,s(l) < +∞, then χω×R∞ ∈ D(C∗Ai ). Moreover, if k ∈ N, i ∈ In+r and σ ∈ B(R∞) satisfy σ ⊆ A−i(Ωk × R∞), then χσ ∈ D(C∗Ai ). Fix i ∈ In+r. Let m ∈ N and ω ∈ B(Rm). Then, applying Step 1, inequality (5.5) and Proposition 3.4, we deduce that χω×R∞ ∈ D(C∗Ai ). Now fix k ∈ N. It follows from condition (vi) that µG(cid:18)(cid:16)A−i(Ωk × R∞)(cid:17)△(cid:16)A−i p (Ωk × Rs(p)−s(κ)) × R∞(cid:17)(cid:19) = 0, p > ep, and thus, by condition (v), we have lim sup l→∞ ZRs(l) χAi l (A−i l (Ωk×Rs(l)−s(κ))) h2 Ai l dµG,s(l) = lim sup l→∞ ZRs(l) χΩk×Rs(l)−s(κ) h2 Ai l dµG,s(l) < +∞. Hence, χA−i(Ωk×R∞) ∈ D(C∗Ai ) by the first part of the claim. Now, that χσ ∈ D(C∗Ai ) follows easily from (3.2) and the definition of the domain of a weighted composition operator. Step 3. For all i ∈ In+r and k ∈ N, and {lj}∞j=1 ⊆ N such that limj→∞ lj = ∞, we have (5.6) kχΩk×R∞ h Aik 6 lim sup j→∞ kχΩk×Rs(lj )−s(κ) h Ai lj k. Fix i ∈ In+r and k ∈ N. By Step 2, χA−i(Ωk×R∞) ∈ D(C∗Ai ), and thus using (3.2) we get f ∈ D(CAi ). Aii = hf, C∗Ai χA−i(Ωk×R∞)i = hCAi f, χA−i(Ωk×R∞)i, hf, χΩk×R∞ h (5.7) According to Lemma 3.1, equalities in (5.7) are satisfied for every cylindrical step function f . Moreover, for every cylindrical step function f , by condition (vi) and Step 1, we have (cid:12)(cid:12)hCAi f, χA−i(Ωk×R∞)i(cid:12)(cid:12) 6 kfk lim sup which together with (5.7) gives χAi lj j→∞ (cid:18)ZRs(lj ) Aii(cid:12)(cid:12) 6 kfk lim sup (cid:12)(cid:12)hf, χΩk×R∞ h j→∞ kχΩk×Rs(lj )−s(κ) h Ai lj k, (A−i lj (Ωk×Rs(lj )−s(κ))) dµG,s(lj )(cid:19)1/2 , h2 Ai lj for every cylindrical step function f . Since every function in L2(µG) may be approximated by cylindrical step functions, we get kχΩk×R∞ h Aik2 6 kχΩk×R∞ h Aik lim sup j→∞ kχΩk×Rs(lj )−s(κ) h Ai lj k, which proves (5.6). 10 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA Step 4. There exists an injective increasing sequence {lj}∞j=1 ⊆ N, such that for all i ∈ In+r and k ∈ N, (5.8) lim j→∞ kχΩk×R∞(cid:0)h Ai − ∆h Ai lj(cid:1)k = 0. First we prove (5.8) with i = k = 1. In view of (5.6) and (v) there exists an injective increasing sequence {l1,1 j }∞j=1 ⊆ N such that lim (5.9) Thus, by [31, Exercise 4.21(a)], it suffices to show that j k = kχΩ1×R∞ hAk. l 1,1 j→∞ kχΩ1×R∞ ∆hA j→∞ZΩ1×R∞ f hA dµG = lim ZΩ1×R∞ f ∆hA l 1,1 j dµG, f ∈ L2(µG). l 1,1 By condition (vi), equality in (5.9) holds when f is a cylindrical step function. Since cylindrical step functions are dense in L2(µG) and supj∈N kχΩ1×R∞ ∆hA j kL2(µG) < ∞, we get (5.9) and consequently (5.8) for i = k = 1. Repeating the same argument (n + r − 1) times (apply Step 3 to }∞j=1 ⊆ {l1,1 consecutive subsequences), we show that there exists a subsequence {ln+r,1 j }∞j=1 such that (5.8) is satisfied with i = 1, . . . , n + r and k = 1. In a similar manner we show that for any }∞j=1 ⊆ {ln+r,1 k0 ∈ N we can find subsequences {ln+r,k0 }∞j=1 such that (5.8) is satisfied with i = 1, . . . , n + r and k = 1, . . . , k0. Now, the sequence {lj}∞j=1 given by lj = ln+r,j does the job. Step 5. CA ∈ S∗n,r. turn, by (3.2), for every k ∈ N, C∗Ak In view of Lemma 3.1, (3.2) and (3.7), CA is densely defined in L2(µG) and C∗A = WA−1,hA. In }∞j=1 ⊆ . . . ⊆ {ln+r,2 = WA−1 k ,hA k . Set j j j j j S = WA−1,hA, and Sk = WA−1 lk ,hA lk , k ∈ N, where {lk}k∈N is as in Step 4 with additional requirement that l1 > κ. We denote by B the family of Borel subsets of R∞ defined as follows: σ ∈ B if and only if there exists eσ ∈ Bc(R∞) such that µG(σ△eσ) = 0 and either eσ = R∞ or there exists k1, . . . , kn+r ∈ N such that eσ ⊆ A−1(Ωk1 × R∞) ∩ . . . ∩ A−(n+r)(Ωkn+r × R∞). According to condition (vi) and the fact that Ωl ր Rκ as l → ∞ the family B satisfies conditions (i)-(iii) of Proposition 3.4. Indeed, conditions (i) and (iii) are clear. For the proof of (ii) we take m ∈ N and σ ∈ B(Rm), and for every k ∈ N we consider the sets ωk =(cid:0)σ × R∞(cid:1) ∩(cid:16)A−1(cid:0)Ωk × R∞(cid:1) ∩ . . . ∩ A−(n+r)(cid:0)Ωk × R∞(cid:1)(cid:17). p (cid:0)Ωk × Rs(p)−m(cid:1) ∩ . . . ∩ A−(n+r) eωk,p =(cid:16)(cid:0)σ × Rs(p)−m(cid:1) ∩ A−1 p ∈ N. Then, using (vi), we deduce that µG(ωk△eωk,p) = 0 for sufficiently large p ∈ N. Hence for every k, p ∈ N, ωk ∈ B and eωl,p ր σ × R∞ as l → ∞. This and the fact that Bc(R∞) generates B(R∞) Let X be a family composed of all characteristic functions χσ ∈ L2(µG) of sets σ ∈ B. Then by (cid:0)Ωk × Rs(p)−m(cid:1)(cid:17) × R∞, prove (ii). p and Step 2 we have (5.10) X ⊆ n+r\i=1 D(C∗Ai ). Thus, by Lemma 3.2 (i), X ⊆ D(Sn+r). For k ∈ N, let Bk denote the family of sets σ ∈ B(Rs(lk)) such that either σ = Rs(lk), or σ ⊆ A−1 (Ωmn+r × Rs(lk)−s(κ)) for some m1, . . . , mn+r ∈ N. Applying condition (v), (3.2) and Lemma 3.2 (i), we show that Xk ⊆ D(Sn+r ) for every sufficiently large k ∈ N, where Xk is composed of all characteristic functions χσ ∈ L2(µG,s(lk)) of sets σ ∈ Bk. (Ωm1 × Rs(lk)−s(κ)) ∩ . . . ∩ A−(n+r) lk lk k COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 11 By condition (vi), X =S∞i=1T∞k=i ∆Xk. In view of (5.10), (3.2) and Proposition 3.4, linX is a core for (C∗A, . . . , C∗An+r ). Thus, by Lemma 3.2, linX is a core for (S, . . . , Sn+r). Notice that, by condition (vii), for any x ∈ Xm with m ∈ N, and i ∈ In+r, we have (∆x)◦A−i = limp→∞(∆x)◦A−i p . Hence, by Lemma 3.2 and Step 4, we have Si∆x = h Ai ·(cid:16)∆x ◦ A−i(cid:17) = lim lk (cid:17) lk ·(cid:16)∆x ◦ A−i i ∈ In+r and x ∈ Xm, m ∈ N, k∆x, k→∞ ∆h Si Ai which implies that = lim k→∞ (5.11) hSi∆f, Sj∆gi = lim k→∞hSi k∆f, Sj k∆gi, for all ∆f, ∆g ∈ linX and i, j ∈ In+r. By subnormality of Sk and Proposition 2.3 we have Sk ∈ Sn,r for every k ∈ N. This and (5.11) imply that inequality in (2.2) (with S in place of T ) is satisfied for all {f k i : i = 1, . . . , m, k = 0, . . . , r} ⊆ X . This and the fact that linX is a core for (S, . . . , Sn+r) imply that CA ∈ S∗n,r. (cid:3) Proposition 5.2. Let n, r ∈ Z+. Let a = (aij )i,j∈N ⊆ R be a matrix such that µG(cid:0) Da(cid:1) = 1, and let s = {s(k)}∞k=1 ⊆ N be an increasing sequence. Assume that (a) A : Da → A(Da), the transformation induced by a, is nonsingular and invertible, and (b) A−1 is B(R∞)-measurable and nonsingular and induced by a matrix a−1 ∈ F(s), (c) there exist κ ∈ N and {Ωl : l ∈ N} ⊆ B(Rs(κ)) such that Ωl ր Rs(κ) as l → ∞ and (d) for every k ∈ N, the matrix (a−1)k is invertible and normal in (cid:0)Rs(k), · (cid:1), (e) for all i ∈ In+r and k ∈ N, inequality (5.1) is satisfied, where Al is the transformation of µG(cid:0)A(Da)(cid:1) = 1, χΩk×R∞ h Ai ∈ L2(µG) for all k ∈ N and i ∈ In+r, Rs(k) induced by the inverse of (a−1)l, l ∈ N. Then CA ∈ S∗n,r. Proof. We begin by showing that for every m ∈ N there exists ep ∈ N such that for all i ∈ In+r and σ ∈ B(Rm) condition (5.2) is satisfied. Indeed, fix i ∈ In+r, m ∈ N and σ ∈ B(Rm). Suppose that there exists p0 ∈ N such that Set ep = p0. Then normality of (a−1)ep+1 implies that rank (a−1)ep+1, ep = 0. This and surjectivity of A−1 yield (5.2). Now, suppose that for every p ∈ N such that s(p) > m we have (5.12) s(p0) > m and rank (a−1)p0,p0+1 = 0. rank (a−1)p,p+1 = s(p) − s(p − 1). It is easily seen that s(p) A−1 (5.13) p > ep, Using (5.13) repeatedly we obtain (5.2). p (cid:16)eσ × Rs(p)−m(cid:17) = πs(p+1) ◦ A−1 µG(cid:18)A−1(cid:16)σ × R∞(cid:17)△A−1 p+1(cid:16)eσ × Rs(p)−m × {0} × . . . × {0}(cid:17), where ep is the smallest integer such that s(ep − 1) > m. This and (5.12) imply that p (cid:16)eσ × Rs(p)−m(cid:17) × R∞(cid:19) = 0, p > ep. (cid:16)χσ×R∞ ◦ A−i(cid:17)(x) = 1 ⇐⇒(cid:16)(cid:0)A−i(x)(cid:1)1, . . . ,(cid:0)A−i(x)(cid:1)m(cid:17) ∈ σ p ◦ πp)(x)(cid:1)1, . . . ,(cid:0)(A−i p (cid:17)(x) = 1, This implies that χσ×R∞ ◦ A−i = limp→∞ ∆(cid:16)χσ×Rs(p)−m ◦ A−i ⇐⇒(cid:16)(cid:0)(A−i ⇐⇒ ∆(cid:16)χσ×Rs(p)−m ◦ A−i Now, if m ∈ N, σ ∈ Rm, i ∈ In+r, then there exists bp ∈ N such that for µG-a.e. x ∈ R∞ we have p ◦ πp)(x)(cid:1)m(cid:17) ∈ σ, p (cid:17), which all proves that (5.3) is p > bp p > bp satisfied. 12 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA Finally, since, by (4.2) and Theorem 4.1, CAk is densely defined and cosubnormal for every (cid:3) k ∈ N, we can apply Theorem 5.1. Remark 5.3. Concerning Theorem 5.1 it is worth noticing that, by Step 3, condition (iii) of Theorem 5.1 is automatically satisfied if lim supl→∞ kh lk < +∞, i ∈ In+r. Ai Below we provide examples of unbounded composition operators induced by infinite matrices that belong to S∗n,r. The first one, with diagonal matrix symbols, is motivated by [25, Theorem 4.1], where cosubnormality of bounded CA's of this kind was shown (by use of different methods). Example 5.4. Let n, r ∈ Z+. Let a = (aij )i,j∈N ⊆ R be a diagonal matrix, i.e. aii = αi with {αm}∞m=1 ⊆ (0,∞), and aij = 0 for all i, j ∈ N such that i 6= j. Assume that and (5.14) 0 < αk < 1 for all k ∈ N such that k > n + r, ∞Xk=1(cid:0)1 − αk(cid:1) is convergent. Clearly, a induces A-measurable invertible transformation A of R∞ such that Da is a set of full measure µG. Let s = {s(k)}∞k=1 ⊆ N be given by s(k) = k. Then A−1 is A-measurable transformation induced by a matrix a−1 ∈ F(s) such that Da−1 is a set of full measure µG. Let {Ωk : k ∈ N} be given by Ωk = [−k, k]n+r. By (4.1), for all i ∈ In+r, we have 1 x1, . . . , α−i h Ai l (x1, . . . , xl) = exp (x1, . . . , xl)2 −(cid:12)(cid:12)(cid:0)α−i exp(cid:16) − l xl(cid:1)(cid:12)(cid:12)2 (x1, . . . , xl) ∈ Rl, l ∈ N, 1 − α2i α2i j (cid:17), x2 j 2 2 j l 1 1 ··· αi αi lYj=1 1 αi j = where Al is the transformation of Rl induced by the matrix al, l ∈ N. This together with the [23, Theorem 8.26]) and (5.14) implies that there is C > 0 such change-of-variable theorem (cf. that kχΩk×Rl−n−r h Ai lk2 L2(µG,l) = dµG,1 α2i j −k j (1−α2i x2 j ) 1 α2i j Z k n+rYj=1 lYj=n+r+1 lYj=n+r+1(cid:16)αi e− √2π Z ∞ jq2 − α2i 1 α2i j 1 −∞ e− j (cid:17)−1 = C = C lYj=n+r+1 j Z ∞ 1 α2i −∞ j (1−α2i x2 j ) α2i j e− dµG,1 j (2−α2i x2 j ) 2α2i j dm1 , k, l ∈ N and l > n + r. jq2 − α2i Since αi j < 1 for every j > n + r and every i ∈ In+r, we infer from (5.14) and [21, lkL2(µG,l) < ∞. In view of [11, Corollary 5.1], the Chapter VII]) that 0 < liml→∞ kχΩk×Rl−n−r h above discussion shows that conditions (a), (b) and (d) of Proposition 5.2 are satisfied. Now, set Ai j x2 j 2 (cid:17), hi(cid:0){xj}j∈N(cid:1) = exp(cid:16) − j = 1 − α2i ∞Yj=1 1 αi j j }∞j=1 with p(i) 1 − α2i α2i j j . Since for every i ∈ In+r, the product Q∞j=1 α−i where pα,i = {p(i) convergent by (5.14) and [21, Theorem 3, p. 219]), and the series P∞j=1 is convergent for every {xj}∞j=1 ∈ ℓ2(pα,i), we see that hi is well-defined. By [1, Lemma 11 and Theorem 1.3, page 92], ℓ2(pα,i) is a set of full measure µG. Hence hi may be treated as a B(R∞)-measurable mapping defined on the whole of R∞. For all i ∈ In+r and k, m ∈ N and σ ∈ B(Rm), we have {xj}∞j=1 ∈ ℓ2(pα,i), i ∈ In+r, 1−α2i j α2i j x2 j 2 is j µG ◦ A−i(cid:0)(σ × R∞) ∩ (Ωk × R∞)(cid:1) = lim l→∞ µG,l(cid:0)A−i l (cid:0)(σ × Rl−m) ∩ (Ωk × Rl−n−r)(cid:1)(cid:1) COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 13 = lim l→∞Zσ×Rl−m χΩk×Rl−n−r h Ai l χΩk×R∞ hi dµG, dµG,l =Zσ×R∞ where the last equality follows from Lebesgue dominated convergence theorem and the fact that χΩk×R∞ hi is essentially bounded. This, together with [2, Theorem 10.3], implies that for every i ∈ In+r, Ai is nonsingular and hi = h Ai a.e. [µG]. Hence the conditions (a) to (e) of Proposition 5.2 are satisfied, and consequently CA ∈ S∗n,r. Remark 5.5. In view of [25, Theorem 4.1], the operator CA from Example 5.4 is bounded and cosubnormal, whenever2 αi ∈ (0, 1) for every i ∈ N. It is worth noticing that, if this is the case, cosubnormality of CA can also be deduced from Proposition 5.2. To this end, we first observe that CA ∈ S∗n,0 for every n ∈ N which follows from Theorem 5.1. On the other hand, by applying [11, Corollary 5.5], we see that CA is a bounded operator on L2(µG). Hence, using Proposition 2.3, we get cosubnormality of CA. Remark 5.6. It should be noted that any nontrivial scalar multiple of the identity mapping on R∞ cannot satisfy assumptions of Proposition 5.2. Indeed, suppose A : R∞ → R∞ is given by A(x) = αx, with α ∈ R \ {0, 1,−1}. Then either A or A−1 is singular, i.e., one of them is not nonsigular. To see this we first fix α ∈ (0, 1). Then we take any sequence {xn}∞n=1 ⊆ (0, 1) such that ∞Xn=1 xn < ∞ and ∞Xn=1 x√α n = ∞ µG,1 is a probability measure due to Poisson (cf. [30, p. 18-19]) we can show that n , n ∈ N. Using the well-known method of proving that the gaussian measure (5.15) Set an = p2 ln x−1 1 − exp(cid:16) − (5.16) a2 2 (cid:17) 6(cid:18) 1 √2π Z[−a,a] exp(cid:16) − x2 2 (cid:17) dm1(cid:19)2 a2 n 0 < ∞Yn=1 in view of (5.16) implies that (cid:18)1 − exp(cid:16) − By (5.15) and [21, Theorem 3, p. 219]) the product Q∞n=1(cid:16)1 − exp(cid:0) − a2 exp(cid:16) − √2πZ[−an,an] [−an, an](cid:1). On the other hand, the product Q∞n=1(cid:16)1 − exp(cid:0) − α2a2 2 (cid:17)(cid:19) 6(cid:18) ∞Yn=1 0 < µG(cid:0) ∞Yn=1 [21, Theorem 3, p. 219]). Hence, by (5.16), we deduce that This, in turn, yields 1 6 1 − exp(−a2), a ∈ (0,∞). n 2 (cid:1)(cid:17) is convergent, which 2 (cid:17) dm1(cid:19)2 6 1. x2 n(cid:1)(cid:17) is divergent to 0 (use again (5.15) and 0 = µG(cid:0) ∞Yn=1 which shows that A is singular. Similar reasoning proves the same for other α's belonging to R \ {0, 1 − 1} (if α > 1, then A−1 is singular). given by A(cid:0){xk}∞k=1(cid:1) = {αkxk}∞k=1, with {αk}∞k=1 ⊆ R such that either lim supk→∞ αk 6= 1 or lim inf k→∞ αk 6= 1, doesn't satisfy assumption of Proposition 5.2. In fact, modifying slightly the above argument, one can show that any transformation A of R∞ [−αan, αan](cid:1), Below, in Proposition 5.9, we provide another set of conditions implying that a satisfies the assumptions of Theorem 5.1. To this end we need an auxiliary lemma in which a class of infinite matrices (suitable for Proposition 5.9) is distinguished. Recall that a matrix (ai,j)i,j∈N is called η-banded, with η ∈ Z+, if ai,j = 0 for all i, j ∈ N such that i − j > η. Below, and later on, δij stands for the Kronecker's delta. 2If αm > 1 for some m ∈ N , then CA is unbounded in L2(µG) (cf. [25, Proposition 2.2]). 14 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA αj pj < ∞. Lemma 5.7. Let p = {pj}∞j=1 and {αj}∞j=1 be sequences of positive real numbers such that (a) {pj}∞j=1 and {αj}∞j=1 belong to ℓ1(N), (b) there are 0 < m < M < ∞ such that pj+1 (c) supj∈N pj ∈ (m, M ) for every j ∈ N, Let η ∈ Z+. Let bb = (bbi,j)i,j∈N be a η-banded matrix with real entries such that Let b = (δij +bbi,j)i,j∈N. Then the following conditions are satisfied (d) bbi,j 6 αi for all i, j ∈ N, (e) bb is symmetric, i.e. bbi,j =bbj,i for all i, j ∈ N. (1) bb induces a trace-class operator bB on ℓ2(N), (2) det b is well-defined; moreover, det b 6= 0 if and only if the operator I + bB is invertible, (3) there exists eC > 0 such that j −(cid:0)bx(cid:1)2 x = {xj}∞j=1 ∈ R∞. ∞Xj=1 ∞Xj=1 x2 j pj, x2 j(cid:12)(cid:12)(cid:12) 6 eC (4) if det b 6= 0, then b induces an invertible transformation of R∞. Proof. If {ei}∞i=1 is the standard orthonormal basis for ℓ2(N), then by (d) we have (cid:12)(cid:12)(cid:12) ∞Xi=1(cid:12)(cid:12)hbBei, eii(cid:12)(cid:12) 6 ∞Xi=1(cid:13)(cid:13)bBei(cid:13)(cid:13) = ∞Xi=1(cid:13)(cid:13)bBei(cid:13)(cid:13) 6 (2η + 1) ∞Xi=1(cid:13)(cid:13)UbBei(cid:13)(cid:13) ∞Xi=1 αi < ∞, = x2 αj pj we have j(cid:12)(cid:12)(cid:12) 6 j −(cid:0)bx(cid:1)2 This, according to [3, p. 46], imply (2). For the proof of (3) we first observe that if S is the shift operator S(x1, x2, . . .) = (0, x1, x2, . . .) acting on ℓ2(p), then, by (b), there exists C > 0 such that where bB = UbB is the polar decomposition of bB. This proves that bB is a trace-class operator. maxi∈Iη{kSik,kS∗ik} < C. Let c = max(cid:8) supj∈N , supj∈N αj(cid:9). Then for every {xj}∞j=1 ∈ ℓ2(p) j + 2xj(cid:12)(cid:12)(cid:0)bbx(cid:1)j(cid:12)(cid:12)(cid:19) 6 2 pj (cid:19)1/2 (cid:12)(cid:12)(cid:12) ηXk=−η (bbx)2 ∞Xj=1 ∞Xj=1 j+k + 2kxkℓ2(p)(cid:18)2 ηXk=−η ∞Xj=1 j+kpj + 2kxkℓ2(p)(cid:18)2c2 ηXk=−η j+k + 2kxkℓ2(p)(cid:18) ∞Xj=1 pj (cid:19)1/2 j+kpj(cid:19)1/2 ∞Xj=1 ℓ2(p) + 2kxkℓ2(p)(cid:18)2c2(2η + 1)C2kxk2 (cid:18)(cid:0)bbx(cid:1)2 ∞Xj=1 ηXk=−η ∞Xj=1 ηXk=−η ∞Xj=1 6 2c2(2η + 1)C2kxk2 ℓ2(p)(cid:19)1/2 6 2c2 α2 j x2 α2 j x2 α2 j x2 6 2 x2 x2 j+k . j This proves (3). If det b 6= 0, then, by (2), the matrix b induces an invertible operator B in ℓ2(N). Let a denote the matrix of the operator B−1 with respect to the standard orthonormal basis of ℓ2(N). In view of [16, Proposition 2.3] and Lemma 4.2, Da is a set of full measure µG. Let A be the transformation of R∞ induced by a, B be the transformation induced by b. It is a matter of simple calculations (based on the fact that BB−1 = B−1B = I) to show that AB and BA coincide with the identity mapping on R∞ up to sets of full measure µG. (cid:3) Regarding Lemma 5.7, we note that if bb is a matrix satisfying assumptions (a) and (d) of Lemma 5.7, then for every k ∈ N we have (cid:12)(cid:12)(cid:12)(cid:16)bbk(cid:17)i,j(cid:12)(cid:12)(cid:12) 6(cid:16) ∞Xl=1 αl(cid:17)k−1 αi, i, j ∈ N. COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 15 As a consequence, we get the following. Corollary 5.8. Under the assumptions of Lemma 5.7 for every k ∈ N we have (bb)k is a trace-class operator, det bk is well-defined and there exists eC > 0 such that ℓ2(p), x = {xj}∞j=1 ∈ ℓ2(p). (cid:12)(cid:12)(cid:12) ∞Xj=1 x2 j −(cid:0)bkx(cid:1)2 j(cid:12)(cid:12)(cid:12) 6 eCkxk2 Proposition 5.9. Let n, r ∈ Z+ and η ∈ N. Let bb = (bbij )i,j∈N, p = {pj}∞j=1 and {αj}∞j=1 satisfy conditions (a) to (e) of Lemma 5.7. Denote by b the matrix (δij +bbij)i,j∈N and by B the transformation of R∞ induced by b. Let s = {s(k)}∞k=1 ⊆ N be an increasing sequence. Assume (5.17) that (f) there exists ρ > 0 such that det bk ∈ [ρ, +∞) for all k ∈ N. Then B is invertible, B−1 is nonsingular and CB−1 ∈ S∗n,r. Proof. Note that, by [3, p. 46], (f) and Lemma 5.7, det b = limk→∞ det bk ∈ [ρ,∞) and det(bi)l 6= 0 for i ∈ In+r and sufficiently large l ∈ N. Hence, by (4) of Lemma 5.7, the transfor- mation B is invertible. According to (3) of Lemma 5.7 and Corollary 5.8 there exists eC > 0 such that the inequality (5.17) holds for every k ∈ In+r. Hence the formula hi(cid:16){xj}∞j=1(cid:17) = det bi exp 1 2(cid:16) ∞Xj=1 j(cid:17), x2 j − (bix)2 {xj}∞j=1 ∈ ℓ2(p), defines a B(R∞)-measurable function on R∞ for every i ∈ In+r. Now, let us choose κ ∈ N so that 1 − 2eCpj > 0 for every j > s(κ), and let {Ωk : k ∈ N} be given by Ωk = [−k, k]s(κ). Then, employing change-of-variable theorem and the fact that ℓ2(p) is a set of full measure µG in R∞, we get 1 (5.18) 1 , x2 x2 (cid:17) dm1 ℓ2(p)(cid:17) dµG (1 − 2eCpj)x2 2 j ZΩk×R∞ =ZΩk =ZΩk exp(cid:16) ∞Xj=1 s(κ)Xj=1 exp(cid:16)eC s(κ)Xj=1 exp(cid:16)eC j(cid:17) dµG 6ZΩk×R∞ exp(cid:16)eC(cid:13)(cid:13){xj}∞j=1(cid:13)(cid:13)2 √2πZR exp(cid:16) − ∞Yj=s(κ)+1 ∞Yj=s(κ)+1 q1 − 2eCpj 1√1−2 eCpj j − (bix)2 x2 j pj(cid:17) dµG,s(κ) · j pj(cid:17) dµG,s(κ) · Since p ∈ ℓ1(N), we deduce that Q∞j=s(κ)+1 is convergent (cf. [21, Chapter VII]). Thus, by (5.18), the function χΩk×R∞ hi belongs to L2(µG) for all i ∈ In+r and k ∈ N. Note that there exists C > 0 such that j(cid:12)(cid:12)(cid:12) 6 C s(l)Xj=1 (see the proof of assertion (3) of Lemma 5.7). Hence arguing as in (5.18) we show that the function belongs to L2(µG,s(l)) for all i ∈ In+r, k ∈ N and every l ∈ N such that l > κ, χΩk×R∞ ∆h l is a transformation of Rs(l) induced by (bi)l. Since, for every x = {xi}∞i=1 ∈ ℓ2(p) we where Bi have j −(cid:0)(bi)lπs(l)x(cid:1)2 x ∈ Rs(l), l ∈ N, k ∈ In+r, i ∈ In+r, k ∈ N. x2 j −(cid:0)(bk)lx(cid:1)2 s(l)Xj=1(cid:16)x2 j(cid:17) − s(l)Xj=1 x2 j pj, (5.19) B−i l (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) ∞Xj=1(cid:16)x2 =(cid:12)(cid:12)(cid:12) j(cid:17)(cid:12)(cid:12)(cid:12) j − (bix)2 s(l)Xj=s(l)−iη x2 j(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12) j −(cid:0)(bi)lπs(l)x(cid:1)2 ∞Xj=s(l)−iη x2 j − (bix)2 j(cid:12)(cid:12)(cid:12). 16 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA This combined with (5.17) and (5.19) proves that hi(x) = lim l→∞ ∆h B−i l (x), x ∈ ℓ2(p). By (5.18) and (5.19), the function Hk,i : ℓ2(p) → C given by Hk,i(x) = sup l∈Nn(cid:12)(cid:12) det(bi)l(cid:12)(cid:12)oχΩk×R∞ exp 1 2(cid:16)Ckxk2 ℓ2(p)(cid:17) is a majorant in L2(µG) for the sequence {χΩk×R∞ ∆h Lebesgue dominated convergence theorem we get B−i l }∞l=1, k ∈ N and i ∈ In+r. Hence, by the (5.20) χΩk×R∞ hi = lim l→∞ χΩk×R∞ ∆h B−i l , k ∈ N, i ∈ In+r. Since for all i ∈ In+r and l ∈ N, the matrices bi are banded and the matrices (bi)l are invertible, we deduce from (5.20) that for all i ∈ In+r, k, m ∈ N, and σ ∈ B(Rm), we have µG(cid:18)Bi(cid:0)(σ × R∞) ∩ (Ωk × R∞)(cid:1)(cid:19) = lim µG,s(l)(cid:18)Bi l→∞Zσ×Rs(l)−m =Zσ×R∞ = lim l→∞ χΩk×R∞ hi dµG. l(cid:0)(cid:0)σ × Rs(l)−m) ∩ (Ωk × Rs(l)−s(κ)(cid:1)(cid:1)(cid:19) χΩk×Rs(l)−s(κ) h B−i l dµG,s(l) This, by [2, Theorem 10.3], implies that for every i ∈ In+r, the transformation B−i is nonsingular and hi = h B−i a.e. [µG]. This, Theorem 4.1, Lemma 4.2 and equality (5.20) imply that conditions (i) to (vii) of Theorem 5.1 hold with A = B−1 and Al = B−1 , l ∈ N (conditions (vi) and (vii) follow easily from η-boundedness of b). Hence, applying Theorem 5.1, we get that CB−1 belongs to S∗n,r. (cid:3) l Remark 5.10. Regarding Proposition 5.9, we note that there is an another way of producing the inverses of an η-bounded matrix b and a transformation B via inductive technique. To this end, instead of conditions (a)-(e) of Lemma 5.7, we assume that (1) there exists ρ > 0 such that σ(bk) ⊆ [ρ, +∞) for all k ∈ N, (2) for every k ∈ N, bk is normal in (Rs(k), · ), (3) for every ε > 0 there is k0 ∈ N such that for every m > l > k0 s(l) x + blx(cid:1), bmιs(m) t : Rt ∋ (x1, . . . , xt) 7→ (x1, . . . , xt, 0, . . . , 0) ∈ Rs for t 6 s. s(l) blx 6 ε(cid:0)x + bmιs(m) s(l) x − ιs(m) where ιs x ∈ Rs(l), Then we proceed as follows. For j ∈ N, let Bj denote the linear mapping Rs(j) → Rs(j) induced by bj. By [18, Theorem 1.1, Lemma 4.3] and (3), the sequence {Bj}∞j=1 induces a closable densely defined operator B∞ acting on ℓ2(N) according to the formula D(B∞) = [k∈N {ιkx : x ∈ Rk such that ιs(j)bjιs(j) k x, B∞ιkx = lim j→∞ lim j→∞ ιs(j)bjιs(j) k x ∈ ℓ2(N)} ιkx ∈ D(B∞), where ιt : Rt ∋ (x1, . . . , xt) 7→ (x1, . . . , xt, 0, . . .) ∈ R∞ for t ∈ N. B∞ is the inductive limit of {Bj}∞j=1. In view of [18, Corollary 2.2, Lemma 4.3] and (1), σ(B∞) ⊆ [ρ, +∞) and thus B∞ is invertible. From this point we proceed as in the proof of (4) of Lemma 5.7. With help of Proposition 5.9 we can deliver an example of composition operators belonging to S∗n,r induced by non-diagonal transformation of R∞. The operator is induced by a 1-banded matrix. COMPOSITION OPERATORS VIA INDUCTIVE LIMITS 17 Example 5.11. Let n, r ∈ Z+. Let b be a matrix given by 0 0 0 q4 ··· q1 1 q2 0 ··· 0 0 q3 1 ··· 0 q2 1 q3 ··· 1 q1 0 0 ···   , ··· ··· ··· ··· ··· b =   √2 2 (cid:1). Let {s(k)}∞k=1 be given by s(k) = k. Set p = {qj}∞j=1 and define αj = qj−1, with q ∈ (cid:0)0, j ∈ N. Clearly, det bl = det bl−1 − q2l−2 det bl−2, l > 3, which implies that 1 −P∞k=2 q2k−2 < det bl < 1 for l > 2. As a consequence, the assumptions of Proposition 5.9 are satisfied. Hence b induces an invertible transformation B of R∞ such that B−1 is nonsingular and CB−1 belongs to S∗n,r. If fact, CB−1 ∈ S∗n,r for every n, r ∈ Z+. Now, let X ⊆ L2(µG) be the family defined as in the Step 5 of proof of Theorem 5.1 with A = B−1. According to Lemma 3.2 and (5.10) we see that X ⊆ D(cid:16)(cid:0)C∗B−1(cid:1)n(cid:17) for every n ∈ Z+. Arguing as in the proof of Step 5 of Theorem 5.1 and using [4, Lemma 3.2] we see that X is linearly dense in L2(µG). Thus every power of C∗B−1 is densely defined. This combined with [10, Theorem 52] proves that D∞(C∗B−1 ) is dense in L2(µG). On the other hand, by Proposition 2.3, the operator C∗B−1D∞(C ∗ Remark 5.12. Concluding the paper we point out that Theorem 5.1 and Propositions 5.2 and 5.9 rely essentially on the very precise knowledge of the Radon-Nikodym derivative hA, which is due to the inequality (5.9). It seems desirable to look for some inductive-limit-based criteria for cosubnormality, which would be independent of the knowledge of hA. B−1 ) is subnormal. References [1] Y. M. Berezansky, Y. G. Kondratiev, Spectral methods in infinite-dimensional analysis, Mathematical Physics and Applied Mathematics, vol. 12, Springer, 1995. [2] P. Billingsley, Probability and measure, John Wiley & Sons, Inc., New York 1979. [3] A. Bottcher, S. M. Grudsky, Spectral Properties of Banded Toeplitz Matrices, SIAM, 2005. [4] P. Budzy´nski, P. Dymek, A. P laneta, Unbounded composition operators via inductive limits: cosubnormal operators with matrix symbols, Filomat (to appear). [5] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Unbounded subnormal weighted shifts on directed trees, J. Math. Anal. Appl. 394 (2012), 819-834. [6] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Unbounded subnormal weighted shifts on directed trees. II, J. Math. Anal. Appl. 398 (2013), 600-608. [7] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, On unbounded composition operators in L2-spaces, Ann. Mat. Pura Appl. 193 (2014), 663-688. [8] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, A multiplicative property characterizes quasinormal composition operators in L2-spaces, J. Math. Anal. Appl. 409 (2014), 576-581. [9] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Unbounded subnormal composition operators in L2-spaces, J. Funct. Anal., 269 (2015), 2110-2165. [10] P. Budzy´nski, Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Unbounded subnormal weighted composition operators in L2-spaces, in preparation, http://arxiv.org/abs/1310.3542. [11] P. Budzy´nski, A. P laneta, Dense definiteness and boundedness of composition operators in L2-spaces via inductive limits, Oper. Matrices (to appear). [12] J. T. Campbell, W. E. Hornor, Seminormal composition operators. J. Operator Theory 29 (1993), 323-343. [13] D. Cicho´n, J. Stochel, F. H. Szafraniec, Extending positive definiteness, Trans. Amer. Math. Soc. 363 (2011), 545-577. [14] J. B. Conway, The theory of subnormal operators, Mathematical Surveys and Monographs, Providence, Rhode Island, 1991. [15] R. E. Curto, Joint hyponormality: a bridge between hyponormality and subnormality, Proc. Sym. Math. 51 (1990), 6991. [16] S. Demko, W. F. Moss, P. W. Smith, Decay rates for inverses of band matrices, Math. Comp. 43 (1984), 491-499. [17] Z. Jab lo´nski, Hyperexpansive composition operators, Math. Proc. Camb. Phil. Soc. 135 (2003), 513-526. [18] J. Janas, Inductive limit of operators and its applications, Studia Math. 90 (1988), 87-102. [19] I. Jung, M. Lee and S. Park, Separating classes of composition operators, Proc. Amer. Math. Soc. 135 (2007), 39553965. 18 P. BUDZY ´NSKI, P. DYMEK, AND A. P LANETA [20] I. B. Jung, S. H. Park, J. Stochel, L(n)-hyponormality: a missing bridge between subnormality and paranor- mality, J. Aust. Math. Soc. 88 (2010), 193-203. [21] K. Knopp, Theory and Application of Infinite Series (English translation), Dover Publications, 1990. [22] S. McCullough and V. I. Paulsen, k-hyponormality of weighted shifts, Proc. Amer. Math. Soc. 116 (1992), 165-169. [23] W. Rudin, Real and Complex Analysis, McGraw-Hill, 1987. [24] R. K. Singh, J. S. Manhas, Composition Operators on Function Spaces, North-Holland, 1993. [25] J. Stochel, Seminormal composition operators on L2 spaces induced by matrices, Hokkaido Math. J. 19 (1990), 307-324. [26] J. Stochel, F. H. Szafraniec, On normal extensions of unbounded operators. I, J. Operator Theory 14 (1985), 31-55. [27] J. Stochel, F. H. Szafraniec, On normal extensions of unbounded operators. II, Acta Sci. Math. (Szeged), 53 (1989), 153-177. [28] J. Stochel, F. H. Szafraniec, On normal extensions of unbounded operators. III. Spectral properties, Publ. RIMS, Kyoto Univ. 25 (1989), 105-139. [29] J. Stochel, F. H. Szafraniec, The complex moment problem and subnormality: a polar decomposition approach, J. Funct. Anal. 159 (1998), 432-491. [30] C. F. Sturm, Cours d'analyse de l'´ecole polytechnique, Gauthier-Villars, 1868. [31] J. Weidmann, Linear operators in Hilbert spaces, Springer-Verlag, Berlin, Heidelberg, New York, 1980. Katedra Zastosowa´n Matematyki, Uniwersytet Rolniczy w Krakowie, ul. Balicka 253c, 30-198 Krak´ow, Poland E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected]
1109.3304
1
1109
2011-09-15T09:58:21
On compactness of Laplace and Stieltjes type transformations in Lebesgue spaces
[ "math.FA" ]
We obtain criteria for integral transformations of Laplace and Stieltjes type to be compact on Lebesgue spaces of real functions on the semiaxis.
math.FA
math
ON COMPACTNESS OF LAPLACE AND STIELTJES TYPE TRANSFORMATIONS IN LEBESGUE SPACES ELENA P. USHAKOVA Abstract. We obtain criteria for integral transformations of Laplace and Stieltjes type to be compact on Lebesgue spaces of real func- tions on the semiaxis. 1. INTRODUCTION Let Lr(I) denote the Lebesgue space of all measurable functions f (x) integrable to a power 0 < r < ∞ on an interval I ⊆ [0, ∞) =: R+, that is If I = R+ we write Lr := Lr(R+), and kf kr means kf kr,R+. For r = ∞ we denote f (x)rdx(cid:17)1/r < ∞o. Lr(I) =nf : kf kr,I :=(cid:16)ZI L∞(I) =nf : kf k∞,I := ess sup t∈I f (t) < ∞o. Take λ > 0, p ≥ 1, q > 0 and put p′ := p/(p − 1). Assume v ∈ Lp′ loc(R+) and w ∈ Lq In this article we study compactness properties of two particular cases of an integral transformation T : Lp → Lq of the form loc(R+) are weight functions. kT (x, y)f (y)v(y)dy, x ∈ R+, with a non-negative kernel kT (x, y) decreasing in variable y. We take as T the Laplace integral operator (1.1) (1.2) T f (x) := w(x)ZR+ Lf (x) :=ZR+ e−xyλ f (y)v(y)dy, x ∈ R+, 1991 Mathematics Subject Classification. 47G10. Key words and phrases. Compactness, Boundedness, Lebesgue space, Integral operator, Laplace transformation, Stieltjes transformation. 1 2 ELENA P. USHAKOVA with the outer weight function w(x) = 1 and kL(x, y) := e−xyλ, and a Stieltjes type transformation of the form (1.3) Sf (x) := w(x)ZR+ f (y)v(y)dy xλ + yλ , x ∈ R+, with kS(x, y) := (xλ + yλ)−1. These operators are related to each other by (4.7). With an appropriate choice of λ, v and w transformations L and S become special cases of conventional convolution transformation F (x) = −∞ f (t)G(x − t)dt, −∞ < x < ∞ [27, Ch.8, §§ 8.5, 8.6]. The Stieltjes type operator (1.3) has also connections with Hilbert's double series R ∞ theorem (see [4] for details). Some interesting properties and applica- tions of the Laplace type transform (1.2) to differential equations are indicated in [7, Ch. 5]. In this work we find explicit necessary and sufficient conditions for Lp − Lq -- compactness of L and S expressed in terms of kernels kL, kS, weight functions v, w and properties of the Lebesgue spaces. The re- sults may be useful for study of characteristic values of the transfor- mations. All cases of summation parameters p ≥ 1 and q > 0 are considered. If 0 < p < 1 then T : Lp → Lq is compact in trivial case only (see [18, Theorem 2]). Note that L2 − L2 compactness of (1.2) and (1.3) was studied in [25, 26]. We generalize these results for all positive p and q. Our main method is well-known and consists in splitting an initial operator into a sum of a compact operator and operators with small norms (see e.g. [8], [13], [16]). The article is organized as follows. Section 3 is devoted to the com- pactness of the Laplace transformation (1.2). Criteria for the compact- ness of the Stieltjes operator (1.3) appear in Section 4. Note that the case of negative λ in S ensues from the results for positive λ by simple modification of the weight functions v and w. In Section 5 we discuss cases p = ∞ and q = ∞. Some auxiliary results are collected in Section 2. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 3 Throughout the article we assume v, w to be non-negative. Products of the form 0 · ∞ are supposed to be equal to 0. An equivalence A ≈ B means either A = c0B or c1A ≤ B ≤ c2A, where ci, i = 0, 1, 2 are constants depending on λ, p, q only. The symbol Z denotes integers, χE stands for a characteristic function of a subset E ⊂ R+. In addition, we use =: and := for marking new quantities. 2. PRELIMINARIES In this part we set auxiliary results concerning boundedness of the transformations and adduce some known results on compactness of integral operators we shall need later on in our proofs. 2.1. Boundedness. We start from a statement for an integral opera- tor T from Lp to Lq, when p = 1. Theorem 2.1. [23, Theorem 3.2] Suppose 0 < q < ∞ and a non- negative kernel kT (x, y) of the operator (1.1) is non-increasing in y for each x. For all f ≥ 0 the best constant C in the inequality (2.1) (cid:18)ZR+(cid:18)ZR+ kT (x, y)f (y)v(y)dy(cid:19)q wq(x)dx(cid:19)1/q ≤ CZR+ is unchanged when v(y) is replaced by ¯v0(y), where f (y)dy ¯vc1(t) := ess sup c1<x<t v(x). Consider the Laplace operator L. Denote r := p q p−q , q′ := q q−1, Vc1(t) :=Z t c1 vp′ , e−xyλ f (y)v(y)dy, αq := min{2, 2q−1}, 1 < q ≤ 2, q > 2, 2 q−1, 2q−1, βq := f (t)v(t)dt, F (x) :=Z c2 2 (cid:1)1/q(cid:2)Vc1(t)(cid:3)1/p′ c1 c1 Fc1(y) :=Z y AL,hc1,c2i(t) :=(cid:0)t−λ−c−λ AL := AL,R+, BL,hc1,c2i :=(cid:18)Z c2 , AL,hc1,c2i := sup AL,hc1,c2i(t), c1<t<c2 c1 ht−λ − c−λ 2 ir/q(cid:2)Vc1(t)(cid:3)r/q′ vp′ (t)dt(cid:19)1/r , 4 ELENA P. USHAKOVA BL := BL,R+ =(cid:18)ZR+ Bq(t) := t−λ/qv(t), ¯Bq(t) := t−λ/q ¯v0(t), Bp :=(cid:18)ZR+ BL(t)dt(cid:19)1/r , BL(t) := t−λr/q(cid:2)V0(t)(cid:3)r/q′ 2 ]q/(1−q)t−λ−1¯vc1(t)q/(1−q)dt(cid:19)(1−q)/q [t−λ − c−λ y−λp′ vp′ vp′ , (y)dy(cid:19)1/p′ , (t), Bq′,hc1,c2i :=(cid:18)Z c2 Bq′ = Bq′,R+ =(cid:18)ZR+ c1 , Bq′(t) := t−λ/(1−q)−1¯v0(t)q/(1−q), Bq′(t)dt(cid:19)(1−q)/q Dhc1,c2i := c−λ/q 2 (cid:2)Vc1(c2)(cid:3)1/p′ . Various conditions were found for the boundedness of the Laplace transformation (1.2) in Lebesgue spaces (see e.g. [3], [5]). Convenient for our purposes Lp − Lq criterion for L was obtained in [24, Theorem 1] (see also [17, Theorem 1]). Theorem 2.2. [17, 24] The following estimates are true for the norm of L : (i) Denote α1 := αq−2/q and β1 := β(q′)1/p′. If 1 < p ≤ q < ∞ then α1AL ≤ kLkLp→Lq ≤ β1AL. (ii) Let 1 ≤ q < p < ∞. If q = 1 then kLkLp→L1 = Bp. If q > 1 then α2BL ≤ kLkLp→Lq ≤ β2BL with α2 := α(p′q/r)1/q′q−1/q, β2 := β(p′)1/q′. (iii) Put α3 := q−1/q, β3 := p1/p(p′)1/q′q−2/qr1/r. If 0 < q < 1 < p < ∞ then α3 kBqkp′ ≤ kLkLp→Lq ≤ β3BL. (iv) Let 0 < q ≤ 1 = p. If 0 < q < 1 then α4 ess sup Bq(t) ≤ kLkL1→Lq ≤ β4Bq′, t∈R+ where α4 := q−1/q and β4 := λ(1−q)/qq−2/q(1 − q)−(1−q)/q. If q = 1 then 2−λ ess sup t∈R+ (2.2) t−λ¯v0(t) ≤ kLkL1→L1 t−λ¯v0(t) ≤ 2−λ sup t∈R+ t−λ¯v0(t) ≤ sup t∈R+ ≤ ess sup t∈R+ t−λ¯v0(t). ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 5 Remark 2.3. Integration by parts gives the equality Br L = (λp′/q)ZR+(cid:2)V0(t)(cid:3)r/p′ t−λr/q−1dt, which is true either v ∈ Lp′[0, t], t > 0, or q > 1 (see [23, Remark, p. 8]). Remark 2.4. Bq(t) in (iv) may be replaced by ¯Bq(t) (see Theorem 2.1). Remark 2.5. The lower estimates in (2.2) can be proved by applying a function ft(y) = t−1χ(t,2t)(y), t > 0 into (2.1) with v replaced by ¯v0. The results (i) and (ii) of Theorem 2.1 rest on [24, Lemma 1]. The following statement is its modification for the Laplace operator of the form f → L(f χhc1,c2i), where 0 ≤ c1 < c2 ≤ ∞ and h·, ·i denotes any of intervals (·, ·), [·, ·], [·, ·) or (·, ·]. Lemma 2.6. Let 1 < q < ∞. Assume f ≥ 0 and suppose the following lim t→c1 t−λ(cid:2)Vc1(t)(cid:3)q/p′ = 0, lim t→c2 conditions are satisfied: RR+ F q(x)dx < ∞, t−λ(cid:2)Vc1(t)(cid:3)q/p′ αqλq−2Z c2 F q c1(y)y−λ−1dy + αqq−2F q < ∞. Then c1 ≤ (βqλ/q)Z c2 c1 c1(c2)c−λ F q(x)dx 2 ≤ZR+ F q c1(y)y−λ−1dy + (βq/q)F q c1(c2)c−λ 2 . Lemma 2.6 modifies Theorem 2.2 as follows. Theorem 2.7. (i) Let 1 < p ≤ q < ∞. The operator L is bounded from Lphc1, c2i to Lq if and only if AL,hc1,c2i + Dhc1,c2i < ∞. Moreover, α1(cid:2)AL,hc1,c2i + Dhc1,c2i(cid:3) ≤ kLkLphc1,c2i→Lq ≤ β1(cid:2)AL,hc1,c2i + Dhc1,c2i(cid:3). (ii) If 1 < q < p < ∞ then L is bounded iff BL,hc1,c2i + Dhc1,c2i < ∞, where α2(cid:2)BL,hc1,c2i + Dhc1,c2i(cid:3) ≤ kLkLphc1,c2i→Lq ≤ β2(cid:2)BL,hc1,c2i + Dhc1,c2i(cid:3). 6 ELENA P. USHAKOVA (iii) Let 0 < q < 1 < p < ∞. The Laplace operator L is bounded from Lphc1, c2i to Lq if BL,hc1,c2i + Dhc1,c2i < ∞. If L : Lphc1, c2i → Lq is bounded then kBqkp′,hc1,c2i < ∞. We also have α3 kBqkp′,hc1,c2i ≤ kLkLphc1,c2i→Lq ≤ β3(cid:2)BL,hc1,c2i + Dhc1,c2i(cid:3). (iv) Let 0 < q < 1 = p. If L is L1 − Lq -- bounded then ess sup t∈hc1,c2i Bq(t) < ∞. The operator L is bounded from L1 to Lq if Bq′,hc1,c2i < ∞ and Dhc1,c2i < ∞. Besides, α4 ess sup t∈hc1,c2i Bq(t) ≤ kLkL1hc1,c2i→Lq ≤ β4Bq′,hc1,c2i + q−2/qDhc1,c2i. Remark 2.8. The constants αi, βi, i = 1, 2, 3, 4, in Theorem 2.7 are defined as in Theorem 2.2. Moreover, if hc1, c2i = R+ and (2.3) (i) lim t→0 AL(t) = 0, (ii) lim t→∞ AL(t) = 0 we have DR+ = 0 and Theorem 2.2 has become a case of Theorem 2.7. Now we start to consider the Stieltjes operator S. Add some deno- tations Vt(∞) :=Z ∞ t vp′(y)dy yλp′ , Wc1(t) :=Z t c1 wq, Wt(c2) :=Z c2 t wq(x)dx xλq . Boundedness criteria for S in Lebesgue spaces were found in [2, 10, 21]. The following Theorem 2.9 contains some of them. Theorem 2.9. (i) [2, Theorem 1] The operator S is bounded from Lp to Lq for 1 < p ≤ q < ∞ if and only if AS := supt∈R+ AS(t) < ∞ with AS(t) := tλ(cid:18)ZR+ wq(x)dx (xλ + tλ)q(cid:19)1/q(cid:18)ZR+ vp′(y)dy (tλ + yλ)p′(cid:19)1/p′ . Moreover, AS ≤ kSkLp→Lq ≤ γS · AS with a constant γS depending on p, q and λ only. If p = 1 ≤ q < ∞ then kSkLp→Lq ≈ A1,S, where A1,S := sup t∈R+ tλ(cid:18)ZR+ wq(x)dx (xλ + tλ)q(cid:19)1/q ess sup y∈R+ v(y) tλ + yλ . ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 7 (ii)[21, Theorem 2.1] S is bounded from Lp to Lq for 1 < q < p < ∞ if and only if BS < ∞ with BS :=(cid:18)ZR+(cid:18)ZR+ wq(x)dx (xλ + tλ)q(cid:19)r/q(cid:18)ZR+ vp′(y)dy (1 + [y/t]λ)p′(cid:19)r/q′ vp′ (t)dt(cid:19)1/r . Besides, kSkLp→Lq ≈ BS with constants of equivalence depending, pos- sibly, on p, q and λ. If q = 1 then kSkLp→L1 =(cid:18)ZR+(cid:18)ZR+ w(x)dx xλ + tλ(cid:19)p′ vp′ (t)dt(cid:19)1/p′ . In view of the relation (2.4) 1 2 [Hf (x) + H ∗f (x)] ≤ Sf (x) ≤ Hf (x) + H ∗f (x), f ≥ 0, some properties of S can be interpreted through the Hardy operator Hf (x) := x−λw(x)Z x 0 f (y)v(y)dy and its dual transformation H ∗f (x) := w(x)R ∞ x f (y)y−λv(y)dy. Alter- native criteria for Lp − Lq -- boundedness of S, in comparison with The- orem 2.9, follow from results for Hardy integral operators and cover even the case 0 < q < 1 (see theorems 4.1 and 4.2 for details). Theorem 2.10. (i) If 1 < p ≤ q < ∞ then the operator S is bounded from Lp to Lq if and only if AH + AH ∗ < ∞, where kSkLp→Lq ≈ AH + AH ∗ and (2.5) (2.6) AH := sup t∈R+ AH ∗ := sup t∈R+ AH (t) := sup AH ∗(t) := sup t∈R+(cid:2)V0(t)(cid:3)1/p′(cid:2)Wt(∞)(cid:3)1/q, t∈R+(cid:2)Vt(∞)(cid:3)1/p′(cid:2)W0(t)(cid:3)1/q . (ii) If 0 < q < 1 < p < ∞ or 1 < q < p < ∞ then S : Lp → Lq if and only if BH + BH ∗ < ∞, where kSkLp→Lq ≈ BH + BH ∗ and BH :=(cid:18)ZR+(cid:2)V0(t)(cid:3)r/p′(cid:2)Wt(∞)(cid:3)r/pt−λqwq(t)dt(cid:19)1/r BH ∗ :=(cid:18)ZR+(cid:2)Vt(∞)(cid:3)r/p′(cid:2)W0(t)(cid:3)r/pwq(t)dt(cid:19)1/r , . 8 ELENA P. USHAKOVA Besides, if q > 1 then BH = (q/p′)1/r(cid:18)ZR+(cid:2)V0(t)(cid:3)r/q′(cid:2)Wt(∞)(cid:3)r/p BH ∗ = (q/p′)1/r(cid:18)ZR+(cid:2)Vt(∞)(cid:3)r/q′(cid:2)W0(t)(cid:3)r/qt−λp′ vp′ vp′ , (t)dt(cid:19)1/r (t)dt(cid:19)1/r . (iii) Let 0 < q ≤ 1 = p. If 0 < q < 1 then the Stieltjes transformation S is L1 − Lq -- bounded if and only if B1,H + B1,H ∗ < ∞, where kSkL1→Lq ≈ B1,H + B1,H ∗ and B1,H :=(cid:18)ZR+ ¯v0(t)q/(1−q)(cid:2)Wt(∞)(cid:3)q/(1−q)t−λqwq(t)dt(cid:19)(1−q)/q B1,H ∗ :=(cid:18)ZR+(cid:2)t−λ¯vt(∞)(cid:3)q/(1−q)(cid:2)W0(t)(cid:3)q/(1−q)wq(t)dt(cid:19)(1−q)/q ¯vt(∞)t−λZ t x−λw(x)dx + sup t∈R+ ¯v0(t)Z ∞ kSkL1→L1 ≈ sup t∈R+ 0 t , . If q = 1 then w(x)dx. We conclude the paragraph by giving a general boundedness criterion for an integral operator T defined by (1.1) and acting from L1 to Lq, when 1 < q < ∞. Theorem 2.11. [11, Ch. XI, §1.5, Theorem 4] Let 1 < q < ∞. The operator T with measurable on R+ × R+ kernel kT (x, y) ≥ 0 is bounded from L1 to Lq if and only if kkT kL∞[Lq] := ess sup kw(·)kT (·, t)v(t)kq < ∞. t∈R+ Besides, kT kL1→Lq = kkT kL∞[Lq]. 2.2. Compactness. Suppose I ⊂ R+ and J ⊂ R+ are intervals of finite Lebesgue measure, that is mes I :=RI dx < ∞ and mes J < ∞. Let K be an integral operator from Lp(I) to Lq(J) of the form Kf (x) =ZI k(x, y)f (y)dy, x ∈ J. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 9 Assume K0 : Lp(I) → Lq(J) is a positive operator K0f (x) =ZI k0(x, y)f (y)dy, x ∈ J, such that Kf ≤ K0f . Then K is called a regular operator (see [12, § 2.2] or [19, Definition 3.5]) with a majorant operator K0. Note that the last inequality ensues from the expression k(x, y) ≤ k0(x, y) (see [12, § 5.6]). Below we adduce two remarkable results on compactness of linear regular integral operators by M.A. Krasnosel'skii, P.P. Zabreiko, E.I. Pustyl'nik and P.E. Sobolevskii [12]. The first theorem gives conditions for the compactness of integral operators from Lp(I) to Lq(J), when 0 < q ≤ 1 < p < ∞. Theorem 2.12. [12, p. 94, Theorem 5.8] Every linear regular integral operator K : Lp(I) → Lq(J) is compact, when 0 < q ≤ 1 < p < ∞. The second theorem is on compactness of linear regular integral op- erators with compact majorants. The statement as well as Theorem 2.12 is true for any I ⊆ R+ and J ⊆ R+. Theorem 2.13. [12, p. 97, Theorem 5.10] Take k(x, y) ≤ k0(x, y) and suppose K0 is compact from Lp(I) to Lq(J) for 1 < p ≤ ∞ and 1 ≤ q < ∞. Then the operator K : Lp(I) → Lq(J) is compact as well. The next theorem is on relative compactness of a subset of Lq(R+), 1 ≤ q < ∞. Theorem 2.14. [1, Theorem 2.21] Let 1 ≤ q < ∞ and Ω be a bounded subset of Lq(R+). Then Ω is relatively compact if and only if for all ε > 0 there exists δ > 0 and a subset I ⊂ R+ such that for every g ∈ Ω and every h ∈ R+ with h < δ g(x + h) − g(x)q dx < εq, g(x)q dx < εq. ZR+\I ZR+ We finalize the section by giving a statement obtained in [13, Lemma 4] for Banach function spaces X and Y with absolutely continuous (AC) 10 ELENA P. USHAKOVA norms. A space X has AC-norm if for all f ∈ X, kf χEnkX → 0 for every sequence of sets {En} ⊂ R+ such that χEn(x) → 0 a.e. We shall apply this statement when X = Lp, 1 < p < ∞, and Y = Lq, 1 ≤ q < ∞. Lemma 2.15. [13, Lemma 4] Let 1 < p < ∞, 1 ≤ q < ∞ and T : Lp → Lq be an integral operator of the form (1.1). T is compact if (2.7) MT :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)w(x)k(x, ·)v(·)(cid:13)(cid:13)p′(cid:13)(cid:13)(cid:13)q < ∞. 3. COMPACTNESS OF THE LAPLACE TRANSFORM Theorem 3.1. (i) If 1 < p ≤ q < ∞ then the operator L : Lp → Lq is compact if and only if AL < ∞ and the conditions (2.3) are fulfilled. (ii) Let 1 ≤ q < p < ∞. If q > 1 then L : Lp → Lq is compact if and only if BL < ∞. If q = 1 then L is compact if and only if Bp < ∞. (iii) Let 0 < q < 1 < p < ∞. The operator L : Lp → Lq is compact if BL < ∞. If L is compact from Lp to Lq then kBqkp′ < ∞. (iv) Let 0 < q < 1 = p. L is compact from L1 to Lq if Bq′ < ∞. If L is L1 − Lq -- compact then ess supt∈R+ Bq(t) < ∞. (v) Let 1 = p ≤ q < ∞. The operator L is L1 − Lq-compact if and only if ess sup t∈R+ ¯Bq(t) < ∞ and lim t→0 ¯Bq(t) = lim t→∞ ¯Bq(t) = 0. Proof. (i) Sufficiency. Suppose AL < ∞ and lim t→0 0. Put 0 < a < b < ∞ and denote AL(t) = lim t→∞ AL(t) = L0f := L(f χ(a,b)), L1f := L(f χ[0,a]), L2f := L(f χ[b,∞)). Obviously, (3.1) Lf (x) = Lif (x). 2Xi=0 Since AL < ∞ then L is bounded from Lp to Lq by Theorem 2.2(a). This yields Lp − Lq -- boundedness of the operator L0f, which is regular with a majorant operator L0f bounded from Lp to Lq as well. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 11 According to Lemma 2.15 the operator L0 : Lp → Lq is compact if (3.2) ML0 :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)χ(a,b)(·)kL(x, ·)v(·)(cid:13)(cid:13)p′(cid:13)(cid:13)(cid:13)q Since 0 < a < b < ∞ and vp′ ∈ Lloc(R+) we have < ∞. (3.3) M q L0 ≤ 1 aλq(cid:2)Va(b)(cid:3)q/p′ < ∞. Therefore, L0 is compact from Lp to Lq for any 0 < a < b < ∞. Now consider the operators Li, i = 1, 2. By Theorem 2.7(a) we have: (3.4) (3.5) kL1kLp→Lq ≤ 2β1 sup 0≤t≤a kL2kLp→Lq ≤ β1 sup b≤t<∞ The conditions (2.3) yield , t−λ/q(cid:2)V0(t)(cid:3)1/p′ t−λ/q(cid:2)Vb(t)(cid:3)1/p′ . lim a→0 sup 0≤t≤a lim b→∞ sup b≤t<∞ t−λ/q(cid:2)V0(t)(cid:3)1/p′ t−λ/q(cid:2)Vb(t)(cid:3)1/p′ = lim t→0 AL(t) = 0, ≤ lim t→∞ AL(t) = 0. Together with (3.4) and (3.5) this gives: (3.6) lim a→0 kL1kLp→Lq = 0, lim b→∞ kL2kLp→Lq = 0. Therefore, (3.1) implies (3.7) kL − L0kLp→Lq ≤ kL1kLp→Lq + kL2kLp→Lq, and now the operator L : Lp → Lq is compact as a limit of compact operators, when a → 0 and b → ∞. Necessity. Suppose now L is compact from Lp to Lq. Then L is Lp − Lq -- bounded and AL < ∞ by Theorem 2.2(a). To prove (2.3) we assume {zk}k∈Z ⊂ R+ is an arbitrary sequence. To establish the claim (i) in (2.3) suppose limk→∞ zk = 0 and put fk(t) = χ[0,zk](t)[v(t)]p′−1(cid:2)V0(zk)(cid:3)−1/p. Since kfkkp = 1 then (cid:12)(cid:12)(cid:12)(cid:12)ZR+ fk(y)g(y)dy(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:18)Z zk 0 g(y)p′ dy(cid:19)1/p′ → 0, k → ∞, 12 ELENA P. USHAKOVA for any g ∈ Lp′. Therefore, the sequence {fk}k∈Z converges weakly to 0 in Lp. Compactness of L : Lp → Lq yields strong convergence of {Lfk}k∈Z in Lq, that is limk→∞ kLfkkq = 0. Besides, we have 0 (cid:18)Z ∞ Z ∞ 0 e−xyλ fk(y)v(y)dy(cid:19)q dx ≥Z ∞ 0 e−qxzλ k dx(cid:2)V0(zk)(cid:3)q/p′ = Aq L(zk) q . Hence, limk→∞ AL(zk) = 0. Since {zk}k∈Z is arbitrary, then (2.3)(i) is proved. For the proof of (2.3)(ii) we suppose limk→∞ zk = ∞ and put gk(t) = χ[0,z−λ k ](t)zλ/q′ k . Since kgkkq′ = 1 we have → 0, k → ∞, for any f ∈ Lq, which means weak convergence of {gk}k∈Z in Lq′. Com- pactness of L : Lp → Lq, 1 < p, q < ∞, implies Lq′ − Lp′ -- compactness of the dual operator f (x)qdx(cid:19)1/q (cid:12)(cid:12)(cid:12)(cid:12)Z ∞ 0 k 0 f (x)gk(x)dx(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:18)Z z−λ L∗g(y) := v(y)Z ∞ 0 e−xyλ g(x)dx. Therefore, {L∗gk}k∈Z strongly converges in Lp′ : (3.8) We obtain lim k→∞ kL∗gkkp′ = 0. Z ∞ 0 vp′ ≥ e−p′ 0 e−xyλ (y)(cid:18)Z ∞ V0(zk)(cid:18)Z z−λ gk(x)dx(cid:19)p′ dx(cid:19)p′ 0 k dy ≥ V0(zk)(cid:18)Z z−λ 0 k e−xzλ k dx(cid:19)p′ zλp′/q′ k zλp′/q′ k = e−p′ z−λp′/q k V0(zk) = e−p′ Ap′ L (zk). Together with (3.8) this implies lim k→∞ AL(zk) = 0, and now the condition (2.3)(ii) is fulfilled by the arbitrariness of {zk}k∈Z. Necessity in (ii), (iii) and (iv) follows by Theorem 2.2 from the hy- pothesis of compactness and, therefore, boundedness of L. (ii) Sufficiency of the conditions BL < ∞ (if 1 < q < p < ∞) and Bp < ∞ (if q = 1) for the compactness of L is provided by Lemma ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 13 2.15 and Theorem 2.7(b). Namely, if 1 < q < p < ∞ then Lemma 2.15 yields Lp − Lq-compactness of L0 (see (3.3)), while norms L1 and L2 are estimated by Theorem 2.7(b) as follows: kL1kLp→Lq ≤ β2 maxn2,(cid:0)r/p′(cid:1)1/ro(cid:18)ZR+ kL2kLp→Lq ≤ β2(cid:18)ZR+ (3.9) χ[0,a](t)BL(t)dt(cid:19)1/r χ[b,∞)(t)BL(t)dt(cid:19)1/r , . Thus, the condition BL < ∞ and the estimate (3.7) implies compact- ness of L as a → 0, b → ∞. If q = 1 then L is compact by Lemma 2.15. Sufficiency in (iii) and (iv) can be established as follows. Let BL < ∞ if 0 < q < 1 < p < ∞ or Bq′ < ∞ if 0 < q < 1 = p. By Theorem 2.7(c) we obtain the estimate (3.9) for the case 0 < q < 1 < p < ∞ (with β3 instead of β2). By the part (iv) of the same theorem we have for 0 < q < 1 = p : kL2kLp→Lq ≤ β4(cid:18)ZR+ χ[b,∞)(t)Bq′(t)dt(cid:19)1/r . Thus, the condition BL < ∞ (or Bq′ < ∞) yields kL2kLp→Lq → 0 as b → ∞. Now consider the operator Lbf := L0f + L1f = L(f χ[0,b]). The hypothesis BL < ∞ (or Bq′ < ∞) suffices for the boundedness of L (see Theorem 2.2). Therefore, Lb is bounded as well. To prove the compactness of Lb we shall use an extension of Theorem 2.12 for the case when an operator K is acting to Lq on the whole R+. Similar to [12, Theorem 5.8] we consider first a set Mh := {f ∈ Lp(a, b) : f ≤ h}, where h is an arbitrary positive number. Under this condition and in view of mes[0, b] < ∞ the operator Lb is bounded from L∞[0, b] to Lq. Compactness of Lb : Mh → Lq can be proved similar to [12, Theorem 5.2]. It remains to note that the rest transformation Lb : {Lp[0, b] \ Mh} → Lq has a norm tending to 0 as h → +∞ (see [12, Theorem 5.8] for details). Therefore, the operator Lb : Lp[0, b] → Lq, 14 ELENA P. USHAKOVA 0 < q < 1 ≤ p < ∞, is compact as a limit of compact operators as h → +∞. Summing up we can claim that (1.2) is compact from Lp to Lq, 0 < q < 1 ≤ p < ∞, on the strength of kL2kLp→Lq → 0, b → ∞, compactness of Lb and (3.10) kL − LbkLp→Lq = kL2kLp→Lq. (v) Sufficiency. Suppose ess supt∈R+ ¯Bq(t) < ∞ and (3.11) Put (i) lim t→0 ¯Bq(t) = 0, (ii) lim t→∞ ¯Bq(t) = 0. Laf (x) := e−xaλZ b a f (y)v(y)dy, x ∈ R+, where 0 < a < b < ∞, and note that La is the operator of rank 1 with the norm kLakL1→Lq = q−1/qa−λ/q ¯va(b) < ∞. Besides, La is a majorant for the operator L0, which is L1−Lq -- bounded with the norm estimated as follows: kL0kL1→Lq = kLkL1(a,b)→Lq = q−1/q ess sup a<t<b ¯Bq(t) =: q−1/qM < ∞. Suppose {fn}n∈Z is an arbitrary bounded sequence in L1(a, b) and as- sume {fnk} is its Cauchy subsequence, that is for any ε0 > 0 there exists N(ε0) such that kfnk − fmk k1,(a,b) < ε0, nk, mk > N(ε0). any ε > 0 : Put Enk,mk(ε) :=nx ∈ R+ :(cid:12)(cid:12)L0fnk(x) − L0fmk (x)(cid:12)(cid:12) > εo. We have for ZEnk ,mk (ε) a e−xyλ dx ≤ ε−1ZR+(cid:12)(cid:12)(cid:12)Z b ≤ ε−1Z b a [fnk (y) − fmk (y)]v(y)dy(cid:12)(cid:12)(cid:12)dx y−λ(cid:12)(cid:12)fnk(y) − fmk (y)(cid:12)(cid:12)v(y)dy kfnk − fmk k1,(a,b). ≤ ε−1Ma−λ/q′ ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 15 If ε0 = εδaλ/q′M −1 then µLq (Enk,mk(ε)) < δ as nk, mk > N(ε, δ) for any ε > 0, δ > 0. Therefore, L0 is compact in measure. Thus, L0 is L1 −Lq -- compact as a transformation majorated by the compact operator La (see [12, Ch. 2, §5.6]). Further, kL1kL1→Lq = q−1/q ess sup t∈[0,a] kL2kL1→Lq ≤ q−1/q ess sup t∈[b,∞) Bq(t) ≤ q−1/q sup t∈[0,a] t−λ/q ¯vb(t) ≤ q−1/q sup t∈[b,∞) ¯Bq(t), ¯Bq(t). Since the conditions (3.11) are fulfilled we can state that L1 and L2 are operators with small norms, when a → 0, b → ∞. Together with compactness of L0 this implies the compactness of L from L1 to Lq for all 1 ≤ q < ∞. Necessity. Suppose L is L1−Lq -- compact. Then the claim ess sup ¯Bq(t) t∈R+ < ∞ follows from theorems 2.2(iv), 2.1 and 2.11. As for necessity of (3.11)(i), note that Lf = L(f χ[0,x−1/λ]) + L(f χ[x−1/λ,∞)) := Lxf + Lxf and Lx, Lx are compact. Besides, the condition (3.11)(i) is equivalent to (3.12) lim k→−∞ 2−λk/q¯v0(2k) = 0. Now suppose the contrary. Then, similar to [8, p. 84], given γ ∈ (0, 1) there is a sequence kj → −∞, some ε > 0 and functions fkj ≥ 0, kfkj kL1 ≤ 1, such that Z 2kj 0 fm(y)v(y)dy ≥ γ ¯v0(2kj ), and 2−kjλ/q ¯v0(2kj ) ≥ ε. By continuity of the integral, there are βkj ∈ (0, 2kj) such that Z 2kj βkj fkj (y)v(y)dy ≥ γ2 ¯v0(2kj ). 16 ELENA P. USHAKOVA Set Fkj = fkj χ(βkj ,2kj ). Then we have for ki and kj such that 2ki+1 < βkj : kLxFki − LxFkj kq 0 0 dx ds q q e−xyλ e−(y/s)λ q =ZR+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z x−1/λ s−λ−1(cid:12)(cid:12)(cid:12)(cid:12)Z s = λZR+ =: kfLxFki −fLxFkj kq = λZ 2ki +1 ≥ e−1λZ 2ki +1 [Fki(y) − Fkj (y)]v(y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) [Fki(y) − Fkj (y)]v(y)dy(cid:12)(cid:12)(cid:12)(cid:12) q ≥ kχ(2ki ,2ki+1)(fLxFki −fLxFkj )kq = kχ(2ki ,2ki+1)fLxFkikq s−λ−1 Z 2ki fki(y)v(y)dy!q fki(y)v(y)dy!q s−λ−1ds Z 2ki e−(y/s)λ βki βki q q ds 2ki 2ki ≥ γ2q 2λ − 1 2λe 2−λki[¯v0(2ki)]q ≥ γ2q 2λ − 1 2λe εq > 0, and Lx is not compact. Necessity of (3.11)(ii) can be established by the similar way obtaining a contradiction with the compactness of Lx. Another way to prove (3.11)(ii) for q > 1 is analogous to the proof of necessity (2.3)(ii) in the part (i) of this theorem. (cid:3) Remark 3.2. The condition ess supt∈R+ ¯Bq(t) < ∞ in (v) may be re- placed by ess supt∈R+ Bq(t) < ∞. 4. COMPACTNESS OF THE STIELTJES TRANSFORM We start from boundedness and compactness criteria for the Hardy f (y)φ(y)dy, 0 ≤ c1 ≤ x ≤ c2 ≤ ∞, operator Hhc1,c2if (x) := ψ(x)Z x Φc1(c2) :=Z c2 φp′ c1 c1 with weight functions φp′ ∈ Lloc(R+) and ψq ∈ Lloc(R+). Denote (y)dy, Ψc1(c2) :=Z c2 c1 ψq(x)dx. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 17 Theorem 4.1. (i) [6, 8, 14, 15, 20] Let 1 < p ≤ q < ∞. The operator Hhc1,c2i is bounded from Lphc1, c2i to Lqhc1, c2i if and only if Ahc1,c2i := sup c1<t<c2 Ahc1,c2i(t) := sup c1<t<c2(cid:2)Φc1(t)(cid:3)1/p′(cid:2)Ψt(c2)(cid:3)1/q < ∞, where kHhc1,c2ikLphc1,c2i→Lqhc1,c2i ≈ Ahc1,c2i. Besides, Hhc1,c2i is compact if and only if Ahc1,c2i < ∞ and lim t→c1 Ahc1,c2i(t) = lim t→c2 Ahc1,c2i(t) = 0. (ii) [4, 14] If 0 < q < 1 < p < ∞ or 1 < q < p < ∞ then Hhc1,c2i is bounded from Lphc1, c2i to Lqhc1, c2i if and only if Bhc1,c2i :=(cid:18)Z c2 c1 (cid:2)Φc1(t)(cid:3)r/p′(cid:2)Ψt(c2)(cid:3)r/p ψq(t)dt(cid:19)1/r < ∞, where kHhc1,c2ikLphc1,c2i→Lqhc1,c2i ≈ Bhc1,c2i. Moreover, finiteness of Bhc1,c2i gives a necessary and sufficient condition for the compactness of Hhc1,c2i if 1 < q < ∞. (iii) [23] Let 0 < q < 1 = p. The Hardy operator Hhc1,c2i is bounded from L1hc1, c2i to Lqhc1, c2i if and only if Bq<1,hc1,c2i < ∞, where Bq<1,hc1,c2i :=(cid:18)Z c2 c1 [ess sup c1<y<t φ(y)]q/(1−q)(cid:2)Ψt(c2)(cid:3)q/(1−q) ψq(t)dt(cid:19)(1−q)/q and kHhc1,c2ikL1hc1,c2i→Lqhc1,c2i ≈ Bq<1,hc1,c2i. (iv) [8] If p = 1 ≤ q < ∞ then Hhc1,c2i is bounded from L1hc1, c2i to Lqhc1, c2i if and only if B1≤q,hc1,c2i < ∞, where B1≤q,hc1,c2i := sup c1≤t≤c2(cid:2)ess sup c1<y<t φ(y)(cid:3)(cid:2)Ψt(c2)(cid:3)1/q and kHhc1,c2ikL1hc1,c2i→Lqhc1,c2i ≈ B1≤q,hc1,c2i. Some of the results collected in Theorem 4.1 are also referenced in [22] and [23]. Besides, similar statement is true for the operator H ∗ hc1,c2if (x) := ψ(x)Z c2 x f (y)φ(y)dy, 0 ≤ c1 ≤ x ≤ c2 ≤ ∞. 18 ELENA P. USHAKOVA Theorem 4.2. (i) If 1 < p ≤ q < ∞ then H ∗ Lqhc1, c2i if and only if hc1,c2i : Lphc1, c2i → A∗ hc1,c2i := sup c1<t<c2 A∗ hc1,c2i(t) := sup hc1,c2ikLphc1,c2i→Lqhc1,c2i ≈ A∗ with kH ∗ if and only if A∗ hc1,c2i < ∞ and c1<t<c2(cid:2)Φt(c2)(cid:3)1/p′(cid:2)Ψc1(t)(cid:3)1/q < ∞, hc1,c2i. Moreover, H ∗ hc1,c2i is compact lim t→c1 A∗ hc1,c2i(t) = lim t→c2 A∗ hc1,c2i(t) = 0. (ii) Hhc1,c2i is bounded from Lphc1, c2i to Lqhc1, c2i for 0 < q < 1 < p < ∞ or 1 < q < p < ∞ if and only if B∗ hc1,c2i :=(cid:18)Z c2 c1 (cid:2)Φt(c2)(cid:3)r/p′(cid:2)Ψc1(t)(cid:3)r/p ψq(t)dt(cid:19)1/r < ∞. hc1,c2ikLphc1,c2i→Lqhc1,c2i ≈ B∗ Besides, kH ∗ ∞ is necessary and sufficient for the compactness of H ∗ ∞. hc1,c2i and the condition B∗ hc1,c2i < hc1,c2i if 1 ≤ q < (iii) [23] Let 0 < q < 1 = p. The Hardy operator H ∗ from L1hc1, c2i to Lqhc1, c2i if and only if B∗ q<1,hc1,c2i < ∞, where hc1,c2i is bounded B∗ q<1,hc1,c2i :=(cid:18)Z c2 c1 [ess sup t<y<c2 φ(y)]q/(1−q)(cid:2)Ψc1(t)(cid:3)q/(1−q)ψq(t)dt(cid:19)(1−q)/q hc1,c2ikL1hc1,c2i→Lqhc1,c2i ≈ B∗ and kH ∗ (iv) [8] If p = 1 ≤ q < ∞ then H ∗ Lqhc1, c2i if and only if B∗ q<1,hc1,c2i. 1≤q,hc1,c2i < ∞, where hc1,c2i is bounded from L1hc1, c2i to B∗ 1≤q,hc1,c2i := sup and kH ∗ hc1,c2ikL1hc1,c2i→Lqhc1,c2i ≈ B∗ SH := sup t∈R+ Now put Λ :=(cid:18)RR+(cid:18)RR+ ¯v0(t)(cid:2)Wt(∞)(cid:3)1/q, SH,a(t) := ¯v0(t)(cid:2)Wt(a)(cid:3)1/q, SH,b(t) := ¯vb(t)(cid:2)Wt(∞)(cid:3)1/q, t<y<c2 c1≤t≤c2(cid:2)ess sup xλ+yλ(cid:19)p′ φ(y)(cid:3)(cid:2)Ψc1(t)(cid:3)1/q vp′(y)dy(cid:19)1/p′ 1≤q,hc1,c2i. w(x)dx , SH ∗ := sup t∈R+ ¯vt(∞)t−λ(cid:2)W0(t)(cid:3)1/q, SH ∗,a(t) := ¯vt(a)t−λ(cid:2)W0(t)(cid:3)1/q, SH ∗,b(t) := ¯vt(∞)t−λ(cid:2)Wb(t)(cid:3)1/q. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 19 Compactness criteria for the Stieltjes transformation S : Lp → Lq read Theorem 4.3. (i) If 1 < p ≤ q < ∞ then S : Lp → Lq is com- pact if and only if AH + AH ∗ < ∞ and limt→0[AH(t) + AH ∗(t)] = limt→∞[AH(t) + AH ∗(t)] = 0. (ii) Let 0 < q < p < ∞ and p > 1. If q 6= 1 then S is compact if and only if BH + BH ∗ < ∞. If q = 1 then S is Lp − L1 -- compact iff Λ < ∞. (iii) If 0 < q < 1 = p then S is Lp − Lq -- compact if and only if B1,H + B1,H ∗ < ∞. (iv) If p = 1 ≤ q < ∞ then the operator S : Lp → Lq is compact if and only if SH + SH ∗ < ∞ and lima→0 sup0<t<a[SH,a(t) + SH ∗,a(t)] = limb→∞ supb<t<∞[SH,b(t) + SH ∗,b(t)] = 0. Proof. (i) Let 1 < p ≤ q < ∞ and suppose AH + AH ∗ < ∞ (see (2.5) and (2.6)). Besides, assume (4.1) (i) lim t→0 [AH(t) + AH ∗(t)] = 0, (ii) lim t→∞ [AH(t) + AH ∗(t)] = 0. By theorems 4.1(i) and 4.2(i) these conditions guaranty the compact- ness of the operator H(f ) + H ∗ (f ) , which is a majorant for the transformation S (see the relation (2.4)). From here the compactness of S : Lp → Lq ensues by Theorem 2.13. The condition AH+AH ∗ < ∞ and (4.1) are also necessary for Lp−Lq -- compactness of S, when 1 < p ≤ q < ∞, by standard arguments for the Hardy operators H and H ∗. (ii), (iii): Let 0 < q < p < ∞ and p ≥ 1. If q 6= 1 we suppose BH + BH ∗ < ∞ for p > 1 and B1,H + B1,H ∗ < ∞ for the case p = 1. Compactness of S in the case p > 1 is guaranteed by BH + BH ∗ < ∞ (see theorems 2.12 and 2.13). If p = 1 and B1,H + B1,H ∗ < ∞ the compactness of S can be stated similarly to sufficiency of the conditions (iv) in Theorem 3.1. If q = 1 then S is compact by Lemma 2.15 provided Λ < ∞. 20 ELENA P. USHAKOVA Necessity of the condition BH + BH ∗ < ∞ in (ii) and B1,H + B1,H ∗ < ∞ in (iii) ensues from the compactness and, therefore, boundedness of the operator S. (iv) It remains to consider the case p = 1 ≤ q < ∞. Suppose SH + SH ∗ < ∞, (4.2) (i) lim a→0 (ii) lim b→∞ sup 0<t<a sup b<t<∞ [SH,a(t) + SH ∗,a(t)] = 0, [SH,b(t) + SH ∗,b(t)] = 0, and prove sufficiency of these assumptions for the L1 −Lq -- compactness of S. In view of (4.2) given ε > 0 there exist 0 < r < R < ∞ such that (4.3) (4.4) sup 0<t<r SH,r < ε/7, sup R<t<∞ SH,R < ε/7, sup 0<t<r sup R<t<∞ SH ∗,r < ε/7, SH ∗,R < ε/7. Now we divide S into a sum Sf = Sr,Rf + [Sr,if + SR,if ] 2Xi=1 of compact operators Sr,Rf := χ(r,R)S(f χ(r,R)), Sr,1f := χ[0,R)S(f χ([0,r]), SR,1f := χ[R,∞)S(f χ[0,R)), Sr,2f := χ[0,r]S(f χ(r,∞)), SR,2f := χ(r,∞)S(f χ[R,∞)). To confirm the compactness of these operators we shall use a combina- tion of Theorem 2.14 and [9, Corollary 5.1]. That is we need to show that for a given ε > 0 there exist δ > 0 and points 0 < s < t < ∞ such that for almost all y ∈ R+ and for every h > 0 with h < δ (4.5) (i) J q (ii) J q s (y) :=Z s t (y) :=Z ∞ 0 t kS(x, y)q dx < εq, kS(x, y)q dx < εq ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 21 and (4.6) J q h(y) :=ZR+ kS(x + h, y) − kS(x, y)q dx < εq, where kS(x, y) := w(x)kS(x, y)v(y). We start from the operator S := Sr,1 + Sr,R + SR,2. Suppose h < δ(ε) and write For simplicity consider the case λ = 1 and denote Jh,S(y) = v(y)(cid:18)ZR+ wq(x)(cid:20) (c1,c2)(y, h) :=Z r 0 I q 1 xλ + yλ − 1 (x + h)λ + yλ(cid:21)q dx(cid:19)1/q . wq(x)dx (x + y)q(x + y + h)q . We have Jh,S(y) = hχ[0,r)(y)v(y)I(0,r)(y, h) + hχ[0,r)(y)v(y)I(r,R)(y, h) +hχ(r,R)(y)v(y)I(r,R)(y, h) + hχ[R,∞)(y)v(y)I(r,R)(y, h) +hχ[R,∞)(y)v(y)I(R,∞)(y, h) =: Jh,i(y). 5Xi=1 The conditions (4.3) and (4.4) imply Jh,1(y) ≤ 2ε/7, and Jh,5(y) ≤ 2ε/7. To estimate Jh,2(y) note that Jh,2(y) ≤ hr−1¯v0(r)(cid:2)Wr(∞)(cid:3)1/q ≤ hr−1SH . From here, with δ = εr/7SH we obtain Jh,2 ≤ ε/7. Analogously, Jh,4 ≤ ε/7 if δ = εr/7SH ∗. For Jh,3 note that ¯vr(R)(cid:2)Wr(R)(cid:3)1/q < M < ∞ provided w ∈ Lq loc(R+) loc(R+). Therefore, Jh,3(y) ≤ hMr−2 and Jh,3(y) ≤ ε/7 if and v ∈ L∞ δ = εr2/7M. Summing up, we obtain Jh,S(y) < ε for almost all y ∈ R+, that is the condition (4.6) is satisfied. Fulfillment of the claims (4.5) ensues from (4.3) and (4.4) with s = r and t = R. Thus, the sum Sr,1 + Sr,R + SR,2 is compact. 22 ELENA P. USHAKOVA Compactness of the operator Sr,2 can be demonstrated as follows. The condition (4.5)(ii) is automatically fulfilled with t = r. To demon- strate (4.5)(i) note that kSr,2kL1→Lq ≈ ¯vr(∞)r−λ(cid:2)W0(r)(cid:3)1/q ≤ SH ∗ < ∞. Hence, given ε > 0 there exists 0 < s ≤ r such that Js,Sr,2(y) < ε. The condition (4.6) may be shown analogously with δ = εrλ/SH ∗. Sim- ilar arguments work for the operator SR,1. Necessity of the conditions [SH + SH ∗] < ∞ and (4.2) follow from [8, (cid:3) Lemma 1, Theorem 1] and the relation (2.4). Remark 4.4. In some special cases the compactness of S can be es- tablished through the Laplace transformation (1.2). Indeed, by the representation (4.7) S = L∗ wLv with Lv ≡ L and L∗ the compactness of S : Lp → Lq if the conditions of Theorem 3.1 are fulfilled for either Lv : Lp → Lr or Lw : Lq′ → Lr′ of the form wf (x) := w(y)RR+ e−xyλf (x)dx we are able to state Lwf (x) := RR+ e−xyλf (y)w(y)dy. Here the parameter r′ is such that r′ = r/(r − 1) for r > 1. In particular, if w ≡ v and p = q′ ≤ q = p′ then S : Lp → Lp′ is compact if A = sup t∈R+ A(t) := sup t∈R+ t−λ/2(cid:2)V0(t)(cid:3)1/p′ < ∞ and limt→0 A(t) = limt→∞ A(t) = 0. Necessity of these conditions can be proved by Theorem 2.9 and the inequality AS(t) ≥ A(t)2. 5. CASES p = ∞ AND q = ∞. In view of the representation (4.7) compactness criteria for the opera- tor L from p = ∞ and/or to q = ∞ may be useful to state compactness of the transformation S. To obtain the criteria we traditionally start from boundedness. Let L be the transform from L∞ to L∞. Then L is bounded if and only if (5.1) C1 := kvk1 < ∞. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 23 If L is acting from L∞ to Lq, 1 < q < ∞, we remind that by Lemma 2.6 Together with [14, §1.3.2, Th.1] this gives e−xyλ ZR+(cid:18)ZR+ kLkL∞→Lq ≈(cid:18)ZR+ f (y)v(y)dy(cid:19)q t−λ(cid:18)Z t kLkL∞→L1 =ZR+ 0 If q = 1 then L : L∞ → Lq boundedly then y−λ−1dy. 0 dx ≈ZR+(cid:18)Z y v(y)dy(cid:19)q−1 f (t)v(t)dt(cid:19)q v(t)dt(cid:19)1/q =: Cq>1. y−λv(y)dy =: Cq=1. Let q < 1. Then L is bounded from L∞ to Lq if Cq>1 < ∞. If Cq<1 :=ZR+ t−λ/qv(t)dt < ∞. Suppose L is an operator from Lp to L∞, when p ≥ 1. If p = 1 then kLkL1→L∞ = kvk∞ =: C∞. If p > 1 then L∗ is acting from L1 to Lp′, p′ > 1. This yields kLkLp→L∞ = kL∗kL1→Lp′ = kvkp′ =: Cp′. The estimate kLkLp→L∞ ≤ Cp′ here and in (5.1) ensues from Theorem 2.11. The reverse inequality follows by applying a function g(x) = χ[0,t−λ)(x)tλ into the L1 − Lp′ -- norm inequality for the operator L∗. Let Cq denote Cq>1 or Cq=1 depending on a range of the parameter q. Now we can state the required compactness criteria for L. Theorem 5.1. (i) Let 1 ≤ q < ∞. L is compact from L∞ to Lq iff Cq < ∞. (ii) Let q < 1. The operator L is L∞ − Lq -- compact if Cq>1 < ∞. If L is compact from L∞ to Lq then Cq<1 < ∞. (iii) If 1 < p ≤ ∞ then L is Lp − L∞ -- compact if and only if Cp′ < ∞. Proof. Proof of the results rests on [12, Theorem 5.2] and [1]. (cid:3) 24 ELENA P. USHAKOVA Remark 5.2. The operator L can not be compact from L1 to L∞ for any v. Since the Stieltjes transformation S is two-weighted then its com- pactness criteria for p = ∞ and/or q = ∞ can be derived from the results of Sec. 4. Acknowledgements. The European Commission is grant-giving authority for the research of the author (Project IEF-2009-252784). The work is also partially supported by Russian Foundation for Basic Research (Project No. 09-01-98516) and Far-Eastern Branch of the Russian Academy of Sciences (Project No. 09-II-CO-01-003). References [1] R.A. Adams, Sobolev Spaces, Academic Press, New York, 1975. [2] K.F. Andersen, Weighted inequalities for the Stieltjes transformation and Hilbert's double series, Proc. Roy. Soc. Edinburgh, Sect. A 86(1980), 75-84. [3] K.F. Andersen, H.P. Heinig, Weighted norm inequalities for certain integral operators, SIAM J. Math. Anal., 14(1983), 834-844. [4] T. Ando, On compactness of integral operators, Nederl. Akad. Wetensch. Proc. Ser. 65 = Indag. Math. 24(1962), 235-239. [5] S. Bloom, Hardy integral estimates for the Laplace transform, Proc. Amer. Math. Soc., 116(1992), 417-426. [6] J.S. Bradley, Hardy inequalities with mixed norms, Canad. Math. Bull., 21(1978), 405-408. [7] D.L. Kreider, R.G. Kuller, D.R. Ostberg, F.W. Perkins, An introduc- tion to linear analysis, Addison-Wesley Publishing Co., Inc., Reading, Mass.- Don Mills, Ont. 1966. [8] D. E. Edmunds, P. Gurka, L. Pick, Compactness of Hardy-type integral operators in weighted Banach function spaces, Studia Math., 109(1994), 73-90. [9] S.P. Eveson, Compactness criteria for integral operators in L∞ and L1 spaces, Proc. Amer. Math. Soc., 123(1995), 3709-3716. [10] A. Erd´elyi, The Stieltjes transformation on weighted Lp spaces, Appl. Anal., 7(1978), 213-219. [11] L.V. Kantorovich, G.P. Akilov, Functional Analysis in Normed Spaces, International Series of Monographs in Pure and Applied Mathematics, Vol. ON COMPACTNESS OF LAPLACE AND STIELTJES TRANSFORMATIONS 25 46. Oxford/London/Edinburgh/New York/Paris/Frankfurt, Pergamon Press, 1964. [12] M.A. Krasnosel'skii, P.P. Zabreiko, E.I. Pustyl'nik, P.E. Sobolevskii, Integral operators in spaces of summable functions, No- ordhoff International Publishing, Leyden, 1976. [13] E. Lomakina, V. Stepanov, On the Hardy-type integral operators in Banach function spaces, Publ. Mat., 42(1998), 165-194. [14] V.G. Maz'ya, Sobolev Spaces, Springer-Verlag, Berlin/Heidelberg, 1985. [15] B. Muckenhoupt, Hardy's inequality with weights, Studia Math., 44(1972), 31-38. [16] R. Oinarov, Boundedness and compactness of Volterra type integral opera- tors, Siber. Math. J., 48(2007), 884-896. [17] L.E. Persson, V.D. Stepanov, E.P. Ushakova, On integral operators with monotone kernels, Doklady Mathematics, 72(2005), 491-494. [18] D.V. Prokhorov, V.D. Stepanov, Weighted estimates for Riemann- Liouville operators and some of their applications, Proc. Steklov Inst. Math., 243(2003), 278-301. [19] H. Rafeiro, S. Samko, Dominated compactness theorem in Banach function spaces and its applications, Compl. Anal. Oper. Theory, 2(2008), 669-681. [20] S.D. Riemenschneider, Compactness of a class of Volterra operators, Tohoku Math. J., 26(1974), 385-387. [21] G. Sinnamon, A note on the Stieltjes transformation, Proc. Roy. Soc. Edin- burgh, Sect. A 110(1988), 73-78. [22] G.J. Sinnamon, Weighted Hardy and Opial-type inequalities, J. Math. Anal. Appl., 160(1991), 434-445. [23] G. Sinnamon, V.D. Stepanov, The weighted Hardy inequality: new proofs and the case p = 1, J. London Math. Soc., 54(1996), 89-101. [24] V.D. Stepanov, E.P. Ushakova, Weighted estimates for the integral opera- tors with monotone kernels on a half-axis, Sib. Math. J., 45(2004), 1124-1134. [25] E.P. Ushakova, On singular numbers of generalized Stieltjes transformation, Doklady Mathematics, 81(2010), 214-215. [26] E.P. Ushakova, Estimates for the singular values of transformations of Stielt- jes type, Siber. Math. J., (1)52(2011), 159-166. [27] A.H. Zemanian, Generalized integral transformations. Pure and Applied Mathematics, Vol. XVIII. Interscience Publishers [John Willey & Sons, Inc.], New York-London-Sydney, 1968. 26 ELENA P. USHAKOVA Current address: Department of Mathematics, University of York, York, YO10 5DD, UK. E-mail address: [email protected] Computing Centre of the Far-Eastern Branch of the Russian Acad- emy of Sciences, Khabarovsk, 680000, RUSSIA. E-mail address: [email protected]
1212.6672
2
1212
2013-01-02T20:55:49
A note on the hypercontractivity of the polynomial Bohnenblust--Hille inequality
[ "math.FA" ]
For $\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$ and $m$ a positive integer, we remark that there is a constant $C$ so that, for all $r\in\lbrack1,\frac {2m}{m+1}],$ the supremum of the ratio between the $\ell_{r}$ norm of the coefficients of any nonzero $m$-homogeneous polynomial $P:\ell_{\infty}% ^{n}(\mathbb{K}) \rightarrow\mathbb{K}$ and its supremum norm is dominated by $C^{m}\cdot n^{(\frac{m}{r}-\frac{m+1}{2})}$ and, moreover, we prove that the exponent $\frac{m}{r}-\frac{m+1}{2}$ is optimal.
math.FA
math
A NOTE ON THE HYPERCONTRACTIVITY OF THE POLYNOMIAL BOHNENBLUST -- HILLE INEQUALITY DANIEL PELLEGRINO Abstract. For K = R or C and m a positive integer, we remark that there is a constant C so that, for all r ∈ [1, 2m m+1 ], the supremum of the ratio between the ℓr norm of the coefficients of any nonzero m-homogeneous polynomial P : ℓn ∞ (K) → 2 ) and, moreover, we K and its supremum norm is dominated by Cm prove that the exponent m is optimal. · n( m m+1 m+1 − r − r 2 1. Introduction If P : ℓn ∞ (C) → C is an m-homogeneous polynomial defined by and Dn is the unit polydisc in Cn, let P (z) = Xα=m aαzα 1/r kPkr :=  Xα=m aαr  and kPk∞ := sup z∈Dn P (z) . The Sidon constant S1(m, n) (see [6, 12]) is the smallest constant satisfying the inequality (1.1) for all m-homogeneous polynomials P : ℓn references therein) we know that there is an absolute constant CC > 0 such that ∞ (C) → C. From [6, 12] (see also the kPk1 ≤ S1(m, n)kPk∞ (1.2) S1(m, n) ≤ C m C · n m−1 2 . It is also known that this result is sharp (we refer to [6, 12] and the references therein). The inequality (1.1) is related to the famous polynomial Bohnenblust -- Hille in- equality for complex scalars (see [2]). Since the polynomial Bohnenblust -- Hille in- equality is hypercontractive for both real and complex scalars (see [4, 6]), there exists a constant CK,BH > 1 such that kPk 2m m+1 ≤ C m K,BH kPk∞ (1.3) Key words and phrases. Sidon constants, homogeneous polynomials, Bohnenblust -- Hille inequality. 2010 Mathematics Subject Classification. 32A70, 32A50, 46G25, 47H60. 1 2 DANIEL PELLEGRINO for all positive integers n and all m-homogeneous polynomials P : ℓn K = R or C. ∞ (K) → K, with It is well-known that (1.2) is a corollary of (1.3). In fact, according to [5, (2.1)] there are constants C, CC > 0 such that (1.4) Xα=m aα ≤ 2m  Xα=m 1 ≤(cid:16)C m(cid:16)1 + ≤ C m C · n m−1  m(cid:17)m(cid:17) 2 kPk∞ . m−1 n m−1 2m   Xα=m m−1 2m C m C,BH kPk∞ m+1 2m aα 2m m+1  For related recent results we refer to [11, 13, 16] and [15] for a panorama of the subject. Both (1.2) and the polynomial (and multilinear) Bohnenblust -- Hille inequalities were originally conceived for complex scalars; the reason is that these inequalities were motivated by problems arising over the complex scalar field. In the last years, however, the interest in the Bohnenblust -- Hille inequality encompassed the case of real scalars, mainly due to its connections with Quantum Information Theory (see [8]). In this note we remark that (1.2) and (1.3) are particular cases of a continuum m+1 ], m, n be positive integers and K = R or C. There is family of sharp inequalities for both complex and real scalars: Theorem 1. Let r ∈ [1, 2m an universal constant LK > 0 such that kPkr ≤ Lm K · n( m for all m-homogeneous polynomials P : ℓn is optimal. 2 ) kPk∞ r − m+1 (1.5) ∞ (K) → K. Moreover, the power m r − m+1 2 It is worth mentioning that, in general, the adaptation of asymptotic results involving homogeneous polynomials from the complex setting to real scalars is not a straightforward task. In fact, sometimes polynomials present a completely different behavior when we change the scalar field from C to R (we refer to [3, page 58] for an illustrative example of this fact). 2. The proof Let r ∈ [1, 2m ing the argument used in (1.4). For real scalars, according to [4], if P : ℓn is an m-homogeneous polynomial, then m+1 ]. The proof of (1.5) for complex scalars is easily obtained by adapt- ∞ (R) → R kPCk∞ ≤ 2m−1 kPk∞ , where PC is the complexification of P ; this result goes back to the Visser's paper [17]. So, we obtain (1.4) for real scalars. It also simple to see that the constant LK can be chosen independent of m, r, n. Now let us prove the optimality of the ON THE HYPERCONTRACTIVITY OF THE BOHNENBLUST -- HILLE INEQUALITY 3 r − m+1 exponent m r − m+1 q < m 2 . For each m, n, let 2 ; for this task let us suppose that the result holds for a power Pm,n : ℓn ∞ (K) → K εαwα Pm,n(w) = Xα=m be the m-homogeneous Bernoulli polynomial satisfying the Kahane -- Salem -- Zygmund inequality (note that this inequality is also valid for real scalars, see [10]). The proof follows the lines of [10, Theorem 10.2]; the essence of this argument can be traced back to Boas' classical paper [1]. We can suppose n > m. As in [10], we have Pα=m εαr = p(n) + 1 m! m−1 Qk=0 (n − k), where p (n) > 0 is a polynomial of degree m − 1. If (1.5) was valid with the power q, then there would exist a constant Cq,K > 0 so that 1 r   Xα=m εαr  ≤ C m ≤ C m q,K · nq kPm,nk∞ q,K · nq · CKSZ · n(m+1)/2plog m, where CKSZ > 0 is the universal constant from the Kahane -- Salem -- Zygmund in- equality. Hence 1 nq · n(m+1)/2√log m(cid:18)p(n) + C m q,KCKSZ ≥ for all n. Raising both sides to the power of r and letting n → ∞ we obtain n→∞ q,KCKSZ(cid:1)r ≥ lim (cid:0)C m nqr · nr(m+1)/2(cid:0)√log m(cid:1)r + Qk=0 nqr · nr(m+1)/2(cid:0)√log m(cid:1)r! , (n − k)(cid:19)1/r 1 m! with p(n) s(n) m−1 s(n) = 1 m! m−1 Qk=0 (n − k). Since q < m is infinity, a contradiction. r − m+1 2 , we have deg s = m > qr + r(m + 1)/2 and thus the limit above References [1] H.P. Boas, The football player and the infinite series, Notices of the American Mathematical Society 44 (1997), 1430 -- 1435. [2] H.F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Annals of Mathematics 32 (1931), 600 -- 622. [3] G. Botelho, H.-A. Braunss, H. Junek and D. Pellegrino, Holomorphy types and ideals of mul- tilinear mappings, Studia Mathematica 177 (2006), 43 -- 65. 4 DANIEL PELLEGRINO [4] J.R. Campos, G.A. Munoz-Fern´andez, D. Pellegrino and J.B. Seoane-Sep´ulveda, The Bohnenblust-Hille inequality for real homogeneous polynomials is hypercontractive and this result is optimal, arXiv:1209.4632. [5] A. Defant and L. Frerick, Hypercontractivity of the Bohnenblust-Hille inequality for polyno- mials and multidimensional Bohr radii, arXiv:0903.3395. [6] A. Defant, L. Frerick, J. Ortega-Cerd´a, M. Ounaıes and K. Seip, The polynomial Bohnenblust -- Hille inequality is hypercontractive, Annals of Mathematics 174 (2011), 485 -- 497. [7] D. Diniz, G.A. Munoz-Fern´andez, D. Pellegrino and J.B. Seoane-Sep´ulveda, The asymptotic growth of the constants in the Bohnenblust -- Hille inequality is optimal, Journal of Functional Analysis 263 (2012), 415 -- 428. [8] A. Montanaro, Some applications of hypercontractive inequalities in quantum information theory, Journal of Mathematical Physics 53, 122206 (2012). [9] D. Nunez-Alarc´on, D. Pellegrino and J.B. Seoane-Sep´ulveda, On the Bohnenblust-Hille inequal- ity and a variant to Littlewood's 4/3 inequality, Journal of Functional Analysis 264 (2013), 326 -- 336. [10] D. Nunez-Alarc´on, D. Pellegrino, J.B. Seoane-Sep´ulveda and D. Serrano-Rodr´ıguez, There exist n=1 with limn→∞ (Cn+1 − Cn) = 0, Journal of multilinear Bohnenblust -- Hille constants (Cn)∞ Functional Analysis 264 (2013), 429 -- 463. [11] D. Nunez-Alarc´on, A note on the polynomial Bohnenblust -- Hille inequality, arXiv:1208.6238. [12] J. Ortega-Cerd´a, M. Ounaıes and K. Seip, The Sidon constant for homogeneous polynomials, arXiv:0903.1455. [13] J. Ortega-Cerd´a and K. Seip, A lower bound in Nehari's theorem on the polydisc, Journal d'Analyse Mathematique 118 (2012), 339 -- 342. [14] K. Seip, The Bohnenblust -- Hille inequality for homogeneous polynomials, http://www.matdat.life.ku.dk/henrikp/fnu2011/Seip-copenhagen.pdf [15] K. Seip, Estimates for Dirichlet polynomials, EMS Lecturer, CRM, 20 -- 23 February 2012, http://www.euro-math-soc.eu/system/files/Seip CRM.pdf [16] D. Serrano-Rodr´ıguez, Improving the closed formula for subpolynomial constants in the mul- tilinear Bohnenblust-Hille inequality, Linear Algebra and its Applications, in press. [17] C. Visser, A generalization of Tchebychef's inequality to polynomials in more than one variable, Indagationes Mathematicae 8 (1946), 310 -- 311. Departamento de Matem´atica, UFPB, Joao Pessoa, PB, Brazil E-mail address: [email protected] and [email protected]
1709.08836
4
1709
2019-05-20T17:39:59
Conjugate Phase Retrieval on ${\mathbb C}^M$ by real vectors
[ "math.FA", "cs.IT", "cs.IT" ]
In this paper, we will introduce the notion of {\it conjugate phase retrieval}, which is a relaxed definition of phase retrieval allowing recovery of signals up to conjugacy as well as a global phase factor. It is known that frames of real vectors are never phase retrievable on ${\mathbb C}^M$ in the ordinary sense, but we show that they can be conjugate phase retrievable in complex vector spaces. We continue to develop the theory on conjugate phase retrievable real frames. In particular, a complete characterization of conjugate phase retrievable real frames on ${\mathbb C}^2$ and ${\mathbb C}^3$ is given. Furthermore, we show that a generic real frame with at least $4M - 6$ measurements is conjugate phase retrievable in ${\mathbb C}^M$ for $ M \ge 4.$
math.FA
math
CONJUGATE PHASE RETRIEVAL ON CM BY REAL VECTORS LUKE EVANS AND CHUN-KIT LAI Abstract. In this paper, we will introduce the notion of conjugate phase retrieval, which is a relaxed definition of phase retrieval allowing recovery of signals up to conjugacy as well as a global phase factor. It is known that frames of real vectors are never phase retrievable on CM in the ordinary sense, but we show that they can be conjugate phase retrievable in complex vector spaces. We continue to develop the theory on conjugate phase retrievable real frames. In particular, a complete characterization of conjugate phase retrievable real frames on C2 and C3 is given. Furthermore, we show that a generic real frame with at least 4M − 6 measurements is conjugate phase retrievable in CM for M ≥ 4. 1. Introduction The phase retrieval problem concerns reconstruction of a signal from linear measure- ments with noisy or corrupt phase information. The classical formulation comes from applications such as X-ray crystallography where a signal must be recovered from the magnitudes of its Fourier coefficients [14]. Phase retrieval also occurs in numerous other applications such as diffraction imaging [8, 9], optics [14, 13], speech processing [6], deep learning [23, 17], and quantum information theory [15, 16]. In 2006, Balan, Casazza and Edidin introduced the following mathematical formulation for the phase retrieval problem within a complex Hilbert space H [6]: Definition 1.1. Let H be a Hilbert space over the field C. We say that a set of vectors {ϕn}n∈I ⊆ H with index set I ⊆ N, is complex phase retrievable if (1.1) If H is over the real numbers, then we say that {ϕn}n∈I is real phase retrievable if x = eiθy is replaced by x = ±y in (1.1). hx, ϕni = hy, ϕni for all n ∈ I =⇒ x = eiθy, for some constant θ. One of the main questions in phase retrieval on complex vector spaces is to determine the minimal number N for which a generic frame (See Definition 2.1) in CM with N vectors can achieve complex phase retrieval. This means also that with probability one, a ran- domly chosen frame with at least N vectors can perform conjugate phase retrieval. Balan, Casazza and Edidin introduced the complement property as a geometric characterization of real phase retrievability and showed that for H = RM any generic frame with at least 2M − 1 vectors is real phase retrievable [6]. In comparison, complex phase retrievability is a much more difficult problem. There is no known geometric characterization of complex phase retrievability, and for H = CM we know that 4M − 4 vectors are sufficient, with the necessary number of vectors of the order 4M − o(1) [7, 11]. There has also been intensive research about the stability of phase retrieval and other different type of generalizations. 2010 Mathematics Subject Classification. Primary 42C15, Secondary 15A63. Key words and phrases. Phase retrieval, Frame, Conjugate, Generic numbers. 1 Interested readers may refer to [5] for a more detailed discussion and summary of the recent results in phase retrieval. 1.1. Conjugate Phase retrieval. Despite the wide applicability of complex phase re- trieval, satisfactory descriptions for complex phase retrievable frames are still lacking. In particular, a set of vectors {ϕn : n ∈ I} taken from RM can never be complex phase retrievable on CM , regardless of how many vectors we take. Real frames fail because real measurement vectors completely ignore conjugation: if ϕn ∈ RM , then hx, ϕni = hx, ϕni, for all x ∈ CM . However, x 6= eiθx in general (for example, take x = (1 i i ··· i)T ∈ CM .) This introduces also an additional difficulty to geometrically visualize complex phase retrievable vectors, which all lie inside CM \ RM . 2 , 1 Phase retrieval problem is also defined on the Paley-Wiener space P W consisting of all of entire functions band-limited to [ 1 2 ]. We say that a sequence {λn}n∈Z ⊆ R is a set of real unsigned sampling if for any two real-valued f, g ∈ P W , f (λn) = g(λn) for all n ∈ Z implies that f = ±g. It is proved that [20] (see also [1, 3]) if λn are taken to be twice of the Nyquist rate (e.g. 1 Z), then it forms a set of real unsigned sampling. 2 However, the natural extension of our definition of unsigned sampling sets to complex valued P W cannot be resolved so easily. Given any band-limited complex-valued function f , the function g(x) = f (x) is also a function in P W . Clearly f (λ) = g(λ) for any λ ∈ R, but it is not true in general that f (x) = eiθf (x) for all x ∈ R and global constant θ. Thus, there cannot exist a sequence of reals {λn}k∈N that is a set of complex unsigned sampling as defined. Both cases we discussed share the same problem that real samples and measurements cannot distinguish conjugate vectors or functions. Yet, with the phase information avail- able, real frames on CM can span the complex vector spaces and real samples Z can per- fectly reconstruct bandlimited functions by the well-known Shannon Sampling Theorem. This means that if we want to close the gap between classical and phaseless reconstruction using real measurements, we need to accept conjugacy as one of our ambiguities. We thus propose the following definition: Definition 1.2. We say that a set of vectors {ϕn}n∈I ⊆ CM with index set I ⊆ N, is conjugate phase retrievable if hx, ϕni = hy, ϕni for all n ∈ I =⇒ there exists θ such that x = eiθy or x = eiθy. (Here y means taking the conjugate over each coordinates) It is clear that frames that are complex phase retrievable must be conjugate phase retrievable. Recently, the concept of norm retrieval with the implication requiring only kfk = kgk was proposed in [4] as another relaxed version of phase retrieval. The following implication is obvious: Complex phase retrieval =⇒ Conjugate phase retrieval =⇒ Norm retrieval. From this implication, we believe that conjugate phase retrieval would not lose more generality in the reconstruction as norm retrieval does. We now discuss our main result of conjugate phase retrieval. 2 1.2. Contribution. We will be focusing mainly on finite dimensional vector space CM . The main conclusion of this paper is that frames of real vectors can be conjugate phase retrievable, for example, the frame with vectors in the column of Φ =(cid:20)1 0 1 0 1 1(cid:21) is con- jugate phase retrievable when considered over C2 (see Theorem 2.4). We will explore in detail the conjugate retrievability of real frames lying in CM . On C2 and C3, we fully solve the number of real measurement vectors needed for conjugate phase retrievability, and characterize all conjugate phase retrievable frame as real algebraic varieties. Building from the recent results by Wang and Xu [22], we prove that 4M − 6 is a sufficient number of generic measurements for conjugate phase retrieval in CM for M ≥ 4. The main idea of the proofs will be considering the phase-lift maps (similar to [7, 5]) by identifying a vector x as xx∗ in the space of all Hermitian matrices. We will show that x and y are equivalent up to a phase and conjugacy if and only if the real part of xx∗ and yy∗ are equal (see Theorem 2.1), on which our analysis will be based. We will also explore the conjugate phase retrievable frames that cannot perform complex In phase retrieval. Such frames are called strictly conjugate phase retrievable. particular, real frames belong to this class. On C2, we will show that the only strictly conjugate phase retrievable frames are essentially real frames. We will organize our article as follows: In section 2, we will present our main setup and state our main results rigorously. In section 3, we will review the complement property and study the phase-lift map for conjugate phase retrieval. In section 4, we fully characterize conjugate phase retrieval by real frames on C2 and C3. We prove the generic number of 4M−6 for CM , M ≥ 4 in section 5. For section 6, we will study the strictly conjugate phase retrievable frames. We will end our article with some open questions for conjugate phase retrieval in both finite dimensional and infinite-dimensional Hilbert spaces in Section 7. 2. Setup and Main Results Throughout the rest of the paper, we will use the following equivalence relation on CM : For x, y ∈ CM , x ∼ y if and only if x = eiθy for some θ ∈ [0, 2π) conj x ∼ y if and only if x ∼ y or x ∼ y. Recall also that if y = (y1 ··· yM )T , then y = (y1 ··· yM )T . It is direct to check that the above statements are equivalence relations. A set of vectors Φ = {ϕn : n = 1, . . . , N} is called a frame for CM if there exists 0 < A ≤ B < ∞ such that Akxk2 ≤ hx, ϕni2 ≤ Bkxk2, for all x ∈ CM . NXn=1 Here, ϕn may be taken from RM or CM . No matter where the ϕn are taken, a frame Φ for CM must be a spanning set of CM . The ratio of the frame bounds, B/A, control the robustness of the reconstruction. However, we will not be discussing the stability problem, so we will identify our frame Φ as a full-rank M × N (short-fat) matrix Φ with entries taken over R or C. i.e. Φ = ϕ1 ϕ2 3 ··· ··· ϕN ···  . If all ϕn ∈ RM , we will identify Φ as an element in RM×N . Otherwise, Φ is identified as an element in CM×N . On RM×N we endow it with the standard Euclidean topology. On CM×N = R2M×2N we endow it with the Euclidean topology by considering the real and imaginary parts of each complex entry as separate coordinates. Putting also the standard Lebesgue measure on RM×N and R2M×2N , we have the following definition: Definition 2.1. Let F = R or C and let Y ⊆ FM×N be the set of full-rank M × N matrices over F with some specified property P. If Y is open and dense in FM×N and X := Y c has Lebesgue measure 0, we say that each frame Φ ∈ Y is called a generic frame with property P. For most of the frame theory literature, X := Y c is an real algebraic variety, which means X can be represented as a common zero set of a finite number of polynomial equations. It is well-known that for an algebraic variety, Y is either empty or an open- dense set with full Lebesgue measure. Thus, a frame in Y will be a generic frame. Our goal is not only to show that it is possible for real frames to perform conjugate phase retrieval, but to also determine as much as possible, N∗(M ) := min{N : a generic frame Φ ⊂ RM×N is conjugate phase retrievable on CM} N∗(M ) := min{N : there exists Φ ⊂ RM×N which is conjugate phase retrievable on CM}. The main idea of theory will be to develop the phase-lift setup for the conjugate phase retrieval. Phase-lift has been the central idea for complex phase retrieval [7, and references therein], which linearizes the absolute value of the inner product. C 2.1. Notation. Let HM×M be the set of all complex M ×M Hermitian matrices (H = H∗ with ∗ denotes the conjugate transpose) and let HM×M be the set of all real M × M symmetric matrices (H = H T ). Both sets form vector spaces over the real numbers. Given H ∈ HM×M , we define C R Re(H) = [Re(hij)] ∈ HM×M R , C (respectively S r where Re(z) denote the real part of the complex number z. We will use similar notation as in [5] for spaces of Hermitian/symmetric matrices of lower rank. For 1 ≤ r ≤ M , we (respectively HM×M define also the set S r ) whose rank is at most r. For non-negative integers p, q such that p + q ≤ M , we define also S p,q F = {Q ∈ HM×M where F = C or R. Of particular interest is the subclass S 1,0 have the following representation: : Q has at most p positive eigenvalues and at most q negative eigenvalues} R) to be the set of Q ∈ HM×M C , which is known to C and S 1,1 F C R S 1,0 C = {xx∗ : x ∈ CM} S 1,1 C = S 1,0 C − S 1,0 C (See [5, Lemma 3.7]). 4 2.2. Results on Phase-lift. Below we provide necessary and sufficient conditions for conjugate phase retrievability in terms of the corresponding phase-lift map, with proofs following in Section 3.2. First, we have the following important characterization about (conjugate) equivalent vectors. Theorem 2.1. For any x, y ∈ CM , (1) x ∼ y if and only if xx∗ = yy∗. (2) x conj ∼ y if and only if Re(xx∗) =Re(yy∗). Given a finite set of frame vectors Φ := {ϕn : n = 1, . . . , N} ⊂ RM , we also note that hx, ϕni 2 = ϕ∗n(xx∗)ϕn. Therefore, the following phase-lift map will play an important role: −→ RN , A(Q) := (ϕT Define similarly the phase-lift map AC on HM×M A : HM×M R with transpose replaced by conjugate transpose. It was proved in ([15, Proposition 2], see also [7, Lemma 1.9] and [5, Theorem 2.2]) that C 1 Qϕ1, . . . , ϕT N QϕN )T . Lemma 2.2. A complex frame Φ is complex phase retrievable if and only if ker(AC) ∩ S 1,1 C = {O}. The lemma can be obtained using the linearity of AC and Theorem 2.1 (1) with the fact C can be written as xx∗ − yy∗ C , which means all matrices from S 1,1 C − S 1,0 C = S 1,0 that S 1,1 for some x, y ∈ CM . The following provides the analogous theorem for conjugate phase retrieval. Theorem 2.3. Let Φ := {ϕn : n = 1, . . . , N} ⊂ RM be a finite set of frame vectors. (1) Φ is conjugate phase retrievable if and only if ker(A) ∩ Re(S 1,1 where Re(S 1,1 (2) If ker(A) ∩ S 4 The proof of (2) is obtained by proving Re(S 1,1 C ) = {Re(Q) : Q ∈ S 1,1 C }. R = {O}, then Φ is conjugate phase retrievable. C ) are contained inside S 4 follows from (1). However, we believe that the containment should be strict. C ) = {O}, R and thus (2) 2.3. Results on conjugate phase retrieval on CM . The phase-lift setup gives us the complete solution to M = 2 and 3, provided below and proven in Section 4. Theorem 2.4. N∗(2) = N∗(2) = 3 and N∗(3) = N∗(3) = 6. Moreover, c1 c2(cid:21) in R2×3 is conjugate phase retrievable on C2 if and det 1 2a1a2 a2 a2 2 b2 b2 2 1 c2 c2 1 2  6= 0. 2b1b2 2c1c2 5 (1) If M = 2, Φ =(cid:20)a1 a2 only if b1 b2 (2.1) b1 b2 b3 c1 d1 c2 d2 c3 d3 e1 f1 e2 f2 e3 f3  in R3×6 is conjugate phase retrievable a2 1 a2 2 a2 3 2a1a2 2a1a3 2a2a3 b2 b2 b2 2b2b3 2b1b2 1 2 3 c2 c2 c2 2c1c2 2c2c3 3 2 1 d2 d2 d2 2d1d2 2d1d3 2d2d3 1 2 3 e2 e2 e2 2e1e2 2e2e3 3 2 1 f 2 f 2 f 2 2f2f3 2f1f2 3 2 1 2e1e3 2f1f3 2b1b3 2c1c3  6= 0. a1 a2 a3 on C3 if and only if (2) If M = 3, Φ =  det (2.2) It is easy to find vectors for which the above determinants are non-zero, so the zero set of the determinant is non-empty and forms an algebraic variety. Thus (2) and (3) implies that N∗(2) = 3 and N∗(3) = 6. Moreover, using the recent result of Wang and Xu [22], we show that for M ≥ 4, Theorem 2.5. Let M ≥ 4. Suppose that N ≥ 4M − 6. Then a generic frame Φ = {ϕi : i = 1, . . . , N} ⊂ RM is conjugate phase retrievable. Theorem 2.1 and Theorem 2.3 will be proved in Section 3.2. Theorem 2.4 will be proved in Section 4, and Theorem 2.5 will be proved in Section 5. Finally, we will discuss strict conjugate phase retrievability in Section 6. 3. Complement property and Phase-lift 3.1. The Complement Property. A set of vectors {ϕn}N H is said to have the complement property if for any subset I in {1,. . . ,N}, n=1 in a complex Hilbert space span{ϕn : n ∈ I} = H or span{ϕn : n ∈ I c} = H. The complement property is known to be the fundamental property for phase retrieval [6][7]. We first derive complement property as a necessary condition for conjugate phase retrieval by real-valued vectors. We say that a vector ϕ ∈ CM is real-valued if all entries are real numbers. The relationship between the real span of a real frame in Rm and the complex span of the same frame in CM is crucial for the proof of complement property. Lemma 3.1. A collection of real-valued vectors {ϕn}N n=1 = CM if and only if span n=1 = Rm. n=1 in CM has R {ϕn}N C {ϕn}N span Proof. Let {ϕn}N C {ϕn}N span n=1 be a collection of real-valued vectors in CM . We write n=1 =( NXn=1 znϕn zn ∈ C) =( NXn=1 (an + ibn)ϕn an, bn ∈ R) . After distributing (an + ibn)ϕn we receive C {ϕn}N span n=1 = span R {ϕn}N n=1 ⊕ span R {iϕn}N n=1. Suppose that {ϕn}N that spanR{ϕn}N iRM and we can say that spanC{ϕn}N n=1 spans CM . Since CM is the direct sum of Rm and iRM , we conclude n=1 = (cid:3) n=1 = RM . Conversely, if spanR{ϕn}N n=1 = RM , then spanR{iϕn}N n=1 = CM . 6 Proposition 3.2. Every conjugate phase retrievable frame in CM consisting of all real- valued vectors has the complement property in CM . n=1 be a frame of real vectors in CM allowing conjugate phase Proof. Let Φ = {ϕn}N retrieval. For any x, y ∈ RM , hx, ϕni2 = hy, ϕni2 for all n = 1,··· N implies x ∼ y or x ∼ y. Since y is real we conclude that x ∼ y and x = ±y. Hence, Φ is real phase retrievable on RM and must have the complement property in RM . Since the complement property is defined by spanning properties, Lemma 3.1 implies that Φ must have the complement property in CM . (cid:3) We currently do not know whether a conjugate phase retrievable complex frame must possess the complement property. We will discuss more on complex conjugate phase retrievable frames in section 6. 3.2. Outer Products and Conjugate Equivalence. We will now prove our Theorem 2.1 and Theorem 2.3, which will form our foundation of the paper. Throughout the proof, we will denote by [M ] the set {1, . . . , M} and T the circle group. Proof of Theorem 2.1 (1). This part of the theorem should be well-known, and we provide it here for completeness. Suppose that xx∗ = yy∗. Then, by comparing entries, we have xixj = yiyj for all i, j ∈ [M ]. In particular, if i = j, then xi2 = yi2 for all i ∈ [M ]. This shows that xk = λkyk for some λk ∈ T. Thus, given i, j ∈ [M ], xixj = λiλjyiyj and hence λiλjyiyj = yiyj. If yi, yj 6= 0, it follows that λiλj = 1 and that λi = λj. Thus for any indices i, j with yi, yj 6= 0 we have xi = λyi and xj = λyj for some λ ∈ T. For any index k with yk = 0 we have that xk = 0 and trivially that xk = λyk. Therefore, xk = λyk for all k ∈ [M ] where λ ∈ T, implying that x ∼ y. The converse holds by a direct computation. (cid:3) Proof of Theorem 2.1 (2). To prove part (2) of Theorem 2.1, we first note that x and only if conj ∼ y if xx∗ = yy∗ (x ∼ y) or xx∗ = yy∗ (x ∼ y). (3.1) If x, y ∈ CM with xx∗ = yy∗ we trivially have Re(xx∗) = Re(yy∗). Likewise, xx∗ = yy∗ implies Re(xx∗) + i Im(xx∗) = Re(yy∗) + i Im(yy∗). But note that yiyj = yiyj, which implies that Re(yy∗) = Re(yy∗). We therefore conclude that Re(xx∗) = Re(yy∗). We now prove the converse. Suppose that Re(xx∗) = Re(yy∗). Then we have Re(xixj) = Re(yiyj) for all i, j ∈ [M ]. Using (3.1), we must show that if we write x = (x1 ··· xM )T , y = (y1 ··· yM )T ∈ CM , we have (3.2) (xixj = yiyj for all i, j ∈ [M ]) or (xixj = yiyj for all i, j ∈ [M ]) . We first claim the following weaker statement: Claim 1: Given any i, j ∈ [M ], xixj = yiyj or xixj = yiyj holds. To see this, we first note that by putting i = j in the assumption. (3.3) xi2 = Re(xixi) = Re(yiyi) = yi2 7 Thus, xixj2 = yiyj2. With Re(xixj) = Re(yiyj), we find that Re(xixj)2 + Im(xixj)2 = Re(yiyj)2 + Im(yiyj)2 Im(xixj) = ± Im(yiyj) given any i, j ∈ [M ]. Hence, we have Re(xixj) = Re(yiyj) and Im(xixj) = ± Im(yiyj). Therefore, xixj = yiyj or xixj = yiyj = yiyj and the claim is justified. We now prove by induction on M that (3.2) holds. We first notice that we only need to check (3.2) for i 6= j (since i = j follows from (3.3)). If M = 2, we have only one pair of (i, j), namely (i, j) = (1, 2). Therefore, the statement is true trivially. When M = 3, we may assume without loss of generality that none of the xi are zero. Otherwise, there is only one pair and the equations for other pairs holds trivially as they are all zero. Now, we have three pairs for (i, j) = (1, 2), (2, 3), (1, 3). Using Claim 1 and the pigeonhole principle, one of the two possibilities in Claim 1 must hold twice. Without loss of generality, assume we have Multiplying them together gives x1x2 = y1y2 and x2x3 = y2y3 x1x22x3 = y1y22y3. By (3.3) and x2 6= 0, we can cancel out the moduli of x2 and y2 and conclude that x1x3 = y1y3. Hence, xx∗ = yy∗. If the other possibility holds twice, using the same argument, we will have xx∗ = yy∗. For M ≥ 4, we use induction. Suppose that the claim holds for dimension M − 1. Let x = (x1 ··· xM )T and y = (y1 ··· yM )T be vectors in CM where Re(xixj) = Re(yiyj) for each i, j ∈ [M ]. Consider the vectors (x1 ··· xM−1)T and (y1 ··· yM−1)T in CM−1. Suppose that x1x2 = y1y2 holds. Then by the inductive hypothesis, xixj = yiyj for all i, j ∈ [M − 1]. Similarly, considering the vectors (x2 ··· xM )T and (y2 ··· yM )T in CM−1, we conclude by the induction hypothesis that xixj = yiyj for all i, j ∈ {2, . . . , M}. Hence, combining the conditions on (x1 ··· xM−1)T and (x2 ··· xM )T we have xixj = yiyj for all i, j ∈ [M ], except i = 1 and j = M . We now prove that x1xM = y1yM . Note that if all x2, ..., xM−1 are zero, we essentially have only one choice (i, j) = (1, M ) and the equations for other pairs holds trivially as they are all zero. Therefore, (3.2) holds trivially. Without loss of generality, we assume that x2 6= 0. Then multiply the equation for the pair (1, 2) and (2, M ) and argue in the same way as before in M = 3, we conclude that x1xM = y1yM also holds. Equivalently we have that xx∗ = yy∗. Similarly, if we assume instead that x1x2 = y1y2, we conclude that xixj = yiyj for all i, j ∈ [M ], in other words that xx∗ = yy∗. This completes the proof of (3.2) and hence the whole proof of Theorem 2.1. (cid:3) For Q ∈ HM×M F with F = C or R, we define the vectorization of Q by (3.4) v(Q) = (q11 q22 ··· qM M q12 ··· q1M ··· q(M−1)M )T ∈ C M (M +1) 2 which is a vector with M coordinates from the diagonal of Q and subsequent coordinates given from the remaining row entries above the diagonal of Q given from row 1 to row M . 8 For ϕ ∈ RM , we define ωϕ ∈ R M (M +1) 2 by ωϕ = (ϕ2 1 ··· ϕ2 M 2ϕ1ϕ2 ··· 2ϕ1ϕM ··· 2ϕM−1ϕM )T . The following lemma expresses two different important identities for the magnitudes of frame coefficients: Lemma 3.3. Let ϕ ∈ RM and let x ∈ CM . Then (3.5) = (cid:10)ωT hx, ϕi 2 = ϕT (Re(xx∗))ϕ ϕ , v(Re(xx∗))(cid:11) . Proof. We note that (3.6) hx, ϕi 2 = ϕT (xx∗)ϕ = ϕT (Re(xx∗))ϕ + iϕT (Im(xx∗))ϕ. But the left hand side is real-valued and ϕT is real-valued, which proves the first equality. We have (3.5) proved. To prove (3.6), we let Q = Re(xx∗) = [qij]1≤i,j≤M , then ϕT Qϕ = MXi=1 q2 iiϕ2 ii +Xi<j 2qijϕiϕj = hωϕ, v(Q)i . (cid:3) We now turn to prove Theorem 2.3. The following lemma concerns the rank of the real part of rank 1 complex matrices. Lemma 3.4. For any x, y ∈ CM , rank(Re(xx∗)) ≤ 2, and rank(Re(xx∗ − yy∗)) ≤ 4. Proof. Notice that Re(xx∗) = xx∗+xx∗ for any A, B ∈ CM×N we can say that 2 = xx∗ 2 + xx∗ 2 . Since rank(A+B) ≤ rank(A)+rank(B) rank(Re(xx∗)) ≤ rank(cid:18) xx∗ 2 (cid:19) + rank(cid:18) xx∗ 2 (cid:19) ≤ 2. Similarly, rank(Re(xx∗ − yy∗)) ≤ rank(Re(xx∗)) + rank(− Re(yy∗)) ≤ 2 + 2 = 4. (cid:3) Proof of Theorem 2.3. We first prove (1). Suppose that Φ is conjugate phase retrievable. C , we can say that there exist x, y ∈ CM Take Q ∈ ker(A)∩Re(S 1,1 such that Q = Re(xx∗ − yy∗). Now, for all n = 1, . . . , N , by Lemma 3.4 (1), (3.7) C −S 1,0 n Re(yy∗)ϕn = hx, ϕni 2 − hy, ϕni2. n Qϕn = ϕT C ). Since S 1,1 C = S 1,0 n Re(xx∗)ϕn − ϕT 0 = ϕT Thus, by conjugate phase retrievability of Φ, we have x Re(xx∗) = Re(yy∗), which shows that Q = O, the zero matrix. ∼ y. By Theorem 2.1 (2), conj Conversely, suppose that hx, ϕni 2 = hy, ϕni 2 for all n = 1, . . . , N . Then, with the same computation in (3.7) we have Q = Re(xx∗ − yy∗) ∈ ker(A) and also Q ∈ Re(S 1,1 C ). By our assumption, Q = O. Thus Re(xx∗) = Re(yy∗), which means x ∼ y by Theorem 2.1 (2). conj 9 For (2), we just notice that from Lemma 3.4 (2), any Q ∈ Re(S 1,1 most 4. Thus, Re(S 1,1 and (2) then follows from (1). C ) is a subset of S 4 R. If ker(A)∩S 4 R = {0}, then ker(A)∩Re(S 1,1 C ) must have rank at C ) = {0} (cid:3) 4. Conjugate Phase Retrieval on C2 and C3 In this section, we will give a complete study of conjugate phase retrieval by real frames n=1 in CM we define the N × M (M +1) 2 on C2 and C3. Given a real valued frame Φ = {ϕn}N matrix − ωT ϕ1 − ... − ωT ϕN − ΩΦ =  where the n-th row is the vector ωT ϕn. Notice that if M = 2 or 3 and N = M (M + 1)/2, then the respective ΩΦ are exactly the matrices given in (2.1) and (2.2) in Theorem 2.4. The following proposition gives a strong sufficient condition for conjugate phase retrieval: n=1 be a frame taken from RM . If ker(ΩΦ) = {0}, then Φ Proposition 4.1. Let Φ = {ϕn}N is conjugate phase retrievable. In particular, if N = M (M + 1)/2 and det(ΩΦ) 6= 0, then Φ is conjugate phase retrievable. Proof. Let x, y ∈ CM be such that hx, ϕni2 = hy, ϕni2. Let also Q = Re(xx∗ − yy∗) Then using (3.6) in Lemma 3.3, we obtain 0 = hx, ϕni 2 − hy, ϕni 2 = hωϕn, v(Q)i for all n = 1, . . . , N . Putting all the equations together, we have a system of linear equations: ΩΦ(v(Q)) = 0. If ker ΩΦ = {0}, we must have v(Q) = 0. This is equivalent to Q = O and hence Re(xx∗) = Re(yy∗). By Theorem 2.1 (2), x ∼ y. Thus Φ is conjugate phase retrievable. If N = M (M + 1)/2, then ker(ΩΦ) = {0} if and only if det(ΩΦ) 6= 0, so the second statement follows. conj (cid:3) This theorem tells us that a generic frame with M (M + 1)/2 vectors on RM is conjugate phase retrievable. However, such conditions on N and the determinant in the previous proposition is far from necessary. We will see the number of vectors required for a generic frame to be conjugate phase retrievable is of order 4M in the next section. Nonetheless, this proposition is accurate when M = 2 and 3, which is what we are going to prove now. Proof of Theorem 2.4 when M = 2. We first prove the statement (1) in Theorem 2.4. For the sufficiency, we note that it has been proved in Proposition 4.1. We now prove the necessity. We note that for Φ =(cid:20)a1 a2 b1 b2 c1 c2(cid:21), det ΩΦ = det" a2 1 2a1a2 a2 2 b2 1 2b1b2 b2 2 c2 1 2c1c2 c2 2# = −2(a1b2 − a2b1)(a1c2 − a2c1)(b1c2 − b2c1) (after a computation by Mathematica). Suppose that Φ is conjugate phase retrievable. Then Φ possesses the complement property by Proposition 3.2, which means that any two vectors from Φ are linearly independent. In particular, this implies that none of the factors (a1b2 − a2b1), (a1c2 − a2c1), (b1c2 − b2c1) are zero. Hence, det ΩΦ 6= 0. As det ΩΦ = 0 is an algebraic equation, it defines an algebraic variety. Note that the complement of this 10 algebraic variety clearly cannot be empty. Thus, generic frames of three real vectors is conjugate phase retrievable. This also shows that N∗(2) ≤ 3. We now show that no Φ with two vectors is conjugate phase retrievable. This shows that N∗(2) = N∗(2) = 3. Indeed, if Φ has only two vectors, then it is obvious that Φ cannot have the complement property on C2 (by taking index subsets I,I c having only one element). Hence, Proposition 3.2 tells us that Φ cannot be conjugate phase retrievable. This finishes the proof. (cid:3) From the proof, we also notice that det Ψ 6= 0 if and only if Φ has the complement property, which gives the following simple characterization of conjugate phase retrievable real-valued frames in C2: Theorem 4.2. A real-valued frame Φ ⊆ R2 is conjugate phase retrievable if and only if Φ has the complement property. The proof for C2 is based on the complement property. However, the determinant in (2.2) becomes impossible to factorize. In fact, we will find that the simple characterization by the complement property in Theorem 4.2 cannot hold on C3 or higher. In the following, we will turn to studying the case C3 and prove Theorem 2.4 for M = 3 without using the complement property. Our idea is to first study the set Re(S 1,1 R ) for M = 3 and show that it will take all possibilities of symmetric matrices whose quadratic form is non-empty as a real algebraic variety. Then it will imply that a non-zero element in ker(ΩΦ) will correspond to some non-conjugate equivalent vectors. This idea is also workable for M = 2 and interested readers are invited to complete the same proof for M = 2. Lemma 4.3. Let Wx,y = Re(xx∗ − yy∗) and let Q be any M × M matrix with real entries. Then WQx,Qy = QWx,yQT . Proof. First, note that for any M × M matrix B with complex entries. Writing B = Re(B) + i Im(B), we have (4.1) Re(Q(B)QT ) = Re(Q Re(B)QT + iQ Im(B)QT ) = Q Re(B)QT since Q has real entries. Thus, with B = xx∗ − yy∗, WQx,Qy = Re((Qx)(Qx)∗ − (Qy)(Qy)∗) = Re(Qxx∗QT − Qyy∗QT ) = Re(Q(xx∗ − yy∗)QT ) = Q Re(xx∗ − yy∗)QT (by (4.1)) = QWx,yQT . (cid:3) Proposition 4.4. For any H ∈ H3×3 definite, there exists x, y ∈ C3 such that H = Re(xx∗ − yy∗). R 11 that is not positive semidefinite or negative semi- Proof. Denote by diag(λ1, λ2, λ3) the diagonal matrix with diagonal entries λ1, λ2, λ3. For any H that is not positive semidefinite or negative semidefinite, we can find an real or- thogonal matrix Q such that QT HQ = diag(a, b,−c) or diag(a, 0,−c) where a, b, c > 0. Suppose that we can find x, y ∈ C3 such that Wx,y = Re(xx∗ − yy∗) = diag(a, b. − c). Then H = QT diag(a, b,−c)Q = QT Wx,yQ = WQT x,QT y by Lemma 4.3. Therefore, it suffices to prove this proposition for diagonal matrices. Let x = (x1, x2, x3)T and y = (y1, y2, y3)T be two vectors in C3 and write them in exponential form: x = (x1eiθ1 x2eiθ2 x3eiθ3)T , y = (y1eiψ1 y2eiψ2 y3eiψ3)T . We are trying to solve for xi,yi, θi, ψi, i = 1, 2, 3 satisfying Re(xx∗−yy∗) = diag(a, b,−c), which can be written in the following sets of equations: x12 − y12 = a, x22 − y22 = b, x32 − y32 = −c These equations can be rewritten as ,  Re(x1x2) = Re(y1y2), Re(x1x3) = Re(y1y3), Re(x2x3) = Re(y2y3).  x1 =pa + y12, x2 =pb + y22, y3 =pc + x32 pa + y12pb + y22 cos(θ1 − θ2) = y1y2 cos(ψ1 − ψ2),  pa + y12x3 cos(θ1 − θ3) = y1pc + x32 cos(ψ1 − ψ3), pb + y22x3 cos(θ2 − θ3) = y2pc + x32 cos(ψ2 − ψ3). ,  x1x2 cos(θ1 − θ2) = y1y2 cos(ψ1 − ψ2), x1x3 cos(θ1 − θ3) = y1y3 cos(ψ1 − ψ3), x2x3 cos(θ2 − θ3) = y2y3 cos(ψ2 − ψ3). Putting the first set of equations into the second sets, we have (4.2) (4.3)  Case(i): diag(a, b,−c). We first set θ1 = ψ1, θ2 = ψ2, θ3 = ψ3 and θ1− θ2 = ψ1− ψ2 = π 2 . The above equations are satisfied if and only if which is equivalent to solving y1,y2,x3 satisfying (4.4) = (cid:26) pa + y12x3 = y1pc + x32, pb + y22x3 = y2pc + x32, y1pa + y12 x3pc + x32 R to [0, 1). Indeed, for any y ∈ [0, 1), we just take x =q ky2 y2pb + y22 = . 1−y2 ∈ R. We now notice that for any k > 0, the function f (x) = x√k+x2 is a surjective function from Hence, we can set y1 be free, and we have then x√c+x2 is surjective, we can always find y2 and x3 such that (4.4) holds. With y1,y2 and x3 chosen, we take x1 = pa + y12, x2 = pb + y22 and y3 = pc + x32 with θ1 = ψ1, θ2 = ψ2, θ3 = ψ3 and θ1 − θ2 = ψ1 − ψ2 = π 2 , then (4.2) holds. Hence, we have found x, y ∈ C3 such that Re(xx∗ − yy∗) = diag(a, b,−c). y1√a+y12 ∈ [0, 1). As x√b+x2 and 12 Case(ii): diag(a, 0,−c). In this case, (4.3) becomes pa + y12 cos(θ1 − θ2) = y1 cos(ψ1 − ψ2), pa + y12x3 cos(θ1 − θ3) = y1pc + x32 cos(ψ1 − ψ3), x3 cos(θ2 − θ3) =pc + x32 cos(ψ2 − ψ3).  pa + y12x3 = y1pc + x32 or equivalently = We take θi = ψi for i = 1, 2, 3 and θ1 − θ2 = ψ1 − ψ2 = π/2 and θ2 − θ3 = ψ2 − ψ3 = π/2. Then θ1 − θ3 = ψ1 − ψ3 = π and we have x√c+x2 implies that we can find x3 Hence, taking y1 free and surjectivity of the function satisfying the above equations. Now, taking also x2 = y2, equations (4.2) are satisfied and the proof is complete. (cid:3) y1pa + y12 . x3pc + x32 This proposition shows that the converse of Proposition 4.1 is true when M = 3. n=1 be a real-valued frame over C3. Then Φ is conjugate Theorem 4.5. Let Φ = {ϕn}N phase retrievable over C3 if and only if ker(ΩΦ) = {0}. Proof. We just need to prove that Φ is conjugate phase retrievable over C3 implies that ker(ΩΦ) = {0} since the other side was proved in Proposition 4.1. Suppose that ker(ΩΦ) 6= {0} and we take v ∈ ker(ΩΦ) and v 6= 0. Note that v ∈ RM (M +1)/2 and if we order v as v = (v11 ··· vM M v12 ··· v1M ··· v(M−1)M )T in a way analogous to (3.4), we can associate uniquely and naturally Qv = [vij] ∈ HM×M Hence, ΩΦv = 0 holds if and only if R . 0 = hωϕn, vi = ϕT n Qvϕn for all n = 1, . . . , N. Note that Qv cannot be positive semidefinite or negative semidefinite. If not, the above equation implies that ϕn = 0 for all n, which is impossible since ϕn forms a frame. Hence, Proposition 4.4 implies the existence of x, y ∈ C3 such that Re(xx∗ − yy∗) = Qv. As v 6= 0, so x and y are not conjugate equivalent. However, 0 = ϕT n Qvϕn = ϕT n (Re(xx∗ − yy∗))ϕn = hx, ϕni 2 − hy, ϕni2 by Lemma 3.3 (3.6). This means that Φ is not conjugate phase retrievable as it cannot distinguish x and y. (cid:3) We are now ready to prove Theorem 2.4 for M = 3. Proof of Theorem 2.4 when M = 3. We first prove statement (2) in Theorem 2.4. The sufficiency was proved in Proposition 4.1. For the necessity, we note that det(ΩΦ) 6= 0 if and only if ker(ΩΦ) = {0}. Hence, the necessity follows from Theorem 4.5. Finally, we also note that if N ≤ 5 < 6, then ker(ΩΦ) must be non-trivial. By Theorem 4.5, Φ cannot be conjugate phase retrievable. Hence, there is no conjugate phase retriev- able frame with cardinality less than 6. Combining with statement (2), we conclude that N∗(3) = N∗(3) = 6. (cid:3) The following example illustrates that the complement property is not sufficient to determine conjugate phase retrievable frames when M = 3. 13 hx, ϕi 2 − hy, ϕi 2 = ϕT (Re(xx∗ − yy∗))ϕ = ϕ2 1 + ϕ2 2 − ϕ2 3 = 0. Example 4.6. By Proposition 4.4, we can find x, y ∈ C3 such that Re(xx∗ − yy∗) = diag(1, 1,−1). Hence, x, y are not conjugate equivalent. Let Φ be a finite set of vectors taken from the cone x2 3. Then for any ϕ = (ϕ1, ϕ2, ϕ3)T ∈ Φ, we have 1 + x2 2 = x2 This shows that Φ cannot be conjugate phase retrievable. Since Φ is taken from the cone, it is easy to see that we can take Φ to span R3 or even satisfies the complement prop- erty. Hence, this also shows that the complement property is not sufficient to guarantee conjugate phase retrievability for M ≥ 3. 5. Generic Numbers In this section, we will be proving the generic number required for conjugate phase retrieval using real vectors. To this end, we need some terminology from algebraic geometry and we will use a theorem in a recent paper by Wang and Xu [22]. A subset V ⊂ CM is called a complex algebraic variety if V is the zero set in C of a collection of polynomials in C[x]. Let also VR be the set of all real points of V (i.e. VR = V ∩ RM ). We will be following the definition of dimension of and algebraic variety in [22, Section 3.1] (see also [12, Chapter 9]) and it is denoted by dim(·). For real algebraic variety X, its dimension is denoted by dimR(X). The set V is called a complex projective variety if V is the zero set in C of a collection of homogeneous polynomials in C[x]. Definition 5.1. Let V be a complex projective variety with dim V > 0 and let ℓα : CM → C, α ∈ I (I is an index set), be a family of linear functions. We say that V is called admissible with respect to {ℓα : α ∈ I} if dim(V ∩ {x ∈ CM : ℓα(x) = 0}) < dim V for all α ∈ I. This admissibility is equivalent to the property that for a generic point x ∈ V and any small neighborhood U of x, U ∩V is not completely contained in the hyperplane ℓα(x) = 0. Theorem 5.1. [22, Theorem 3.2 and Corollary 3.3] For j = 1, . . . , N , let Lj : Cn × Cm → C be bilinear functions and Vj be complex projective varieties on Cn. Set V = V1 × ··· × VN ⊂ (Cn)N . Let W be an complex projective variety. Suppose that for each j, Vj is admissible with respect to the linear functions {f w(·) := Lj(·, w) : w ∈ W \{0}}. We have the following conclusions: (1) If N ≥ dim W , then there exists an algebraic variety Z ⊂ V with dim Z < dim V j=1 ∈ V \ Z and w ∈ W , Lj(xj, w) = 0 for all such that for any X = (xj)N j = 1, . . . , N implies w = 0. dimRVR such that for any X = (xj)N j = 1,··· , N implies w = 0. (2) If dim VR = dim V, then there exists a real algebraic variety eZ ⊂ VR with dimReZ < j=1 ∈ VR \eZ and w ∈ W , Lj(xj, w) = 0 for all Proof of Theorem 2.5. The proof is inspired by Theorem 4.1 in [22]. We will let Vr be the complex algebraic variety of M × M complex symmetric matrices (i.e. AT = A, and A has complex entries) with rank at most r. This is a complex projective variety defined by the homogeneous polynomials vanishing on all (r + 1) × (r + 1) minors. The real points (Vr)R are all the real symmetric matrices with rank at most r. In our notation, (Vr)R = S r R. Moreover, dimVr = dimR((Vr)R) = M r − 14 r(r − 1) 2 (For this fact, see Theorem 4.1 [22]). Consider bilinear functions Lj : C M×M × CM×M and V = V1 × .... × V1 ⊂ (CM×M )N . (i.e. N copies of V1). In particular, we will consider Lj(A, Q) = T r(AQ), for j = 1, . . . , N less than that of VR such that for every (ϕnϕT n )N Lj(ϕnϕT n , Q) = ϕT (T r denotes the trace of the matrix). If A ∈ V1 and positive semidefinite, then A = ϕϕT and Lj(A, Q) = ϕT Qϕ. Let W = V4. Then dimV4 = 4M − 6. Assume we can prove that V1 is admissible with respect to {f Q(·) = Lj(·, Q) : Q ∈ W}. Then Theorem 5.1 (2) and the fact that positive semidefinite matrices of rank 1 is open in V1 (See [22, Remark after Theorem 4.1]) imply that there exists an real algebraic variety eZ with dimension strictly n=1 ∈ VR \eZ, the following property holds: But then this implies ker(A) ∩ S 4 conjugate phase retrievable by Theorem 2.3. R ⊂ W ). Hence, generic frame will be It remains to show V1 is admissible with respect to {f Q(·) = Lj(·, Q) : Q ∈ W}. It suffices to show that a generic point A0 ∈ V1 and any non-zero Q0 ∈ W , we must have T r(AQ0) 6≡ O in any small neighborhood of A0 in V1. If T r(A0Q0) 6= 0, then we are done. So we assume that T r(A0Q0) = 0. In this case, we factorize A0 = uvT and for any fixed z, w ∈ CM , consider Then n Qϕn = 0 for all n = 1, .., N and Q ∈ W =⇒ Q = O. R = {O} (since S 4 At = (u + tz)(v + tw)T T r(AtQ0) = T r((u + tz)(v + tw)T Q0) = t(T r(uwT + zvT )Q0) + t2T r(zwT Q0). As Q0 6= O, we can find z, w such that wT Q0z 6= 0, so that T r(zwT Q0) = T r(wT Q0z) 6= 0. Thus, for any sufficiently small t, T r(AQ0) 6≡ O in any small neighborhood of A0 in V1. This completes the whole proof. (cid:3) 6. Strict Conjugate Phase retrievability In this section, we are going to give a systematic study of general frame Φ ⊂ CM (not necessarily real vectors) that are conjugate phase retrievable. Of course, we know that a complex phase retrievable frame must be conjugate phase retrievable. Our interest will be frames in the following definition. Definition 6.1. We say a frame is strictly conjugate phase retrievable if the frame is conjugate phase retrievable but not complex phase retrieval. Complex phase retrieval fails using real vectors because there always exist x and x that are not equivalent up to phase. A natural question that arises is: what are the vectors x which are equivalent to x up to a phase (i.e. x ∼ x)? It turns out that these vectors will all be phased real vectors. Moreover, they will give us an important characterization for strictly conjugate phase retrievable frame. Definition 6.2. We say that y is a phased real vector if y belongs to the following set: ϑRM = {λv λ ∈ T, v ∈ RM} Proposition 6.1. A vector y ∈ CM is equivalent to its conjugate y up to a global phase if and only if y ∈ ϑRM . 15 Proof. It is clear that if y ∈ ϑRM , then y ∼ y. Suppose y ∼ y. Then there exists 0 ≤ θ < 2π such that yn = eiθyn for all n = 1, . . . , M . Writing yn = yneiθn, it follows that eiθn = ei(θ−θn). Thus, 2θn = θ + 2πk for some k ∈ N, which implies that θn = θ 2 + πk. Hence, eiθn = ei θ 2 for each n = 1, . . . , M with sign depending on n. Thus, y = λv with λ = ei θ (cid:3) 2 . Therefore, yn = ±ynei θ 2 or −ei θ 2 and v = (±y1 ± y2 ··· ± ym)T . Note that Proposition 6.1 implies that no frame Φ ⊆ ϑRM is complex phase retrievable. i=1 is a frame over CM that is conjugate phase Theorem 6.2. Suppose that Φ = {ϕn}N retrievable. Then, Φ is strictly conjugate phase retrievable if and only if there exists some y ∈ CM with y /∈ ϑRM but hy, ϕni2 = hy, ϕni 2 for all n ∈ {1, . . . , N}. Proof. Suppose that Φ is strictly conjugate phase retrievable. Then, there exist x, y ∈ Cm such that hx, ϕni 2 = hy, ϕni2 for all n ∈ {1, . . . , N}, with x 6∼ y but x ∼ y. Since ∼ is transitive , y ∼ y would imply that x ∼ y, a contradiction. Hence, y 6∼ y and we conclude that y /∈ ϑRM . With x ∼ y, we can write x = λy for some unimodular scalar λ, which gives hy, ϕni 2 = hx, ϕni 2 = hλy, ϕni 2 = hy, ϕni2. Thus, hy, ϕni 2 = hy, ϕni2 for all n ∈ {1, . . . , N}. This shows the necessity. Conversely, suppose that there exists some y ∈ CM with y /∈ ϑRM and hy, ϕni 2 = hy, ϕni2 for all n = 1, . . . , N. Since y /∈ ϑRM implies y 6∼ y it follows that Φ is not complex phase retrievable and is only strictly conjugate phase retrievable by the original assumption. (cid:3) Strict conjugate phase retrieval relates directly back to phased real vectors. The fol- lowing set of equations characterize those frames which strictly allow conjugate phase retrieval. n=1 be a conjugate phase retrievable frame in CM where Proposition 6.3. Let Φ = {ϕn}N ϕn = (ϕ1n ϕ2n ··· ϕM n)T for n ∈ {1, . . . , N}. Then Φ is strictly conjugate phase retriev- able if and only if there exists some x = (x1 ··· xM )T ∈ CM , with x /∈ ϑRM and Xj<k (6.1) Im(xjxk) Im(ϕjnϕkn) = 0 for each n = 1, . . . , N. Proposition 6.3 will be a consequence of the following lemma. Lemma 6.4. For x = (x1 ··· xM )T , ϕ = (ϕ1 ··· ϕM )T ∈ CM , Im(xjxk) Im(ϕj ϕk) = 0. Proof of Lemma 6.4. Let x = (x1 ··· xM )T , ϕ = (ϕ1 ··· ϕM )T ∈ CM . Expanding using the definition of the conjugate, we may write hx, ϕi2 = hx, ϕi2 if and only if Xj<k hx, ϕi 2 = xjϕj MXk=1 MXj=1 MXk=1 MXj,k=1,j6=k xkϕk2 + = 16 xkϕk! = MXj,k=1 xjϕjxkϕk xjϕjxkϕk. Thus, hx, ϕi 2 − hx, ϕi 2 = = = MXj,k=1, j6=k MXj,k=1, j6=k MXj,k=1, j6=k xjϕjxkϕk − xjϕjxkϕk ϕjϕk(xjxk − xjxk) ϕjϕk(2i Im(xjxk). Now, for any fixed j 6= k, we observe that we have the equality ϕjϕk(2i Im(xjxk)) = ϕkϕj(2i Im(xkxj). Therefore, we can split our sum into a sum over indices with j < k and a sum over indices with k < j, MXj,k=1, j6=k ϕjϕk2i Im(xjxk) =Xj<khϕjϕk2i Im(xjxk) + ϕjϕk2i Im(xjxk)i =Xj<k =Xj<k =Xj<k 4 Re(i(ϕj ϕk Im(xjxk))) −4 Im(ϕjϕk Im(xjxk)) −4 Im(xjxk) Im(ϕjϕk). Therefore, hx, ϕi 2 = hx, ϕi 2 if and only ifXj<k Proof of Proposition 6.3. Suppose Φ is strictly conjugate phase retrievable. By Theo- rem 6.2, there exists some x ∈ CM with x 6∼ x and hx, ϕni 2 = hx, ϕni 2 for n = 1, . . . , N. Using Lemma 6.4 with x and ϕn for each n = 1, . . . , N completes this direction of the proof. Im(xjxk) Im(ϕjϕk) = 0. (cid:3) Suppose there exists a vector x /∈ ϑRM and that Im(xixj) Im(ϕinϕjn) = 0 for each n ∈ [N ]. Xi<j Then, Lemma 6.4 implies that hx, ϕni2 = hx, ϕni2 for each n = 1, . . . , N which gives that Φ is not complex phase retrievable. (cid:3) Note that given any conjugate phase retrievable Φ ⊆ ϑRM , equation (6.1) holds for any ϕ ∈ Φ and x ∈ CM because Im(ϕjϕk) are always zero. Hence, Proposition 6.3 implies Φ is strictly conjugate phase retrievable. In the following, we show that in C2, every strictly conjugate phase retrievable frame is a frame in ϑRM . Theorem 6.5. Any frame over C2 that is strictly conjugate phase retrievable must be a frame contained in ϑRM . Furthermore, we have the following consequence: (1) Any frame Φ 6⊆ ϑRM on C2 that is conjugate phase retrievable must be complex (2) On the other hand, any real-valued frame Φ ⊂ R2 on C2 that is conjugate phase phase retrievable and have at least four vectors. retrievable requires only at least three vectors. 17 (cid:20) ϕ11 ϕ12 ϕ21 ϕ22 ··· ϕ1n ··· ϕ2n (cid:21) = ϕ1 ϕ2 ··· ϕn  . Proof. Let Φ = {ϕ1, ϕ2, . . . , ϕn} be a strictly conjugate phase retrievable frame over C2. We first write the frame matrix of Φ as By Theorem 6.3, there exists y = (y1 y2)T in C2 with y /∈ ϑRM and Im(ϕ11ϕ21) Im(y1y2) = 0 Im(ϕ12ϕ22) Im(y1y2) = 0 ... Im(ϕ1nϕ2n) Im(y1y2) = 0. By assumption, y 6∼ y, and we must have y1y2 6= y1y2 = y1y2 and thus Im(y1y2) 6= 0. To satisfy the above list of equations we must then have Im(ϕ11ϕ21) = ··· = Im(ϕ1nϕ2n) = 0. For any frame vector ϕi, we have Im(ϕ1iϕ2i) = 0, which implies ϕi ∈ ϑRM . Thus, Φ ⊆ ϑRM . Thus, we can say that any strictly conjugate phase retrievable frame over C2 is a frame in ϑRM . To prove (1), suppose that Φ is conjugate phase retrievable and Φ 6⊆ ϑRM . By what we just proved, Φ is not strictly conjugate phase retrievable, Thus, we must have that Φ is complex phase retrievable on C2. In [7] it was proved that a minimum of four vectors is required for complex phase retrieval on C2. This completes the proof. Statement (2) has been proved in Theorem 2.4. (cid:3) 7. Discussions and Open Questions We end this paper with a discussion of some problems concerning conjugate phase retrieval that is also in line with the current research about phase retrieval. M ≥ 4. 7.1. Conjugate Phase Retrieval in High Dimension. For M ≥ 4, the following two questions are naturally raised: (1) Compute N∗(M ) and N∗(M ) for conjugate phase retrieval of real frames when (2) Determine if N∗(M ) < N∗(M ) can happen for conjugate phase retrieval. In Theorem 2.5 we showed that N∗(M ) ≤ 4M − 6 and N∗(M ) ≤ N∗(M ) ≤ 4M − 6 for M ≥ 4. In comparison with complex phase retrieval with the same notation for N∗(M ) and N∗(M ), N∗(M ) ≤ 4M − 4 in any dimension M for complex phase retrieval, but it is also known that when M = 4, there exists a frame of 11 vectors that also does complex phase retrieval [21]. In other words, N∗(4) ≤ 11 < 4(4) − 4 = 12. Notice that the proof for 4M − 6 generic vectors performing phase retrieval uses the R = {O} in Theorem 2.3 (2). However, a weaker C ) = {O} is already enough. Unfortunately, we do not know C ) is a real projective variety, nor its real dimension. Therefore, we cannot use C ) should R when M ≥ 4. In view of this, we believe that N∗(4) < 10 = sufficient condition that ker(A) ∩ S 4 condition that ker(A)∩ Re(S 1,1 if Re(S1,1 Theorem 5.1 to obtain a sharper result. Furthermore, we conjecture that Re(S 1,1 be strictly contained in S 4 4(4) − 6 is highly possible to happen for conjugate phase retrieval with real frames. 18 7.2. Strict Conjugate Phase Retrieval. Another interesting question raised up is to know which vectors perform strict conjugate phase retrieval. We showed that strictly conjugate phase retrievable frames in C2 come entirely from phased real vectors ϑRM . Is this true in higher dimensions? If not, what other frames Φ 6⊂ ϑRM are strictly conjugate phase retrievable for M > 2? 7.3. Conjugate Unsigned Sampling. Recent studies about phase retrieval on real- valued bandlimited functions and also on shift-invariant spaces can be found in [1, 3, 2, 10, 20]. However, as indicated in the introduction, phase retrieval on complex-valued Paley-Wiener space bandlimited on [−b/2, b/2] (P Wb) is impossible using real samples. With the notion of conjugate phase retrieval, we ultimately wish to recover complex-valued functions in P Wb from real samples. We propose the following definition for recovery up to conjugacy in P Wb and a natural question is raised: Definition 7.1. Let Λ be a countable subset of R. We say Λ is a set of conjugate unsigned sampling for P Wb if for any f, g ∈ P Wb f (λ) = g(λ) for all λ ∈ Λ implies that f = eiθg or f = eiθg for some 0 ≤ θ ≤ 2π. (Qu) Does there exist a set Λ ⊆ R that forms a set of conjugate unsigned sampling on P Wb? In [18, 19], the authors proved the possibility of complex phase retrieval on P W and Bernstein spaces under a very specific measurement setup with samples taken over the complex plane. In their work, a sampling density of four times of the bandwidth is also recorded. With the results studied in this paper, conjugate unsigned sampling by real numbers may be possible and its density should be at least four times of the bandwidth. Acknowledgment. Chun-Kit Lai would like to thank professors Deguang Han and Jame- son Cahill for some early enlightening discussions about phase retrievals. References [1] R. Alaifari, I. Daubechies, P. Grohs, and G. Thakur. Reconstructing real-valued functions from un- signed coefficients with respect to wavelet and other frames. Journal of Fourier Analysis and Appli- cations, pages 1 -- 15, 2016. [2] R. Alaifari, I. Daubechies, P. Grohs, and R. Yin. Stable phase retrieval in infinite dimensions. arXiv preprint arXiv:1609.00034, 2016. [3] R. Alaifari and P. Grohs. Phase retrieval in the general setting of continuous frames for banach spaces. Technical Report 2016-21, Seminar for Applied Mathematics, ETH Zurich, Switzerland, 2016. [4] S. Bahmanpour, J. Cahill, P. G. Casazza, J. Jasper, and L. M. Woodland. Phase retrieval and norm retrieval. In Trends in harmonic analysis and its applications, volume 650 of Contemp. Math., pages 3 -- 14. Amer. Math. Soc., Providence, RI, 2015. [5] R. Balan. Reconstruction of signals from magnitudes of redundant representations: The complex case. Found. Comput. Math, 16:677 -- 721, 2016. [6] R. Balan, P. Casazza, and D. Edidin. On signal reconstruction without phase. Appl. Comput. Harmon. Anal., 20(3):345 -- 356, 2006. [7] A. S. Bandeira, J. Cahill, D. G. Mixon, and A. A. Nelson. Saving phase: injectivity and stability for phase retrieval. Appl. Comput. Harmon. Anal., 37(1):106 -- 125, 2014. [8] O. Bunk, A. Diaz, F. Pfeiffer, C. David, B. Schmitt, D. K. Satapathy, and J. F. van der Veen. Diffrac- tive imaging for periodic samples: retrieving one-dimensional concentration profiles across microfluidic channels. Acta Crystallographica Section A, 63(4):306 -- 314, Jul 2007. [9] A. Burvall, U. Lundstrom, P. A. C. Takman, D. H. Larsson, and H. M. Hertz. Phase retrieval in x-ray phase-contrast imaging suitable for tomography. Opt. Express, 19(11):10359 -- 10376, May 2011. 19 [10] Y. Chen, C. Cheng, Q. Sun, and H. Wang. Phase retrieval of real-valued signals in a shift-invariant space. https://arxiv.org/pdf/1603.01592.pdf, 2016. [11] A. Conca, D. Edidin, M. Hering, and C. Vinzant. An algebraic characterization of injectivity in phase retrieval. Appl. Comput. Harmon. Anal., 38(2):346 -- 356, 2015. [12] D. Cox, J. Little, and D. O'Shea. Ideals, variety and algorithms, an introduction to computational algebraic geometry and commutative algebra, 3rd edition. Springer, 2007. [13] J. R. Fienup. Reconstruction of an object from the modulus of its fourier transform. Opt. Lett., 3(1):27 -- 29, Jul 1978. [14] J. R. Fienup. Phase retrieval algorithms: a comparison. Appl. Opt., 21(15):2758 -- 2769, Aug 1982. [15] T. Heinosaari, L. Mazzarella, and M. M. Wolf. Quantum tomography under prior information. Comm. Math. Phys., 318(2):355 -- 374, 2013. [16] M. Kech and M. Wolf. From quantum tomography to phase retrieval and back. In 2015 International Conference on Sampling Theory and Applications (SampTA), pages 173 -- 177, May 2015. [17] S. Mallat and I. Waldspurger. Phase retrieval for the cauchy wavelet transform. Journal of Fourier Analysis and Applications, 21(6):1251 -- 1309, Dec 2015. [18] V. Pohl, F. Yang, and H. Boche. Phaseless signal recovery in infinite dimensional spaces using struc- tured modulations. J. Fourier Anal. Appl., 20:1212 -- 1233, 2014. [19] V. Pohl, F. Yang, and H. Boche. Phase retrieval via structured modulations in paleyvwiener spaces. Proceedings of 10th International Conference on Sampling Theory and Applications (SampTA), July 2013, July 2013. [20] G. Thakur. Reconstruction of bandlimited functions from unsigned samples. J. Fourier Anal. Appl., 17(4):720 -- 732, 2011. [21] C. Vinzant. A small frame and a certificate of its injectivity. In 2015 International Conference on Sampling Theory and Applications (SampTA), pages 197 -- 200, May 2015. [22] Y. Wang and Z. Xu. Generalized phase retrieval : measurement number, matrix recovery and beyond. https://arxiv.org/abs/1605.08034, 2016. [23] T. Wiatowski and H. Bolcskei. A mathematical theory of deep convolutional neural networks for feature extraction. IEEE Transactions on Information Theory, Dec. 2015. Department of Mathematics, University of Maryland, College Park, 4176 Campus Drive, College Park, MD 20742. E-mail address: [email protected] Department of Mathematics, San Francisco State University, 1600 Holloway Avenue, San Francisco, CA 94132. E-mail address: [email protected] 20
1004.3751
1
1004
2010-04-21T17:29:24
On the higher rank numerical range of the shift operator
[ "math.FA" ]
For any n-by-n complex matrix T and any $1\leqslant k\leqslant n$, let $\Lambda_{k}(T)$ the set of all $\lambda\in \C$ such that $PTP=\lambda P$ for some rank-k orthogonal projection $P$ be its higher rank-k numerical range. It is shown that if $\bbS$ is the n-dimensional shift on ${\C}^{n}$ then its rank-k numerical range is the circular disc centred in zero and with radius $\cos\dfrac{k\pi}{n+1}$ if $1<k\leqslant\left[\frac{n+1}{2} \right]$ and the empty set if $\left[\frac{n+1}{2} \right]<k\leqslant n$, where $\left[x \right] $ denote the integer part of $x$. This extends and rafines previous results of U. Haagerup, P. de la Harpe \cite{Haagerup} on the classical numerical range of the n-dimensional shift on${\C}^{n}$. An interesting result for higher rank-$k$ numerical range of nilpotent operator is also established.
math.FA
math
ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR HAYKEL GAAYA centred in zero and with radius cos Abstract. For any n-by-n complex matrix T and any 1 6 k 6 n, let Λk(T ) the set of all λ ∈ CI such that P T P = λP for some rank-k orthogonal pro- jection P be its higher rank-k numerical range. It is shown that if Sn is the n-dimensional shift on CI n then its rank-k numerical range is the circular disc 2 (cid:3) and the empty set if (cid:2) n+1 2 (cid:3) < k 6 n, where [x] denote the integer part of x. This extends and rafines previous results of U. Haagerup, P. de la Harpe [8] on the classical numerical range of the n-dimensional shift on CI n. An interesting result for higher rank-k numerical range of nilpotent operator is also established. if 1 < k 6 (cid:2) n+1 kπ n + 1 1. Introduction Let H be a complex separable Hilbert space and B(H) the collection of all bounded linear operator on H. The numerical range of an operators T in B(H) is the subset W (T ) = {< T x, x >∈ CI ; x ∈ H, kxk6 1} of the plane, where < ., . > denotes the inner product in H and the numerical range of T is defined by We denote by S the unilateral shift acting on the Hardy space HI 2 of the square ω2(T ) = sup {z; z ∈ W (T )} . summable analytic functions. S : HI 2 → HI 2 7→ zf (z) f Beurling's theorem implies that the non zero invariant subspaces of S are of the forme φ HI 2, where φ is some inner function . Let S(φ) denote the compression of S to the space H(φ) = HI 2 ⊖ φ HI 2 : S(φ)f (z) = P (zf (z)), where P denotes the ortogonal projection from HI 2 onto H(φ). The space H(φ) is a finite-dimensional exactly when φ is a finite Blaschke product. The numerical radius and numerical range of the model operator S(φ) seems to be important and have many applications. In [1], Badea and Cassier showed that there is relation- ship between numerical radius of S(φ) and Taylor coefficients of positive rational functions on the torus and more recently in [6], the author gave an extension of this result. However the evaluation of the numerical radius of S(φ) under an explicit form is always an open problem. The reader may consult [6] for an estimate of S(φ) 2000 Mathematics Subject Classification. 47A12, 47B35. Key words and phrases. Operator theory, Numerical radius, Numerical range, higher rank numerical range, Eigenvalues, Toeplitz forms. 1 2 HAYKEL GAAYA where φ is a finite Blashke product with unique zero. In the particular case where φ(z) = zn, S(φ) is unitarily equivalent to Sn where Sn =  0 1 . . . . . . . . . 1 0 .   n+1o In [8]; it is proved that W (Sn) is the closed disc Dn = nz ∈ CI ; z6 cos π and ω2(Sn) = cos π n+1 and more general Theorem 1.1 ([8]). Let T be an operator on H such that T n = 0 for some n ≥ 2. One has: ω2(T ) 6 kT kcos π n + 1 and ω2(T ) = kT kcos π n+1 when T is unitarily equivalent to kT kSn. In this mathematical note, we extend this result to the higher rank-k numerical range of the shift. The notion of the the higher rank-k numerical range of T ∈ B(H) is introduced in [4] and it's denoted by: Λk(T ) = {λ ∈ CI : P T P = λP for some rank-k orthogonal projection P } , The introduction of this notion was motivated by a problem in quantum error correction; see [5]. If P is a rank-1 orthogonal projection then P = x ⊗ x for some x ∈ CI n and P T P =< T x, x > P . Then when k = 1, this concept is reduces to the classical numerical range W (T ), which is well known to be convex by the Toeplitz- Hausdorff theorem; for exemple see [10] for a simple proof. In [2], it's conjectured that Λk(T ) is convex, and reduced the convexity problem to the problem of showing that 0 ∈ Λk(T ′) where T ′ =(cid:18) Ik X Y −Ik (cid:19) for arbitrary X, Y ∈ Mk (the algebra of k × k complex matrix). They further reduced this problem to the existance of a Hermitian matrix H satisfying the matrix equation (1.1) Ik + M H + HM ∗ − HRH = H for arbitrary M ∈ Mk and a positive definite R ∈ Mk. In [16], H. Woerdeman proved that equation (1.1) is equivalent to Ricatti equation: (1.2) HRH − H(M ∗ − Ik/2) − (M − Ik/2)H − Ik = 0k, and using the theory of Ricatti equations (see [9], Theorem 4), the equation (1.2) is solvable which prove the convexity of Λk(T ). In [4], the authors showed that if dimH < ∞ and T ∈ B(H) is a Hermitian matrix with eigenvalues λ1 6 λ2 · · · 6 λn then the rank-k nuemrical range Λk(T ) coincides with [λk, λn+1−k] which is a non- degenerate closed interval if λk < λn+1−k, a singleton set if λk = λn+1−k and an empty set if λk > λn+1−k. In [13], the authors proved that if dimH = n Λk(T ) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθT + e−iθT ∗(cid:1)(cid:9) , ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR 3 for 1 6 k 6 n, where λk(H) denote the kth largest eigenvalue of the hermitian matrix H ∈ Mn. This result establishes that if dimH = n and T ∈ B(H) is a normal matrix with eigenvalues λ1, . . . , λn then Λk(T ) = \ 16j1<···<jn−k+16n conv(cid:8)λj1 , . . . , λjn+1−k(cid:9) . We close this section by the following properties wich are easly checked. The reader may consult [2],[3],[4],[5],[7] and [11]. P1. For any a and b ∈ CI , Λk(aT + bI) = aΛk(T ) + b. P2. Λk(T ∗) = Λk(T ). P3. Λk(T ⊕ S) ⊇ Λk(T ) ∪ Λk(S). P4. For any unitary U ∈ B(H), Λk(U ∗T U ) = Λk(T ). P5. If T0 is a compression of T on a subspace H0 of H such that dimH0 ≥ k, then Λk(T0) ⊆ Λk(T ). P6. W (T ) ⊇ Λ2(T ) ⊇ Λ3(T ) ⊇ . . . . Some results from [1] will be also developed in this context in a forthcoming paper. 2. main theorem In the following theorem we give the higher rank-k numerical range of the n- dimensional shift on CI n. Theorem 2.1. For any n ≥ 2 and 1 6 k 6 n, Λk(Sn) coincides with the circular disc {z ∈ CI : z6 cos kπ n + 1 } if 1 6 k 6(cid:2) n+1 2 (cid:3) and the empty set if(cid:2) n+1 2 (cid:3) < k 6 n. Proof. First observe that Λk(Sn) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθSn + e−iθS ∗ n(cid:1)(cid:9) n(cid:1)(cid:27) λk(cid:0)eiθSn + e−iθS ∗ n(cid:1)(cid:27) λk(cid:0)eiθSn + e−iθS ∗ = \θ∈[0,2π[(cid:26)µ ∈ CI : Re(eiθµ) 6 eiθ(cid:26)z ∈ CI : Re(z) 6 = \θ∈[0,2π[ 1 2 1 2 (2.1) On the other hand, we have eiθSn + e−iθS ∗ n =   0 eiθ 0 ... 0 0 e−iθ 0 eiθ ... 0 0 0 e−iθ 0 ... 0 0 . . . . . . . . . . . . . . . . . . 0 0 0 ... 0 eiθ 0 . . . 0 . . . 0 ... e−iθ 0 .   Note that eiθSn + e−iθS ∗ n is a Toeplitz matrix associated to the Toeplitz form fθ(t) = 2 cos(θ + t). 4 HAYKEL GAAYA The eigenvalues satisfy the caracteristic equation ∆n(λ) = Det(cid:0)eiθSn + e−iθS ∗ n(cid:1) 0 = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −λ e−iθ eiθ −λ eiθ 0 ... ... 0 0 0 0 0 0 0 ... 0 . . . 0 . . . . . . . . . . . . 0 ... . . . . . . −λ e−iθ eiθ −λ . . . e−iθ −λ ... 0 0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Expanding this determinant, we obtain the recurrence relation ∆n(λ) = −λ∆n−1 − ∆n−2, n = 2, 3, 4, . . . , This recurrence relation holds also for n = 1 provided we put ∆0 = 1 and ∆−1 = 0. In order to find an explicit representation of ∆n(λ), we write convenently λ = 2 cos(θ + t) = fθ(t) and form the caracteristic equation ρ2 = −λρ − 1 = −2ρ cos(θ + t) − 1 with the roots −ei(θ+t) and −e−i(θ+t) so that ∆n(2 cos(θ + t)) = (−1)n(Aein(θ+t) + Be−in(θ+t)) where the constants A and B can be determined from the cases n = −1 and n = 0. Thus ∆n(2 cos(θ + t)) = (−1)n sin((n + 1)(θ + t)) sin(θ + t). This yields the eigenvalues λν = 2 cos( νπ n + 1 ), ν = 1, 2, . . . n. This implies of course that Λk(Sn) = \θ∈[0,2π[ eiθ(cid:26)z ∈ CI : Re(z) 6 cos( kπ n + 1 )(cid:27) Thus Λk(Sn) is the intersection of closed half planes. We note that cos( positive if and only if k 6(cid:2) n+1 2 (cid:3). Case 1.If k 6(cid:2) n+1 Case 2. If k >(cid:2) n+1 Λk(Sn) ⊆ (cid:26)z ∈ CI : Re(z) 6 cos( 2 (cid:3) In this case Λk(Sn) is circular disc {z ∈ CI : z6 cos 2 (cid:3), then )(cid:27)\ eiπ(cid:26)z ∈ CI : Re(z) 6 cos( n + 1 kπ = ∅. kπ n + 1 ) is kπ n + 1 }. kπ n + 1 )(cid:27) (cid:3) This completes the proof. On the sequel of this paper, let denote by ρ(k, r) =(cid:26) k/r [k/r] + 1 unless if k/r is is integer ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR 5 where k and r are arbitrary numbers. Lemma 2.2. For a fixed n ≥ 1 and r ≥ 1, let denote by λ1 > · · · > λn; n real numbers and (λ′ p)16p6nr a finite sequence defined by: λ′ 1 = · · · = λ′ r = λ1, . . . , λ′ (n−1)r+1 = · · · = λ′ nr = λn. Then for each 1 6 k 6 nr, the kth largest term of (λ′ t)16t6nr is λρ(k,r). Proof. The claim is obvious in the case where r = 1. We may assume r ≥ 2. We prove the result by induction on k. If k = 1, then the largest term is λ1 = λρ(1,r). So the result hold for k = 1. Assume that k > 1, and the reslut is valid for the mth largest term of (λ′ t)16t6nr whenever m < k. Case 1. Suppose that ρ(k − 1, r) = k−1 k − 1 = pr. By induction assumption, we have λρ(k−1,r) = λ′ that the kth largest term of (λ′ r , then there exists 1 6 p 6 n − 1 such that pr = λp, which implies t)16t6nr is λ′ pr+1 = λp+1 = λ k−1 r +1 = λ[ k r ]+1 = λρ(k,r). Case 2. Suppose that ρ(k − 1, r) = [ k−1 r ] + 1, then there exist 1 6 q 6 n − 1 and 1 6 s 6 r − 1 such that k − 1 = qr + s. First, note that ρ(k − 1, r) = ρ(k, r). On the other hand, by induction assumption, we have λρ(k−1,r) = λ′ qr+s = λq+1. Consequently the kth largest term of (λ′ t)16t6nr is λ′ qr+s+1 = λq+1 = λρ(k−1,r) = λρ(k,r). The proof is now complete. (cid:3) Let DT = (IN − T ∗T )1/2 be the defect operator of T and DT the closed range of DT . Let denote by r = dimDT . Theorem 2.3. Consider T ∈ B(H) such that kT k6 1 and T n = 0. Then Λk(T ) is contained in the circular disc {z ∈ CI : z6 cos( ρ(k,r)π n+1 )} if 1 6 ρ(k, r) 6 [ n+1 2 ] and empty if ρ(k, r) > [ n+1 2 ]. Proof. If T is a contaction with T n = 0, then T can be viewed as a compression of n acting on the Hilbert space DT ⊗ CI n. Consider the isometry H → DT ⊗ CI n, Ir ⊗S ∗ where {el}n l=1 is the canonical basis of CI n. Note that DT T t−1x ⊗ et t=1 V (x) =Xn DT T tx ⊗ et =Xn−1 t=1 V T x =Xn t=1 It follows that DT T tx ⊗ et = (Ir ⊗ S ∗ n)V x. and from (P.5) T = V ∗(Ir ⊗ S ∗ n)V (2.2) Λk(T ) = Λk(V ∗(Ir ⊗ S ∗ n)V ) ⊆ Λk(Ir ⊗ S ∗ n), for any 1 6 k 6 nr. 6 HAYKEL GAAYA Now, Λk(Ir ⊗ S ∗ n) n) + e−iθ(Ir ⊗ S ∗ n)∗(cid:1)(cid:9) n) + e−iθ(Ir ⊗ Sn)(cid:1)(cid:9) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθ(Ir ⊗ S ∗ = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθ(Ir ⊗ S ∗ = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)Ir ⊗ (eiθSn + e−iθS ∗ n)(cid:1)(cid:9) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)⊕r n)(cid:1)(cid:9) = \θ∈[0,2π[ eiθ(cid:26)z ∈ CI : Re(z) 6 cos( i (eiθSn + e−iθS ∗ ρ(k, r)π n + 1 ) (cid:27) where the last equality is due to the lemma (2.2) and theorem (2.1). Thus Λk(Ir ⊗ S ∗ n) =( D(0, cos( ρ(k,r)π n+1 )) ∅ if 1 6 ρ(k, r) 6 [ n+1 2 ] if [ n+1 2 ] < ρ(k, r) 6 n Therefore, if 1 6 k 6 nr, (2.2) implies that Λk(T ) ⊆ D(0, cos( ρ(k,r)π and empty if [ n+1 2 ] < ρ(k, r) 6 n. Finally, if k > nr, Λk(T ) = ∅ from (P6). n+1 )) if 1 6 ρ(k, r) 6 [ n+1 2 ] (cid:3) Corollary 2.4 (U. Haagerup, P. de la Harpe,[8]). Consider T ∈ B(H) such that kT k6 1 and T n = 0. Then we have ω2(T ) 6 cos( π n+1 ). Proof. T = V ∗(Ir ⊗ S ∗ n)V where V : H → DT ⊗ CI n, Now V (x) =Xn t=1 DT T t−1x ⊗ et. W (T ) = Λ1(T ) = Λ1(V ∗(Ir ⊗ Sn)V ) ⊆ Λ1(Ir ⊗ Sn) = D(0, cos π n + 1 ). (cid:3) Acknowledgements: The author would like to express his gratitude to Gilles Cassier for his help and his good advices. References [1] C. Badea and G. Cassier, Constrained von neumann inequalities, Adv. Math. 166 (2002), no. 2, 260 -- 297. [2] M.-D. Choi, M. Giesinger, J. A. Holbrook, and D. W. Kribs, Geometry of higher-rank nu- merical ranges, Linear and Multilinear Algebra 56 (2008), 53-64. [3] M.-D. Choi, J. A. Holbrook, D.W. Kribs, and K. Zyczkowski, Higher-rank numerical ranges of unitary and normal matrices, Operators and Matrices 1 (2007), 409-426. [4] M.-D. Choi, D. W. Kribs, and K. Zyczkowski, Higher-rank numerical ranges and compression problems, Linear Algebra Appl. 418 (2006), 828-839. [5] M.-D. Choi, D. W. Kribs, and K. Zyczkowski, Quantum error correcting codes from the compression formalism, Rep. Math. Phys. 58 (2006), 77-91. [6] H. Gaaya, On the numerical radius of the truncated adjoint Shift. to appear. ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR 7 [7] H.L. Gau, C.K. Li, P.Y. Wu, Higher-rank numerical ranges and dilations, J. Operator Theory, in press. [8] U.Haagerup and P. dela Harpe, The numerical radius of a nilpotent operator on a Hilbert space, Proc. Amer. Math. Soc. 115(1992), 371 -- 379. [9] P. Lancaster and L. Rodman, Algebraic Riccati equations, Oxford Science Publications, The Clarendon Press, Oxford University Press, New York, 1995. [10] C.-K. Li, A simple proof of the elliptical range theorem, Proc. Amer. Math. Soc. 124 (1996), 1985-1986. [11] C.-K. Li, Y. T. Poon and N.-S. Sze, Condition for the higher rank numerical range to be non-empty, Linear and Multilinear Algebra, to appear. [12] C.-K. Li, Y. T. Poon and N.-S. Sze, Higher rank numerical ranges and low rank perturbations of quantum channels, preprint. http://arxiv.org/abs/0710.2898. [13] C.-K. Li, N.-S. Sze, Canonical forms, higher rank numerical ranges, totally isotropic sub- spaces, and matrix equations, Proc. Amer. Math. Soc. Volume 136, Number 9, September 2008, Pages 3013 -- 3023 . [14] C. Pop, On a result of Haagerup and de la Harpe. Rev. Roumaine Math. Pures Appl. 43 (1998), no. 9-10, 869 -- 871. [15] G. W. Stewart and J.-G. Sun, Matrix Perturbation Theory, Academic Press, New York, 1990. [16] H. Woerdeman, The higher rank numerical range is convex, Linear and Multilinear Algebra 56 (2008), 65-67. ‡.Institute Camille Jordan, Office 107 University of Lyon1, 43 Bd November 11, 1918, 69622-Villeurbanne, France. E-mail address: ‡[email protected]
1807.05625
1
1807
2018-07-15T22:41:07
Convex Bodies Associated to Tensor Norms
[ "math.FA" ]
We determine when a convex body in $\mathbb{R}^d$ is the closed unit ball of a reasonable crossnorm on $\mathbb{R}^{d_1}\otimes\cdots\otimes\mathbb{R}^{d_l},$ $d=d_1\cdots d_l.$ We call these convex bodies "tensorial bodies". We prove that, among them, the only ellipsoids are the closed unit balls of Hilbert tensor products of Euclidean spaces. It is also proved that linear isomorphisms on $\mathbb{R}^{d_1}\otimes\cdots \otimes \mathbb{R}^{d_l}$ preserving decomposable vectors map tensorial bodies into tensorial bodies. This leads us to define a Banach-Mazur type distance between them, and to prove that there exists a Banach-Mazur type compactum of tensorial bodies.
math.FA
math
CONVEX BODIES ASSOCIATED TO TENSOR NORMS MAITE FERNÁNDEZ-UNZUETA1 AND LUISA F. HIGUERAS-MONTAÑO2 Abstract. We determine when a convex body in Rd is the closed unit ball of a reasonable crossnorm on Rd1 ⊗ · · · ⊗ Rdl , d = d1 · · · dl. We call these convex bodies "tensorial bodies". We prove that, among them, the only ellipsoids are the closed unit balls of Hilbert tensor products of Euclidean spaces. It is also proved that linear isomorphisms on Rd1 ⊗ · · · ⊗ Rdl preserving decomposable vectors map tensorial bodies into tensorial bodies. This leads us to define a Banach-Mazur type distance between them, and to prove that there exists a Banach-Mazur type compactum of tensorial bodies. 1. Introduction Tensor products of finite dimensional spaces play a fundamental role in a wide range of problems in applications. They arise, among others, in quantum computing [15], in theoretical computer science [10], and in the use of tensor decompositions to extract and explain properties from data arrays (see [18] and the references therein). This fact has motivated the current research into their geometric, topologic and algebraic properties, as can be seen in [6, 12, 14, 19, 29]. On the other hand, there is a well developed theory of norms defined on tensor products of Banach spaces. This theory was established by A. Grothendieck [13]. It has had a great impact in the Geometry of Banach spaces, as can be traced in [7, 8, 9, 23, 25, 28]. Indeed, its impact extends even beyond Mathematical Analysis. By way of example, we refer to the survey [17] where applications of Grothendieck's theorem (usually called Grothendieck's inequality) to the design of polynomial time algorithms for computing approximate solutions of NP problems are detailed. In the other direction, we refer to [5] where results from theoretical computer science are ⊗πℓp3 fails to have used to prove that for some indices p1, p2, p3, the space ℓp1 non trivial cotype. The interested reader can consult [1, 24, 27] for further information about tensor products of Banach spaces and its applications. ⊗πℓp2 In the case of finite dimensions, the Minkowski functional enables the use of convex geometry to study finite dimensional Banach spaces (also known as Minkowski spaces) and vice versa. With it, a bijection between norms and 0-symmetric convex bodies in Rd is established. This result was originally due to H. Minkowski [22], and nowadays is a standard result (see [26, Remark 1.7.7] for a modern statement). Thus, in the context of tensors of finite dimensional spaces, a natural question to ask is if it is possible to determine the convex bodies that are the unit balls of tensor normed spaces, as well as 0-symmetric convex bodies are the unit balls of normed spaces. 2010 Mathematics Subject Classification. 46M05, 52A21, 46N10, 15A69. Key words and phrases. Convex body, Tensor norm, Minkowski space, Banach-Mazur distance, Tensor product of convex sets, Linear mappings on tensor spaces. 1 CONVEX BODIES ASSOCIATED TO TENSOR NORMS 2 The main result of this paper, Theorem 3.2, provides an affirmative answer to this question. This work, as well as [3], lies between the theory of tensor norms and convex geometry. In [3], G. Aubrun and S. Szarek establish connections between tensor norms on finite dimensions and convex geometry to estimate the volume of the set of separable mixed quantum states. We now briefly expose our results. Bringing together the theory of tensor norms and convex geometry, we immediately obtain that the the convex bodies Q ⊂ Rd that are the unit ball of a reasonable crossnorm defined on Rd = Rd1 ⊗ ··· ⊗ Rdl, d = d1 ··· dl, are those Q ⊂ Rd1 ⊗ ··· ⊗ Rdl for which there exist norms k·ki on Rdi such that (1.1) B ⊗π(Rdi ,k·ki) ⊆ Q ⊆ B ⊗ǫ(Rdi ,k·ki), where B denotes the closed unit ball of the projective and the injective tensor norms. In Proposition 3.1, we prove that (1.1) is equivalent to say that Q1 ⊗π ··· ⊗π Ql ⊆ Q ⊆ Q1 ⊗ǫ ··· ⊗ǫ Ql, (1.2) where for each i, Qi ⊂ Rdi is the closed unit ball of (cid:0)Rdi,k·ki(cid:1), and ⊗π,⊗ǫ are the projective and the injective tensor products of 0-symmetric convex bodies, defined by G. Aubrun and S. Szarek in [3, 4]. Our main result (Theorem 3.2) lies much deeper than Proposition 3.1. There, we establish the conditions on Q ⊂ Rd = Rd1 ⊗ ··· ⊗ Rdl that guarantee the existence of the convex sets Qi ⊂ Rdi in (1.2) and give an explicit description of them. Proposition 3.1 and Theorem 3.2 allow us to introduce "tensorial bodies": a 0- symmetric convex body Q ⊂ Rd1 ⊗ ··· ⊗ Rdl is a tensorial body in Rd1 ⊗ ··· ⊗ Rdl if there exist 0-symmetric convex bodies Qi ⊂ Rdi, i = 1, ..., l such that (1.2) holds (see Definition 3.3). Corollary 3.4 shows that the reasonable crossnorms on Rd1 ⊗···⊗ Rdl are the image of the tensorial bodies in Rd1 ⊗ ··· ⊗ Rdl, under the bijection given by the Minkowski functional. With it, we can go further with the study of this class of convex sets. We prove that the polar set of a tensorial body is a tensorial body and prove the stability of tensorial bodies by multiplying for positive scalars (Proposition 3.5). We also show that the convex bodies Qi in (1.2) are essentially unique (see Proposition 3.6). δBM i=1Rdi(cid:17)o , In Theorem 3.12 we prove that the subgroup of linear isomorphisms on Rd1⊗···⊗Rdl preserving decomposable vectors also preserve tensorial bodies. We denote this group i=1Rdi(cid:1), we start a geometric study of the by GL⊗(cid:0)⊗l i=1Rdi(cid:1). By means of GL⊗(cid:0)⊗l ⊗ (P, Q) := infnλ ≥ 1 : Q ⊆ T P ⊆ λQ for T ∈ GL⊗(cid:16)⊗l set of tensorial bodies, defining the following distance: where P, Q ⊂ Rd1 ⊗··· ⊗ Rdl are tensorial bodies. We call δBM the tensorial Banach- ⊗ Mazur distance. We use it to show that there is a Banach-Mazur type compactum of tensorial bodies in Rd1 ⊗ ··· ⊗ Rdl (Theorem 3.13). Finally, we apply the ideas developed through the paper to prove that the only ellipsoids that are also tensorial bodies in Rd1 ⊗ ··· ⊗ Rdl are the unit balls of the Hilbert tensor product of Euclidean spaces (see Theorem 4.2 and Corollary 4.3). CONVEX BODIES ASSOCIATED TO TENSOR NORMS 3 The paper is organized as follows: in Subection 1.1, we introduce the notation and basic results that we will use throughout the paper. In Section 2, we recall the main properties of the projective and the injective tensor product of 0-symmetric convex bodies. In Section 3, we prove Theorem 3.2 and establish the fundamental properties of tensorial bodies. There, we exhibit examples of tensorial bodies and show that not every 0-symmetric convex body is of this type. In Subsection 3.2, we establish the relation between GL⊗(cid:0)⊗l i=1Rdi(cid:1) and the set of tensorial bodies, and settle the fun- damental properties of δBM ⊗ . We finish this section by giving upper bounds for δBM ⊗ (Corollary 3.15). In Section 4, we characterize the ellipsoids in the class of tensorial bodies (Theorem 4.2 and Corollary 4.3). We like to point out that Theorems 3.2 and 3.12 remain true in Cd = ⊗l i=1Cdi , d = d1 ··· dl, when circled convex bodies (i.e. a convex body Q ⊂ Cd s.t. eiθQ = Q) are considered. As a consequence, it is possible to provide the corresponding notion i=1Cdi" as well as the definition of the tensorial Banach-Mazur of "tensorial body in ⊗l distance. Here, for the sake of transparency we will concentrate in the case of 0- symmetric convex bodies in real spaces. 1.1. Preliminaries. Throughout this paper, X, Y or Xi will denote Banach spaces. The closed unit ball of X will be denoted by BX and its dual space by X∗. We write L (X, Y ) to denote the space of bounded linear operators from X to Y. Let Vi, i = 1, . . . , l be vector spaces over the same field R or C. By ⊗l denote its tensor product, and by ⊗ we denote the canonical multilinear map: i=1Vi we ⊗ : V1 × ··· × Vl −→ ⊗l i=1Vi (cid:16)x1, . . . , xl(cid:17) → x1 ⊗ ··· ⊗ xl. i=1Xi is called a reasonable crossnorm if i=1Xi, α(cid:1)∗ and kx∗1 ⊗ ··· ⊗ x∗l k ≤ In the case of Banach spaces, a norm α (·) on the tensor product ⊗l (1) α(cid:0)x1 ⊗ ··· ⊗ xl(cid:1) ≤(cid:13)(cid:13)x1(cid:13)(cid:13)···(cid:13)(cid:13)xl(cid:13)(cid:13) for every xi ∈ Xi with i = 1, ..., l. (2) If x∗i ∈ X∗i for i = 1, ..., l then x∗1⊗···⊗x∗l ∈(cid:0)⊗l kx∗1k··· kx∗l k . If α (·) is a reasonable crossnorm on ⊗l i=1Xi, ⊗l i=1Xi, α(cid:1) , and X1 ⊗α ··· ⊗αXl its completion. (cid:0)⊗l For each u ∈ ⊗l i(cid:13)(cid:13)(cid:13) Xi=1 ǫ (u) := sup(cid:8)x∗1 ⊗ ··· ⊗ x∗l (u) : x∗i ∈ BX∗i , for i = 1, . . . , l(cid:9) . Both the projective and the injective norm are reasonable crossnorms on ⊗l i=1Xi. In- deed, these norms provide the next fundamental characterization of reasonable cross- norms: i=1Xi, the projective norm π and the injective norm ǫ are defined π (u) := inf( n Xi=1 (cid:13)(cid:13)x1 i(cid:13)(cid:13)···(cid:13)(cid:13)(cid:13) α,i=1Xi will denote the normed space n i) x1 i ⊗ ··· ⊗ xl by: and xl : u = A norm α (·) on ⊗l (1.3) i=1Xi is a reasonable crossnorm if and only if ǫ (u) ≤ α (u) ≤ π (u) for every u ∈ ⊗l i=1Xi. CONVEX BODIES ASSOCIATED TO TENSOR NORMS 4 The proof of this equivalence in the case of two normed spaces can be consulted in [25, Proposition 6.3]. For a deeper discussion about tensor norms we also refer to [7]. 1.1.1. Convex bodies in Euclidean spaces. Let E be a real Euclidean space with scalar product h·,·iE and Euclidean ball BE. A subset P ⊂ E is called a convex body if P is a compact convex set with nonempty interior. Every convex body P ⊂ E for which P = −P is called a 0-symmetric (or centrally symmetric) convex body. The set of 0-symmetric convex bodies in E is denoted by B (E) (resp. B (d) if E = Rd). If C is a nonempty subset of E, then its polar set is defined by C◦ := {y ∈ E : supx∈C hx, yiE ≤ 1} . The Minkowski functional (or gauge function) of P ∈ B (E) is defined as gP (x) := inf(cid:26)λ > 0 : 1 λ x ∈ P(cid:27) for x ∈ E. A fundamental result concerning 0-symmetric convex bodies is the bijection between norms defined on E and 0-symmetric convex bodies in E. This result, originally due to H. Minkowksi [22], will be used throughout the paper without making an explicit reference. We will use it in the following form: Theorem 1.1. Let E be a Euclidean space. If A ∈ B (E) , then kxkA := gA (x) for x ∈ E defines a norm k·kA on E for which A is the closed unit ball. Furthermore, for every x ∈ E we have kxkA◦ = kh·, xi : (E,k·kA) → Rk . This statement as well as the theory of convex bodies and convex geometry that will be used in this paper, can be found in [26]. 2. The projective and injective tensor products of 0-symmetric convex bodies To introduce the projective and the injective tensor products of 0-symmetric convex bodies, it is convenient to first recall two well known facts about tensor products of Banach spaces. The first one is that (2.1) BX1 ⊗π··· ⊗πXl = conv {x1 ⊗ ··· ⊗ xl : x1 ∈ BX1 , . . . , xl ∈ BXl}. The second one is the duality between the injective and projective tensor product of Banach spaces given by the canonical isometry: (2.2) X1 ⊗ǫ ··· ⊗ǫ Xl ֒→ (X∗1 ⊗π ··· ⊗π X∗l )∗ , which on finite dimensions is an isometric isomorphism (see [7, pp. 27, 46], respec- tively). Let Qi ⊂ Rdi, i = 1, . . . , l be 0-symmetric convex bodies with associated Minkowski functionals gQi , i = 1, . . . , l. By (2.1), conv{x1 ⊗ ··· ⊗ xl : xi ∈ Qi} is the closed unit i=1(cid:0)Rdi, gQi(cid:1) . This fact provides a natural way to ball of the projective norm on ⊗l define the projective tensor product of 0-symmetric convex bodies: the projective CONVEX BODIES ASSOCIATED TO TENSOR NORMS 5 tensor product of Q1, . . . , Ql is the 0-symmetric convex body in ⊗l by: i=1Rdi defined Q1 ⊗π ··· ⊗π Ql := convnx1 ⊗ ··· ⊗ xl ∈ ⊗l i=1Rdi : xi ∈ Qi, i = 1, . . . , lo . This definition was introduced by G. Aubrun and S. Szarek in [3]. There, the projec- tive tensor product of more general classes of convex sets is considered. Since conv(cid:8)x1 ⊗ ··· ⊗ xl ∈ ⊗l coincides with its closure. Then, i=1Rdi : xi ∈ Qi(cid:9) is compact (see Proposition 2.4), it (2.3) Q1 ⊗π ··· ⊗π Ql = B π,i=1(Rdi ,gQi (·)). ⊗l The duality between the injective and the projective tensor norms given in (2.2) gives rise to a notion of injective tensor product of 0-symmetric convex bodies. To be precise, we first fix the scalar products that will be used through the paper. 2 its associated norm and Euclidean ball respectively. Given d ∈ N, we will denote by h·,·i the standard scalar product on Rd, and by i=1Rdi will be the one associated to the Hilbert tensor The scalar product on ⊗l H,i=1Rdi , that is, h·,·iH will be the bilinear form determined by the relation k·k2 , Bd product ⊗l := Πl i=1(cid:10)xi, yi(cid:11) Dx1 ⊗ ··· ⊗ xl, y1 ⊗ ··· ⊗ ylEH (see [16, Section 2.5 ]). The closed unit ball of ⊗l and its norm by k · kH. In this way, given a 0-symmetric convex body Q ⊂ ⊗l its polar set acquires the form Q◦ =(cid:8)z ∈ ⊗l Now, if Qi ⊂ Rdi, i = 1, . . . , l are 0-symmetric convex bodies, the injective tensor product of Q1, . . . , Ql is the 0-symmetric conex body in ⊗l i=1Rdi defined as follows: Q1 ⊗ǫ ··· ⊗ǫ Ql := (Q◦1 ⊗π ··· ⊗π Q◦l )◦. H,i=1Rdi will be denoted by Bd1,...,dl , 2 i=1Rdi, i=1Rdi : supu∈Q hu, ziH ≤ 1(cid:9) . This definition appeared for the first time in the remarkable monograph [4, Subsec- tion 4.1.4] published in 2017. Later we will use this identity written in the following equivalent ways: Proposition 2.1. Let Qi ⊂ Rdi, i = 1, . . . , l be 0-symmetric convex bodies. Then, (1) (Q1 ⊗ǫ ··· ⊗ǫ Ql)◦ = Q◦1 ⊗π ··· ⊗π Q◦l . (2) (Q1 ⊗π ··· ⊗π Ql)◦ = Q◦1 ⊗ǫ ··· ⊗ǫ Q◦l . Due to the duality between the projective and the injective tensor norms (2.2), along with (2.3), we have that (2.4) Q1 ⊗ǫ ··· ⊗ǫ Ql = B ǫ,i=1(Rdi ,gQi (·)). ⊗l ∞. Proposition 2.2 below, together with (2.3) and ∞, d = d1 ··· dl, are the closed unit balls of 1 , Bd 1 and ℓd 2.1. The unit balls of ℓd (2.4) show that the convex bodies Bd π,i=1ℓdi ⊗l 1 and ⊗l ∞, respectively. In effect, let Bd jioji=1,...,di i = 1, ..., l, let nedi ǫ,i=1ℓdi p be the closed unit ball of ℓd p, d ∈ N and 1 ≤ p ≤ ∞. For each be the standard basis of Rdi. Then, the set of vectors CONVEX BODIES ASSOCIATED TO TENSOR NORMS 6 p and Bd1,...,dl H,i=1Rdi, and it can be identified with jlo is an orthonormal basis in ⊗l ned1 j1 ⊗ ··· ⊗ edl the standard basis of Rd, d = d1 ··· dl. Consequently, for each 1 ≤ p ≤ ∞, the sets := i=1Rdi : Xj1,...,jl(cid:12)(cid:12)(cid:12)Dz, ed1 z ∈ ⊗l  :=(cid:26)z ∈ ⊗l j1 ⊗ ··· ⊗ edl j1 ⊗ ··· ⊗ edl jlEH(cid:12)(cid:12)(cid:12) i=1Rdi : max for p 6= ∞ are naturally identified with the closed unit balls of ℓd 1 ≤ p ≤ ∞. Proposition 2.2. Let d ∈ N. For every factorization of d in natural numbers d = d1 ··· dl, p. Thus, Bd ≤ 1  jlEH(cid:12)(cid:12)(cid:12) ≤ 1(cid:27) j1,...,jl(cid:12)(cid:12)(cid:12)Dz, ed1 p = Bd1,...,dl Bd1,...,dl ∞ p for p Bd 1 = Bd1 1 ⊗π ··· ⊗π Bdl 1 and Bd ∞ = Bd1 ∞ ⊗ǫ ··· ⊗ǫ Bdl ∞. The previous proposition is a well known result, see for instance [25, Excercise 2.6] or [4, pp. 83]. i=1ℓdi i=1ℓdi p . H,i=1Rdi. p is the p . In this case it is not In Subection 3.1, we will treat the case 1 < p < ∞. We will see that Bd closed unit ball associated to a reasonable crossnorm on ⊗l the projective nor the injective tensor norm on ⊗l We finish this section stating without proof two results that will be used throughout the paper. Proposition 2.3 is a well known result (for a proof see [6, Proposition 4.2]). Proposition 2.4 is a direct consequence of the continuity of the canonical multilinear map ⊗ : Rd1 × ··· × Rdl → ⊗l Proposition 2.3. The set of decomposable vectors(cid:8)x1 ⊗ ··· ⊗ xl ∈ ⊗l i=1Rdi : xi ∈ Rdi(cid:9) is a closed subset of ⊗l Proposition 2.4. If Ai ⊆ Rdi, i = 1, ..., l are compact sets then ⊗ (A1, . . . , Al) := (cid:8)x1 ⊗ ··· ⊗ xl ∈ ⊗l i=1Rdi : xi ∈ Ai(cid:9) is a compact subset of ⊗l In this section we characterize the convex bodies in ⊗l i=1Rdi that are the closed unit balls of reasonable crossnorms. They will be called tensorial bodies (Definition 3.3). A main tool to study them is the group of linear isomorphisms that preserve decomposable vectors. With it, we will introduce a Banach-Mazur type distance between tensorial bodies, and prove that there is a Banach-Mazur type compactum associated to them (see Subsection 3.2). 3. tensorial bodies H,i=1Rdi. H,i=1Rdi . Recall that we have already fixed the scalar product h·,·iH on ⊗l i=1Rdi and that gQ denotes the Minkowski functional of a 0-symmetric convex body Q. Whit them, we have: Proposition 3.1. Let Q ⊂ ⊗l convex bodies. Then, gQ (·) is a reasonable crossnorm on ⊗l only if i=1Rdi and let Qi ⊂ Rdi, i = 1, . . . , l be 0-symmetric i=1(cid:0)Rdi, gQi (·)(cid:1) if and (3.1) Q1 ⊗π ··· ⊗π Ql ⊆ Q ⊆ Q1 ⊗ǫ ··· ⊗ǫ Ql. CONVEX BODIES ASSOCIATED TO TENSOR NORMS 7 i=1Rdi we have: In this case, for every decomposable vector x1 ⊗ ··· ⊗ xl ∈ ⊗l gQ(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = gQ1(cid:0)x1(cid:1)··· gQl(cid:16)xl(cid:17) , gQ◦(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = gQ◦1(cid:0)x1(cid:1)··· gQ◦l (cid:16)xl(cid:17) . Proof. Let Qi ⊂ Rdi , i = 1, . . . , l, be 0-symmetric convex bodies. Then, (2.3) and π,i=1(cid:0)Rdi , gQi(cid:1) , and (2.4) tell us that Q1 ⊗π ··· ⊗π Ql is the closed unit ball of ⊗l ǫ,i=1(cid:0)Rdi , gQi(cid:1) . Now, the proof of the first Q1 ⊗ǫ ··· ⊗ǫ Ql is the closed unit ball of ⊗l part follows from the characterization of a reasonable crossnorm (1.3). The second part follows using the two properties that define being a reasonable crossnorm. (cid:3) This proposition can be understood as the definition of a reasonable crossnorm written in terms of convex bodies. It determines when a 0-symmetric convex body in ⊗l i=1Rdi is the unit ball of a reasonable crossnorm when the norms on each Rdi are fixed (gQi). Our next result goes further: it determines when a 0-symmetric convex body in ⊗l i=1Rdi is the unit ball of a reasonable crossnorm, with respect to some norms (not determined a priori) on the spaces Rdi. For every non-zero decomposable vector a ∈ ⊗l i=1Rdi and every 0-symmetric convex i=1Rdi . If a = a1 ⊗ ··· ⊗ al, consider the 0-symmetric convex bodies in body Q ⊂ ⊗l Rdi, i = 1, . . . , l, defined as: (3.2) Qa1,...,al i :=nxi ∈ Rdi : a1 ⊗ ··· ⊗ ai−1 ⊗ xi ⊗ ai+1 ⊗ ··· ⊗ al ∈ Qo . i=1Rdi be a 0-symmetric convex body. Then, there exist Theorem 3.2. Let Q ⊂ ⊗l norms k·ki on Rdi, i = 1, ..., l, such that Q is the closed unit ball of a reasonable i=1(cid:0)Rdi ,k·ki(cid:1) if and only if for an arbitrary decomposable vector a1 ⊗ crossnorm on ⊗l ··· ⊗ al ∈ ∂Q it holds: (3.3) In such a situation, Qa1,...,al ⊗ǫ ··· ⊗ǫ Qa1,...,al Qa1,...,al ⊆ Q ⊆ Qa1,...,al B(Rdi ,k·ki). 1 1 . l l ⊗π ··· ⊗π Qa1,...,al =(cid:13)(cid:13)ai(cid:13)(cid:13)i i 1, and i=1Rdi we have: Proof. Suppose that Q is the closed unit ball of a reasonable crossnorm α (·) on i=1(cid:0)Rdi,k·ki(cid:1) . ⊗l Clearly gQ (·) = α (·) and for each x1 ⊗ ··· ⊗ xl ∈ ⊗l (3.4) =(cid:13)(cid:13)(cid:10)·, x1(cid:11)(cid:13)(cid:13)···(cid:13)(cid:13)(cid:13)D·, xlE(cid:13)(cid:13)(cid:13) and (cid:13)(cid:13)(cid:13)D·, x1 ⊗ ··· ⊗ xlEH(cid:13)(cid:13)(cid:13) gQ(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) =(cid:13)(cid:13)x1(cid:13)(cid:13)1 ···(cid:13)(cid:13)(cid:13) xl(cid:13)(cid:13)(cid:13)l where (cid:10)·, xi(cid:11) is a linear funtional on (cid:0)Rdi ,k · ki(cid:1) , i = 1, . . . , l. Now, if we fix an arbitrary a1⊗···⊗al ∈ ∂Q, then gQ(cid:0)a1 ⊗ ··· ⊗ al(cid:1) =(cid:13)(cid:13)a1(cid:13)(cid:13)1 ···(cid:13)(cid:13)al(cid:13)(cid:13)l = gQ(cid:16)a1 ⊗ ··· ai−1 ⊗ xi ⊗ ai+1 ⊗ ··· ⊗ al(cid:17) =(cid:13)(cid:13)a1(cid:13)(cid:13)1 ···(cid:13)(cid:13)ai−1(cid:13)(cid:13)i−1(cid:13)(cid:13)xi(cid:13)(cid:13)i(cid:13)(cid:13)ai+1(cid:13)(cid:13)i+1 ···(cid:13)(cid:13)(cid:13) al(cid:13)(cid:13)(cid:13)l kaiki (cid:13)(cid:13)xi(cid:13)(cid:13)i for i = (cid:17)◦(cid:0)xi(cid:1) = (cid:0)xi(cid:1) = 1 B(Rdi ,k·ki). Since the latter is equivalent to g Thus, from the definition of Qa1,...,al 1 . . . , l and Qa1,...,al kaiki (cid:13)(cid:13)xi(cid:13)(cid:13)i , we obtain g (cid:16)Qa1,...,al Qa1,...,al = 1 , i . i i i =(cid:13)(cid:13)ai(cid:13)(cid:13)i CONVEX BODIES ASSOCIATED TO TENSOR NORMS 8 and gQ(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = g (cid:13)(cid:13)ai(cid:13)(cid:13)i(cid:13)(cid:13)(cid:10)·, xi(cid:11)(cid:13)(cid:13) , i = 1 . . . , l, then from (3.4) and the previous equalities we have: =(cid:13)(cid:13)(cid:10)·, x1(cid:11)(cid:13)(cid:13) ···(cid:13)(cid:13)(cid:13)D·, xlE(cid:13)(cid:13)(cid:13) (cid:0)x1(cid:1)··· g gQ◦(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) =(cid:13)(cid:13)(cid:13)D·, x1 ⊗ ··· ⊗ xlEH(cid:13)(cid:13)(cid:13) (cid:17)◦(cid:0)x1(cid:1)··· g (cid:17)◦(cid:16)xl(cid:17) . (cid:16)xl(cid:17) (cid:16)Qa1,...,al (cid:16)Qa1,...,al Therefore, by Proposition 3.1, (3.3) holds. Qa1,...,al Qa1,...,al 1 l = g 1 l To prove the converse, suppose that Q satisfies (3.3) for a1⊗···⊗al ∈ ∂Q, then from (cid:19) . Qa1 ,...,al Proposition 3.1, we conclude that gQ is a reasonable crossnorm on ⊗l This completes the proof. i=1(cid:18)Rdi, g (cid:3) i Now, we introduce the formal notion of a tensorial body: Definition 3.3. A 0-symmetric convex body Q ⊂ ⊗l in ⊗l i=1Rdi is called a tensorial body i=1Rdi if there exist 0-symmetric convex bodies Qi ⊂ Rdi , i = 1, ..., l such that Q1 ⊗π ··· ⊗π Ql ⊆ Q ⊆ Q1 ⊗ǫ ··· ⊗ǫ Ql. If Q satisfies the inclusions in Definition 3.3, we will say that Q is a tensorial i=1Rdi is denoted body with respect to Q1, . . . , Ql. The set of tensorial bodies in ⊗l by B⊗(cid:0)⊗l BQ1,...,Ql(cid:0)⊗l In the next corollary, we summarize the relation between tensorial bodies in ⊗l i=1Rdi(cid:1) . The set of tensorial bodies with respect to Q1, ..., Ql is denoted by i=1Rdi(cid:1) . i=1Rdi and reasonable crossnorms. We omit its proof, since it follows directly from Proposi- tion 3.1 and Theorem 3.2. Corollary 3.4. Let Q ⊂ ⊗l equivalent: i=1Rdi be a 0-symmetric convex body. The following are (1) Q is a tensorial body in ⊗l (2) Q satisfies (3.3) for any a1 ⊗ ··· ⊗ al ∈ ∂Q. (3) There exist norms k·ki on Rdi , i = 1, ..., l, such that gQ is a reasonable cross- i=1Rdi. i=1Rdi is a tensorial body with respect to Q1, . . . , Ql then Q1 ⊗π ··· ⊗π Ql ⊆ Q ⊆ Q1 ⊗ǫ ··· ⊗ǫ Ql. Thus, (Q1 ⊗ǫ ··· ⊗ǫ Ql)◦ ⊆ Q◦ ⊆ (Q1 ⊗π ··· ⊗π Ql)◦ . By Proposition 2.1, this im- plies that Q◦1 ⊗π ··· ⊗π Q◦l ⊆ Q◦ ⊆ Q◦1 ⊗ǫ ··· ⊗ǫ Q◦l which is equivalent to Q◦ ∈ BQ◦1,...,Q◦l (cid:0)⊗l i=1Rdi(cid:1) . kaiki k·ki for i = 1, . . . , l. i In this case, g norm on ⊗l Qa1,...,al i=1(cid:0)Rdi ,k·ki(cid:1) . (·) = 1 Proposition 3.5. Let Q ⊂ ⊗l are tensorial bodies. Indeed, if Q ∈ BQ1,...,Ql(cid:0)⊗l bodies Qi ⊂ Rdi, i = 1, ..., l, then (1) Q◦ ∈ BQ◦1,...,Q◦l (cid:0)⊗l i=1Rdi(cid:1) . (2) λQ ∈ BQ1,...,(λQk),...,Ql(cid:0)⊗l i=1Rdi(cid:1) . Proof. (1). If Q ⊂ ⊗l i=1Rdi be a tensorial body. Then Q◦ and λQ, λ > 0, i=1Rdi(cid:1) for some 0-symmetric convex CONVEX BODIES ASSOCIATED TO TENSOR NORMS 9 (2). We assume, w.l.o.g., that k = 1. To prove this part, it is enough to observe that, by definition, for each real number λ > 0, we have λ(Q1 ⊗π ··· ⊗π Ql) = (λQ1) ⊗π ··· ⊗π Ql and λQ1 ⊗ǫ ··· ⊗ǫ Ql = λ (Q◦1 ⊗π ··· ⊗π Q◦l )◦ =(cid:0)λ−1(Q◦1 ⊗π ··· ⊗π Q◦l )(cid:1)◦ = ((λQ1)◦ ⊗π ··· ⊗π Q◦l )◦ = (λQ1) ⊗ǫ ··· ⊗ǫ Ql. i=1Rdi(cid:1) , if Q ∈ BQ1,...,Ql(cid:0)⊗l i=1Rdi(cid:1) . (cid:3) i=1Rdi is a tensorial body with respect to an essentially unique From this, it follows that λQ ∈ B(λQ1),...,Ql(cid:0)⊗l A tensorial body in ⊗l l-tuple of convex bodies. More precisely: Proposition 3.6. Let Pi, Qi ⊂ Rdi, i = 1 . . . , l be 0-symmetric convex bodies. If Q ∈ BP1,...,Pl(cid:16)⊗l i=1Rdi(cid:17) ∩ BQ1,...,Ql(cid:16)⊗l i=1Rdi(cid:17) , then there exist real numbers λi > 0, i = 1, . . . , l such that λ1 ··· λl = 1 and Pi = λiQi for i = 1, ..., l. Proof. Let gQ, gQi and gPi be the Minkowski functionals associated to Q, Qi and Pi respectively. If Q is a tensorial body with respect to Pi, i = 1, . . . , l, and with respect to Qi, i = 1, . . . , l, then Proposition 3.1 implies that: gP1(cid:0)x1(cid:1)··· gPl(cid:16)xl(cid:17) = gQ(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = gQ1(cid:0)x1(cid:1)··· gQl(cid:16)xl(cid:17) . Therefore, if we fix a1 ⊗ ··· ⊗ al ∈ ∂Q, then gQ1(cid:0)a1(cid:1)··· gQl(cid:16)al(cid:17) = gP1(cid:0)a1(cid:1)··· gPl(cid:16)al(cid:17) = 1. Analogously, for a1 ⊗ ··· ⊗ ai−1 ⊗ xi ⊗ ai+1 ⊗ ··· ⊗ al, i = 1, ..., l, we have: gQ1(cid:0)a1(cid:1)··· gQi−1(cid:0)ai−1(cid:1) gQi(cid:0)xi(cid:1) gQi+1(cid:0)ai+1(cid:1)··· gQl(cid:16)al(cid:17) = gP1(cid:0)a1(cid:1)··· gPi−1(cid:0)ai−1(cid:1) gPi(cid:0)xi(cid:1) gPi+1(cid:0)ai+1(cid:1)··· gPl(cid:16)al(cid:17) . Now, if we multiply both sides of the above equation by gQi(cid:0)ai(cid:1) gPi(cid:0)ai(cid:1) , we obtain that gPi(cid:0)ai(cid:1) gQi(cid:0)xi(cid:1) = gQi(cid:0)ai(cid:1) gPi(cid:0)xi(cid:1) , which is equivalent to gPi(cid:0)xi(cid:1) = gQi(ai) gQi (ai) gQi(cid:0)xi(cid:1) . Thus, if λi := gPi (ai) , i = 1, ..., l then we have proved that λ1 ··· λl = 1 and Pi = λiQi, as required. gPi(ai) (cid:3) d1 1 ,...,λe In order to simplify the arguments, we will choose the convex bodies defined (3.2) i=1Rdi , Qi will denote . That is, in a specific way: for every 0-symmetric convex body Q ⊂ ⊗l the convex bodies generated by ed1 gQ(cid:16)e 1 ⊗ ··· ⊗ (cid:16)λedl 1 (cid:17) , λ = d1 1 ⊗···⊗e for i = 1, . . . , l. Qi := Q Proposition 3.7. If Qn, n ∈ N, are tensorial bodies in ⊗l i=1Rdi such that gQn con- verges uniformly on compact sets to gQ, for some 0-symmetric convex body Q, then Q is a tensorial body in ⊗l and g(Qi n)◦ converge uniformly on compact sets to gQi and g(Qi)◦, respectively. i=1Rdi. In this case, gQi 1 (cid:17) dl 1 e i dl l n CONVEX BODIES ASSOCIATED TO TENSOR NORMS 10 Proof. Since Qn, n ∈ N, are tensorial bodies, then (2) of Corollary 3.4 implies that Q1 n ⊗π ··· ⊗π Ql Suppose that we already proved the uniform convergence (on compact sets) of gQi to gQi. From this, it follows that g(Qi n)◦ converges uniformly on compact sets to g(Qi)◦. Thus, we get: n ⊗ǫ ··· ⊗ǫ Ql n ⊆ Qn ⊆ Q1 n, for n ∈ N. n gQ(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = limn→∞gQn(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = limn→∞gQ1 n(cid:0)x1(cid:1)··· gQl n(cid:16)xl(cid:17) = gQ1(cid:0)x1(cid:1)··· gQl(cid:16)xl(cid:17) . Similarly, since the uniform convergence of gQn to gQ implies the convergence gQ◦n to gQ◦, we have gQ◦(cid:0)x1 ⊗ ··· ⊗ xl(cid:1) = g(Q1)◦(cid:0)x1(cid:1)··· g(Ql)◦(cid:0)xl(cid:1) . Therefore, from Propo- sition 3.1, Q is a tensorial body w.r.t. Qi, i = 1, . . . , l. Now, we turn to prove that for each i = 1, . . . , l, gQi converges pointwise to gQi . Then, by [26, Theorem 1.8.12], we know that this implies the uniform convergence. From the convergence of gQn to gQ, and the definition of Ql, Ql n, it follows directly converges pointwise to gQl. For the case i = 1, . . . , l − 1, it is enough to that gQl observe that 1 (cid:17) , 1 ⊗ ··· ⊗ edl for xi ∈ Rdi. Thus, from the definiton of Qi and the convergence of gQn to gQ, we know that gQi 1 (cid:17) 1 ⊗ ··· ⊗ edl 1 ⊗ xi ⊗ edi+1 1 ⊗ ··· ⊗ edi−1 converges pointwise gQi. n(cid:0)xi(cid:1) = gQn(cid:16)ed1 gQn(cid:16)ed1 gQi (cid:3) 1 n n n 3.1. Examples of tensorial bodies. 3.1.1. The trivial case. Every 0-symmetric convex body Q ⊂ R ⊗ Rd is a tensorial body in R ⊗ Rd : Proposition 3.8. Let Q ⊂ R⊗Rd be a 0-symmetric convex body. Then Q = [−1, 1]⊗π Q where [−1, 1] = {λ ∈ R : −1 ≤ λ ≤ 1} and Q := {x ∈ Rd : 1 ⊗ x ∈ Q}. i=1 λi ⊗ xi = 1 ⊗(cid:16)PN i=1 λixi(cid:17) . Thus, u ∈ Q Proof. Let u ∈ R ⊗ Rd, then u = PN i=1 λixi ∈ Q. Therefore, from the definition of if and only if u = 1 ⊗ u with u = PN [−1, 1] ⊗π Q we obtain the desired result. R⊗Rd is the closed unit ball associated to the projective tensor norm on R⊗(cid:16)Rd, g Q(cid:17). It is also worth to notice that on R⊗Rd, the projective and the injective tensor product of 0-symmetric convex bodies are equal. The proposition above and (2.3) show that every 0-symmetric convex body Q in (cid:3) 3.1.2. The closed unit balls of ℓd p. Proposition 3.9. Let d ∈ N. For every factorization of d in natural numbers d = i=1Rdi. It d1 ··· dl and for every 1 ≤ p ≤ ∞, Bd holds that is a tensorial body in ⊗l p = Bd1,...,dl p Bd1 p ⊗π ··· ⊗π Bdl p ⊆ Bd p = Bd1,...,dl p ⊆ Bd1 p ⊗ǫ ··· ⊗ǫ Bdl In the cases where 1 < p < ∞, di ≥ 2, i = 1, . . . , l, Bd1,...,dl the injective tensor product of Bdi p , i = 1, . . . , l. p p , 1 ≤ p ≤ ∞. is not the projective nor CONVEX BODIES ASSOCIATED TO TENSOR NORMS 11 Proof. We will use the notation fixed in Example 2.1. The cases p = 1,∞ were already proved in Proposition 2.2. We will give the proof for 1 < p < ∞. Let xi ∈ Rdi, i = 1, . . . , l then: 1 p j1 ⊗ ··· ⊗ edl p(cid:17) jlEH(cid:12)(cid:12)(cid:12) (cid:17)◦ = Bd1,...,dl p∗ we have . Therefore, by Proposition 3.1, . B g d1 ,...,dl p (cid:16)x1 ⊗ ··· ⊗ xl(cid:17) =(cid:16)Xj1,...,jl(cid:12)(cid:12)(cid:12)Dx1 ⊗ ··· ⊗ xl, ed1 Thus, from the last equality for p∗ and the relation (cid:16)Bd1,...,dl that g(cid:16)B Bd1,...,dl =(cid:13)(cid:13)x1(cid:13)(cid:13)p ···(cid:13)(cid:13)(cid:13) is a tensorial body w.r.t. Bdi xl(cid:13)(cid:13)(cid:13)p (cid:17)◦(cid:0)x1 ⊗ ··· ⊗ xl(cid:1) = (cid:13)(cid:13)x1(cid:13)(cid:13)p∗ ···(cid:13)(cid:13)xl(cid:13)(cid:13)p∗ p ⊗ǫ ···⊗ǫ Bdl π,i=1ℓdi p ⊗ǫ ···⊗ǫ Bdl p , i = 1 . . . , l. B1,...,dl di ≥ 2. d1 ,...,dl p p p p To prove the other statement, first observe that if di = 1, i = 1, . . . , l − 1 then p . To avoid this case, we assume that each p = B1 = B1 1 p 1 , ed1 1 + a2ed1 Let E ⊂ ⊗l 1 ⊗ ··· ⊗ edl 1 ⊗ ··· ⊗ edl 2 ⊗ ··· ⊗ edl p}. Hence, for each u = a1ed1 r . Therefore, Bd1,...,dl p be the vector space generated byned1 2 o . 2 ⊗ ··· ⊗ edl Then, from [2, Theorem 1.3,], it follows that E is isometrically isomorphic to ℓ2 r for 1 r = min{1, l 2 ∈ E we have π (u) = (a1r + a2r) p Proposition 2.1, Bd1,...,dl we must have Bd1,...,dl p ⊗π ··· ⊗π Bdl p . p for some 1 < p < ∞. Then, from p∗. Since the latter equality is not possible, (cid:3) Now, suppose that Bd1,...,dl = Bd1 p∗ ⊗π ···⊗π Bdl = Bd1 p ⊗ǫ ··· ⊗ǫ Bdl 6= Bd1 p ⊗ǫ ··· ⊗ǫ Bdl p for all 1 < p < ∞. Proposition 3.9 together with Corollary 3.4 imply that Bd i=1ℓdi closed unit ball associated to a reasonable crossnorm on ⊗l d = d1 ··· dl. 3.1.3. A convex body in Rmn = Rm ⊗ Rn which is not a tensorial body in Rm ⊗ Rn. Let m, n ∈ N, m, n ≥ 2. Let p , 1 ≤ p ≤ ∞, is the p for any factorization 6= Bd1 p∗ p m,n z = zijem i ⊗ en E = Xi,j=1  1 ⊗ √3en 1 ,√3en =(cid:8)x ∈ Rm : x ⊗ √3en em 1 = λ1E em 1 and em 1 1 j : z112 3 + zmn2 2 m ⊗ √2en 1 ∈ E(cid:9) , E em 1 m,√2en . + X(i,j)6=(1,1),(m,n) zij2 ≤ 1  =(cid:8)x ∈ Rm : x ⊗ √2en n Then E ∈ B(Rm ⊗ Rn) \ B⊗(Rm ⊗ Rn). To verify this, consider the convex bodies generated by em n, according to the relation (3.2): E n ∈ E(cid:9) . We will proceed by contradiction. Suppose that E is a tensorial body in Rm ⊗ Rn, then Corollary 3.4 and Proposition 3.6 imply that there exists λ1 > 0 such that E 1 ,√3en em 1 m,√2en n 1 . However, for every x = (x1, . . . , xm) ∈ Rm we have: (x) = gE(cid:16)x ⊗ (x) = gE(cid:16)x ⊗ √3en 1(cid:17) =rx12 + 3Xm i=2 xi2, √2en n(cid:17) =rxm2 + 2Xm−1 i=1 xi2. ,√3en 1 g E em 1 1 g m ,√2en E em n 1 CONVEX BODIES ASSOCIATED TO TENSOR NORMS 12 1 n 1 1 ,√3en em 1 6= λE em for all λ > 0. This is a contradiction, hence Q is not a Analogous examples E so that E ∈ B(⊗l i=1Rdi when l ≥ 2. m,√2en Thus E tensorial body in Rm ⊗ Rn. in ⊗l Remark 3.10. The previous example, in contrast with Example 3.1.1 (the trivial case), shows that if d = mn is not a trivial factorization of d (i.e. if m 6= 1 or n 6= 1) then B⊗(Rm ⊗ Rn) ( B(Rm ⊗ Rn). As a consequence being a tensorial body depends on the tensor decomposition defined on Rd. i=1Rdi), can be constructed i=1Rdi) \ B⊗(⊗l i=1Rdi i=1Rdi preserving decomposable vectors. To shorten i=1Rdi preserves decomposable vectors if T (cid:0)x1 ⊗ ··· ⊗ xl(cid:1) is a decomposable vec- i=1Rdi(cid:1) we denote the set of linear 3.2. Linear isomorphisms preserving tensorial bodies. A linear map T : ⊗l → ⊗l tor for every xi ∈ Rdi, i = 1, . . . , l. By GL⊗(cid:0)⊗l i=1Rdi → ⊗l isomorphisms T : ⊗l notation we usually write GL⊗. Linear mappings preserving decomposable vectors have been deeply studied. For an account on this topic as well as for the fundamentals about it, we refer the reader to [20, 21, 30, 31]. In [20, Corollary 2.14], it is proved that if di ≥ 2, i = 1, . . . , l, then for each element T ∈ GL⊗(cid:0)⊗l i=1Rdi(cid:1) there exists a permutation σ on {1, ..., l} and linear isomorphisms Ti : Rdσ(i) → Rdi , i = 1, ..., l such that for every x1 ⊗···⊗ xl ∈ ⊗l i=1Rdi , (3.5) T(cid:16)x1 ⊗ ··· ⊗ xl(cid:17) = T1(cid:16)xσ(1)(cid:17) ⊗ ··· ⊗ Tl(cid:16)xσ(l)(cid:17) . H,i=1Rdi(cid:17) . H,i=1Rdi (see Proposition 2.3), we can easily obtain the next result. Using this characterization and the fact that the set of decomposable vectors is closed in ⊗l Proposition 3.11. Let di ≥ 2, i = 1, . . . , l be natural numbers. Then GL⊗(cid:0)⊗l i=1Rdi(cid:1) is a closed subgroup of GL(cid:16)⊗l Theorem 3.12. (GL⊗ preserves tensorial bodies). Assume di ≥ 2 for i = 1, ..., l. If T ∈ GL⊗(cid:0)⊗l Proof. Suppose that Q is a tensorial body in ⊗l 0-symmetric convex bodies Qi ⊂ Rdi, i = 1, . . . , l. On the other hand, let T be an element in GL⊗(cid:0)⊗l be as in (3.5). Then, by the definition of ⊗π, we have: (3.6) i=1Rdi(cid:1) and let Ti, i = 1, ..., l, i=1Rdi(cid:1) , then T Q ∈ B⊗(cid:0)⊗l i=1Rdi(cid:1) and Q ∈ B⊗(cid:0)⊗l i=1Rdi, then (3.1) holds for suitable i=1Rdi(cid:1) . T (Q1 ⊗π ··· ⊗π Ql) = T1Qσ(1) ⊗π ··· ⊗π TlQσ(l). Similarly, T (Q1 ⊗ǫ ··· ⊗ǫ Ql) =(cid:0)(T t)−1 (Q◦1 ⊗π ··· ⊗π Q◦l )(cid:1)◦ =(cid:16)(T t 1)−1Q◦σ(1) ⊗π ··· ⊗π (T t = T1Qσ(1) ⊗ǫ ··· ⊗ǫ TlQσ(l). l )−1Q◦σ(l)(cid:17)◦ Therefore, T1Qσ(1) ⊗π ··· ⊗π TlQσ(l) ⊆ T Q ⊆ T1Qσ(1) ⊗ǫ ··· ⊗ǫ TlQσ(l). This proves that T Q is a tensorial body in ⊗l i=1Rdi. (cid:3) CONVEX BODIES ASSOCIATED TO TENSOR NORMS 13 A Banach-Mazur type distance. From now on, we will assume that each space Rdi, i = 1, . . . , l has dimension di ≥ 2. Using Theorem 3.12 we are able to define a distance δBM i=1Rdi , which is the analogue, for tensorial bodies, of the Banach-Mazur distance. ⊗ between tensorial bodies in ⊗l Recall that the Banach-Mazur distance between isomorphic Banach spaces X and Y is defined as: δBM (X, Y ) := inf(cid:8)kTk(cid:13)(cid:13)T −1(cid:13)(cid:13) : T ∈ L (X, Y ) and T −1 ∈ L (Y, X)(cid:9) . Between 0-symmetric convex bodies in a Euclidean space E, it is defined as: δBM (P, Q) := inf {λ ≥ 1 : T : E → E is a bijective linear map and Q ⊆ T P ⊆ λQ} . A complete exposition of the Banah-Mazur distance and its properties can be found in [28]. Let P, Q be tensorial bodies in ⊗l distance δBM (3.7) ⊗ (P, Q) as follows: ⊗ (P, Q) := infnλ ≥ 1 : Q ⊆ T P ⊆ λQ, for T ∈ GL⊗(cid:16)⊗l δBM i=1Rdi(cid:17)o . i=1Rdi. We define the tensorial Banach-Mazur It is well defined, since for every P, Q ∈ B⊗(cid:0)⊗l r1, r2 > 0 such that Q ⊆ r1P ⊆ r2Q. It holds that (3.8) δBM (P, Q) ≤ δBM ⊗ (P, Q) for P, Q ∈ B⊗(cid:16)⊗l i=1Rdi(cid:17) . i=1Rdi(cid:1) there exist real numbers Using Proposition 3.11 and Theorem 3.12, it can be directly proved that for each pair P, Q ⊂ ⊗l i=1Rdi of tensorial bodies, the infimum in (3.7) attains its value at some λ > 0 and some T ∈ GL⊗(cid:0)⊗l equivalence relation: For every P, Q ∈ B⊗(cid:0)⊗l We denote BM⊗(cid:0)⊗l i=1Rdi(cid:1) . Indeed, it is possible to define the following i=1Rdi(cid:1) , P ∼ Q if and only if δBM i=1Rdi(cid:1) the set of equivalence classes of tensorial bodies deter- mined by this relation. Elementary arguments show that log δBM is a metric on this set. Moreover, this metric gives rise to a Banach-Mazur type compactum of tensorial bodies: ⊗ (P, Q) = 1. ⊗ a compact metric space. Theorem 3.13. (The compactum of tensorial bodies) (cid:0)BM⊗(cid:0)⊗l Proof. The proof is essentially a standard argument of compactness. Given a tensorial body P ⊂ ⊗l Let {[Pn]} be a sequence in BM⊗(cid:0)⊗l duced in Proposition 3.7. By Corollary 3.4, P 1 Thus, from [11, Proposition 2.4], it follows that n be the convex sets intro- n ⊆ Pn ⊆ P 1 n ⊗ǫ ··· ⊗ǫ P l n. i=1Rdi, [P ] denotes its associate equivalence class. i=1Rdi(cid:1) and let P i n ⊗π ··· ⊗π P l i=1Rdi(cid:1) , log δBM ⊗ (cid:1) is (3.9) P 1 n ⊗π ··· ⊗π P l n ⊆ Pn ⊆ d dl P 1 n ⊗π ··· ⊗π P l n, for every n ∈ N. On the other hand, from a general well known fact, for every n, i = 1, ..., l there exists a linear isomorphism Ti,n : Rdi → Rdi, such that Bdi P i 1 ⊆ n ⊆ diBdi Ti,nP i 1 . Hence, applying T1,n ⊗···⊗ Tl,n to (3.9) together with (3.6), we have 1 ⊗π ··· ⊗π Bdl Bd1 1 ⊆ (T1,n ⊗ ··· ⊗ Tl,n) Pn ⊆ 1 ⊗π ··· ⊗π Bdl Bd1 1 . d2 dl CONVEX BODIES ASSOCIATED TO TENSOR NORMS 14 Now, for each n ∈ N denote by Qn the tensorial body (T1,n ⊗ ··· ⊗ Tl,n) Pn. By the Arzela-Ascoli theorem, there is a subsequence ngQnko converging uniformly (on compact sets of ⊗l Proposition 3.7, Q is a tensorial body in ⊗l i=1Rdi) to gQ for some 0-symmetric convex body Q. Hence, by What is left is to show that [Pnk ] converges to [Q]. To prove this, notice that the uni- i=1Rdi . form convergence of gQnk i=1Rdi , gQ(cid:1) is such that limk→∞ kIkk(cid:13)(cid:13)I−1 (cid:0)⊗l ⊗ (Pnk , Q) = δBM to gQ implies that the indentity map Ik :(cid:16)⊗l i=1Rdi , gQnk(cid:17) → k (cid:13)(cid:13) = 1. Thus, limk→∞δBM ⊗ (Qnk , Q) = 1. ⊗ (Qnk , Q) , we conclude that [Pnk ] converges to [Q] as re- Since δBM quired. (cid:3) We finish this section by giving some upper bounds for the tensorial Banach-Mazur distance δBM ⊗ . Proposition 3.14. Let Pi, Qi ⊂ Rdi, i = 1, . . . , l be 0-symmetric convex bodies. Then, (1) δBM (P1 ⊗π ··· ⊗π Pl, Q1 ⊗π ··· ⊗π Ql) ≤ δBM (P1, Q1)··· δBM (Pl, Ql) . (2) δBM (P1 ⊗ǫ ··· ⊗ǫ Pl, Q1 ⊗ǫ ··· ⊗ǫ Ql) ≤ δBM (P1, Q1) ··· δBM (Pl, Ql) . Proof. We give the proof only for the projective tensor product of 0-symmetric convex bodies. The proof for ⊗ǫ is analogous. First, we will show that for each i ∈ {1, ..., l} the following inequality holds: (3.10) δBM (P1 ⊗π ··· ⊗π Pi ⊗π ··· ⊗π Pl, P1 ⊗π ··· ⊗π Qi ⊗π ··· ⊗π Pl) ≤ δBM (Pi, Qi) . Let λ ≥ δBM (Pi, Qi) . Then, Qi ⊆ Ti (Pi) ⊆ λQi for some linear isomorphism Ti : Rdi → Rdi. By (3.6) we have P1 ⊗π ··· ⊗π TiPi ⊗π ··· ⊗π Pl = IRd1 ⊗ ··· ⊗ Ti ⊗ ··· ⊗ IRdl (P1 ⊗π ··· ⊗π Pi ⊗π ··· ⊗π Pl) . Therefore, if S = IRd1 ⊗ ··· ⊗ Ti ⊗ ··· ⊗ IRdl then P1 ⊗π ··· ⊗π Qi ⊗π ··· ⊗π Pl ⊆ S (P1 ⊗π ··· ⊗π Pi ⊗π ··· ⊗π Pl) ⊆ P1 ⊗π ··· ⊗π λQi ⊗π ··· ⊗π Pl = λ (P1 ⊗π ··· ⊗π Qi ⊗π ··· ⊗π Pl) . From this, it follows (3.10). To prove (1), observe that from the multiplicative triangle inequality of δBM and (3.10) we have: δBM (P1 ⊗π ··· ⊗π Pl, Q1 ⊗π ··· ⊗π Ql) ≤ Yl i=1 δBM (Q1 ⊗π ··· ⊗π Qi−1 ⊗π Pi ⊗π Pi+1 ⊗π ··· ⊗π Pl, Q1 ⊗π ··· ⊗π Qi−1 ⊗π Qi ⊗π Pi+1 ⊗π ··· ⊗π Pl) ≤ δBM (P1, Q1)··· δBM (Pl, Ql) . Using the previous proposition and [11, Proposition 2.4], we obtain the following (cid:3) upper bound for the tensorial Banach-Mazur distance. Corollary 3.15. For every pair of tensorial bodies P, Q ⊂ ⊗l (1) δBM (2) δBM ⊗ (P, Q) ≤ (d1 ··· dl−1)2(cid:16)Ql ⊗ (P, Q) ≤ (d1 ··· dl−1)2 (d1 ··· dl) . i=1δBM(cid:0)P i, Qi(cid:1)(cid:17) . i=1Rdi we have: CONVEX BODIES ASSOCIATED TO TENSOR NORMS 15 4. tensorial ellipsoids 2 2 by a linear isomorphism T : Rd → V . i=1Rdi, alternatively, we will say that E ⊂ ⊗l (cid:17) for some linear isomorphism T : ⊗l i=1Rdi → ⊗l In this section we give a complete description of the ellipsoids in ⊗l i=1Rdi which are also tensorial bodies (Corollary 4.3). To this end, we first introduce some definitions. Recall that an ellipsoid E ⊂ V in a vector space of dimension d is defined as the image of the Euclidean ball Bd i=1Rdi is In the case of ellipsoids in ⊗l an ellipsoid if E = T (cid:16)Bd1,...,dl i=1Rdi , providing we have identified Bd i=1Rdi is called a tensorial ellipsoid in ⊗l Definition 4.1. An ellipsoid E ⊂ ⊗l if E is also a tensorial body in ⊗l i=1Rdi. i=1Rdi will be denoted by E⊗(cid:0)⊗l The set of tensorial ellipsoids in ⊗l i=1Rdi(cid:1) . If Ei = Ti(cid:16)Bdi Hilbertian tensor product of E1, . . . ,El, introduced in [3], is defined as 2 (cid:17) , i = 1, . . . , l are ellipsoids in Rdi , i = 1, ..., l respectively, then the (see Subsection 2.1). 2 = Bd1,...,dl i=1Rdi 2 E1 ⊗2 ··· ⊗2 El := T1 ⊗ ··· ⊗ Tl(cid:16)Bd1,...,dl It can be directly proved that E1 ⊗2 ··· ⊗2 El is the closed unit ball of the Hilbert tensor product ⊗l H,i=1(cid:0)Rdi , gEi(cid:1). Thus, Hilbertian tensor products of ellipsoids are the first examples of tensorial ellipsoids. In particular for the Euclidean ball we have: (cid:17) . 2 2 ⊗π 2 . From this, we obtain that the only tensorial ellipsoids is the only ellipsoid between Bd1 2 i=1Rdi(cid:17) . = Bd1 Bd1,...,dl 2 and Bd1 2 ∈ E⊗(cid:16)⊗l 2 ⊗ǫ···⊗ǫ Bdl 2 ⊗2 ··· ⊗2 Bdl Actually, in Theorem 4.2 we prove that Bd1,...,dl ···⊗π Bdl are the Hilbertian tensor product of ellipsoids (Corollary 4.3). Theorem 4.2. Let E ⊂ ⊗l (4.1) then E = Bd1,...,dl i=1Rdi be an ellipsoid such that 2 ⊗π ··· ⊗π Bdl Bd1 2 ⊂ E ⊂ Bd1 2 ⊗ǫ ··· ⊗ǫ Bdl 2 , 2 2 . We will give the proof of the theorem at the end of the section. Before, we will prove Corollary 4.3 and several related results. Corollary 4.3. If E is a tensorial ellipsoid in ⊗l morphisms Ti : Rdi → Rdi for i = 1, . . . , l such that (cid:17) = T1(cid:16)Bd1 i=1Rdi(cid:1) . Then there exist 0-symmetric convex E = T1 ⊗ ··· ⊗ Tl(cid:16)Bd1,...,dl Proof. Assume that E belongs to E⊗(cid:0)⊗l bodies Ai ⊂ Rdi , i = 1, ..., l such that 2 (cid:17) ⊗2 ··· ⊗2 Tl(cid:16)Bdl 2 (cid:17) . i=1Rdi, then there exist linear iso- 2 A1 ⊗π ··· ⊗π Al ⊂ E ⊂ A1 ⊗ǫ ··· ⊗ǫ Al. Since E is an ellipsoid we must have that all Ai, i = 1, . . . , l are ellipsoids. Thus, there exist linear isomorphisms Ti : Rdi → Rdi, i = 1, . . . , l with Ai = Ti(cid:16)Bdi 2 (cid:17) . From CONVEX BODIES ASSOCIATED TO TENSOR NORMS 16 this and Theorem 3.12, we obtain: l 2 T1(cid:16)Bd1 2 (cid:17) ⊗π ··· ⊗π Tl(cid:16)Bdl 2 ⊗π ··· ⊗π Bdl 2 ⊂(cid:0)T −1 Bd1 2 (cid:17) ⊂ E ⊂ T1(cid:16)Bd1 1 ⊗ ··· ⊗ T −1 (cid:17) or equivalently E = T1(cid:16)Bd1 2 (cid:17) ⊗ǫ ··· ⊗ǫ Tl(cid:16)Bdl 2 (cid:17) 2 ⊗ǫ ··· ⊗ǫ Bdl (cid:1)E ⊂ Bd1 2 . (cid:1)E = Bd1,...,dl 1 ⊗ ··· ⊗ T −1 . Thus, E = 2 (cid:17) ⊗2 ··· ⊗2 Tl(cid:16)Bdl 2 (cid:17) . Therefore, Theorem 4.2 implies that (cid:0)T −1 T1 ⊗ ··· ⊗ Tl(cid:16)Bd1,...,dl Every ellipsoid E = T (cid:16)Bd1,...,dl scalar product h·,·iE :=(cid:10)T −1 (·) , T −1 (·)(cid:11)H. In view of this, the following proposition describes the relation between h·,·iE and h·,·iH on decomposable vectors, when E is a tensorial ellipsoid in Rm ⊗ Rn. Proposition 4.4. Let m, n be natural numbers. If E = T (Bm,n ) ⊂ Rm ⊗ Rn is an ellipsoid, then i=1Rdi is the closed unit ball associated to the (cid:17) ⊂ ⊗l (cid:3) 2 2 2 l (4.2) if and only if for L = T −1, T t the following relations hold : 2 ⊂ E ⊂ Bm Bm 2 ⊗π Bn 2 ⊗ǫ Bn 2 2 hx, zi hy, wi = hL (x ⊗ y) , L (z ⊗ w)iH + hL (x ⊗ w) , L (z ⊗ y)iH (4.3) for each x, z ∈ Rm and y, w ∈ Rn. ) ⊂ Rm⊗Rn is an ellipsoid then E◦ =(cid:0)T t(cid:1)−1 (Bm,n Proof. Recall that if E = T (Bm,n ) . Assume that (4.3) holds for T −1 and T t. Then, if we make x = y and z = w in (4.3), we have gE (x ⊗ y) = kxk2kyk2 and gE◦ (x ⊗ y) = kxk2kyk2. Thus, from Proposition 3.1, we get that (4.2) holds. Assume that (4.2) holds. Let x, z ∈ Rm and y, w ∈ Rn. From Proposition 3.1, we know that gE (x ⊗ y) = kxk2kyk2. Thus, (cid:13)(cid:13)T −1 (x ⊗ y)(cid:13)(cid:13)H = kxk2 kyk2 . Now, the polarization formula applied to (cid:10)T −1 (x ⊗ y) , T −1 (x ⊗ w)(cid:11)H and the lat- ter equality imply: 2 2 (4.4) From the polarization formula and (4.4), we have (cid:10)T −1 (x ⊗ y) , T −1 (x ⊗ w)(cid:11)H = kxk2 hx, zi hy, wi = kx + zk2 2 − kx − zk2 4 2 2 hy, wi . ! hy, wi Thus, using (4.4) in the last equality, we get that (4.3) holds for L = T −1. To finish the proof, observe that E◦ also satisfies (4.2), see Proposition 3.5. Hence, (4.3) holds for T t. Lemma 4.5. Let E be a tensorial ellipsoid in ⊗l i=1Rdi. For every zl ∈ ∂Bdl 2 , let (cid:3) and Ezl := i−1 zl (E) . Then, if izl : Rd1 ⊗ ··· ⊗ Rdl−1 → Rd1 ⊗ ··· ⊗ Rdl−1 ⊗ Rdl x1 ⊗ ··· ⊗ xl−1 → x1 ⊗ ··· ⊗ xl−1 ⊗ zl, 2 ⊗π ··· ⊗π Bdl Bd1 2 ⊂ E ⊂ Bd1 2 ⊗ǫ ··· ⊗ǫ Bdl 2 CONVEX BODIES ASSOCIATED TO TENSOR NORMS 17 one has Bd1 2 ⊗π ··· ⊗π B dl−1 2 ⊂ Ezl ⊂ Bd1 2 ⊗ǫ ··· ⊗ǫ B dl−1 2 . Proof. Let h·,·iE be the scalar product associated to E. From the definition of Ezl, we know that it is an ellipsoid. By h·,·izl , gEzl (·) we denote the scalar product and the Minkowski functional determined by Ezl. Thus, for every u ∈ Rd1 ⊗ ··· ⊗ Rdl−1 , we have gEzl (u) = gE (izl (u)) . Since E is a tensorial ellipsoid, from Proposition 3.1, gEzl (cid:0)x1 ⊗ ··· ⊗ xl−1(cid:1) =(cid:13)(cid:13)x1(cid:13)(cid:13)2 ···(cid:13)(cid:13)xl−1(cid:13)(cid:13)2 g(Ezl)◦(cid:16)x1 ⊗ ··· ⊗ xl−1(cid:17) = sup . We also have: = sup gEzl (a)≤1(cid:12)(cid:12)(cid:12)Da, x1 ⊗ ··· ⊗ xl−1EH(cid:12)(cid:12)(cid:12) gE(izl (a))≤1(cid:12)(cid:12)(cid:12)Dizl (a) , x1 ⊗ ··· ⊗ xl−1 ⊗ zlEH(cid:12)(cid:12)(cid:12) ≤ gE◦(cid:16)x1 ⊗ ··· ⊗ xl−1 ⊗ zl(cid:17) =(cid:13)(cid:13)x1(cid:13)(cid:13)2 ···(cid:13)(cid:13)(cid:13) xl−1(cid:13)(cid:13)(cid:13)2 . Therefore, from Proposition 3.1, we know Ezl is a tensorial body w.r.t. Bdi 2 , i = (cid:3) 1, . . . , l. Proof. (of Theorem 4.2) The proof will be divided into two parts. First, we will prove the theorem for tensorial ellipsoids in Rm ⊗ Rn. Then, for the general case we will use induction on l, the number of factors on the tensor product ⊗l 2 ⊗ǫ Bn 2 . If E = T (Rm ⊗ Rn) for some linear isomorphism on Rm ⊗ Rn, then from Proposition 4.4, (4.3) holds for T −1, T t. Thus, for x, z ∈ Rm and y, w ∈ Rn and S = T T t we have: Step 1. Suppose E ⊂ Rm ⊗ Rn is an ellipsoid such that Bm 2 ⊂ E ⊂ Bm 2 ⊗π Bn i=1Rdi . hx, zi hy, wi = (cid:10)S−1 (x ⊗ y) , z ⊗ w(cid:11)H +(cid:10)S−1 (x ⊗ w) , z ⊗ y(cid:11)H hx, zi hy, wi = hS (x ⊗ y) , z ⊗ wiH + hS (x ⊗ w) , z ⊗ yiH 2 2 . , On the other hand, for the canonical basis {eσ}σ=1,...,d ⊆ Rd, {ei}i=1,...,m ⊆ Rm and {ej}j=1,...,n ⊆ Rn let Φ(m,n) : Rm ⊗ Rn → Rd be the bijective map such that Φ(m,n) (ei ⊗ ej) = e(i−1)n+j for 1 ≤ i ≤ m and 1 ≤ j ≤ n. Clearly, Φ(m,n) preserves h·,·iH and the standard scalar product h·,·i on Rd. Hence if S := Φ(m,n)SΦ−1 (m,n) we have: (cid:10)e(i−1)n+j , e(k−1)n+l(cid:11) = D S−1e(i−1)n+j , e(k−1)n+lE +D S−1e(i−1)n+l, e(k−1)n+jE (cid:10)e(i−1)n+j , e(k−1)n+l(cid:11) = D Se(i−1)n+j , e(k−1)n+lE +D Se(i−1)n+l, e(k−1)n+jE 2 2 , , CONVEX BODIES ASSOCIATED TO TENSOR NORMS 18 for 1 ≤ i, k ≤ m and 1 ≤ j, l ≤ n. Therefore, for W = S, S−1 : 1 0 0 (4.5) −(cid:10)W e(i−1)n+l, e(k−1)n+j(cid:11) if k = i, l = j if k = i, l 6= j if k 6= i, l = j if k 6= i, l 6= j Hence, the positive definite matrices associated to S, S−1 can be written using the ma- . (cid:10)W e(i−1)n+j , e(k−1)n+l(cid:11) =  trices: Aki :=(cid:16)D Se(i−1)n+j , e(k−1)n+lE(cid:17)l,j and Bki :=(cid:16)D S−1e(i−1)n+j , e(k−1)n+lE(cid:17)l,j That is, S = (Aki)k,i and S−1 = (Bki)k,i for 1 ≤ k, i ≤ m. Clearly, Aki, Bki ∈ Mn,n (R) for all 1 ≤ k, i ≤ m. Moreover, from (4.5), it follows that Aki and Bki, k 6= i, are antysimmetric matrices and Akk = Bkk = In, k = i (In is the identity matrix). From this and the symmetry of S, S−1, we know that Aik = −Aki. Thus, S, S−1 satisfy (4.6) (see Lemma 4.6 at the end of this section) and S = Id. The latter implies that the linear isomorphism T is such that T T t = Id, so it is an orthogonal map on ⊗l H,i=1Rdi and E = Bm,n Step 2. As we mentioned at the beginning of the proof, this case will be proved by induction on the number l of factors on the tensor product. To simplify the notation, in this part of the proof we use the symbol k·kQ to denote the Minkowski functional associated to a 0-symmetric convex body Q. The case l = 2 was already proved. Now we assume that the result holds for l − 1. This means that for every tensorial ellipsoid E ⊂ ⊗l−1 2 ⊗π ··· ⊗π dl−1 d1,...,dl−1 2 ⊂ E ⊂ Bd1 2 i=1Rdi satisfying (4.1), and let k·kE be its Let E be a tensorial ellipsoid in ⊗l Minkowski functional. By Lemma 4.5, for every zl ∈ ∂Bdl 2 ⊗π ··· ⊗π dl−1 . Applying the induction hypothesis we obtain B 2 i ⊗ ··· ⊗ xl−1 Ezl = B . This finishes the first part of the proof. ⊂ Ezl ⊂ Bd1 d1,...,dl−1 2 i=1Rdi such that Bd1 , we have E = B 2 ⊗ǫ ··· ⊗ǫ B 2 we have Bd1 i=1Rdi we have i=1 x1 dl−1 2 dl−1 2 B 2 . Since, by Proposition 3.5, E◦ also satisfies (4.1) , we also have i=1 i=1 i ⊗ ··· ⊗ xl−1 x1 i ∈ ⊗l−1 i ⊗ ··· ⊗ xl−1 x1 2 ⊗ǫ ··· ⊗ǫ B . Therefore, for every PN i ⊗ zl(cid:13)(cid:13)(cid:13)(cid:13)E (cid:13)(cid:13)(cid:13)(cid:13)XN =(cid:13)(cid:13)(cid:13)(cid:13)XN =(cid:13)(cid:13)(cid:13)(cid:13)XN =(cid:13)(cid:13)(cid:13)(cid:13)XN i (cid:13)(cid:13)(cid:13)(cid:13)Ezl i (cid:13)(cid:13)(cid:13)(cid:13)H i (cid:13)(cid:13)(cid:13)(cid:13)H (cid:13)(cid:13)(cid:13)(cid:13)XN ψ :(cid:16)Rd1 ⊗ ··· ⊗ Rdl−1(cid:17) ⊗ Rdl → Rd1 ⊗ ··· ⊗ Rdl−1 ⊗ Rdl i ⊗ xl, i ⊗ zl(cid:13)(cid:13)(cid:13)(cid:13)E◦ i ⊗ ··· ⊗ xl−1 x1 i ⊗ ··· ⊗ xl−1 x1 i ⊗ ··· ⊗ xl−1 x1 i ⊗ ··· ⊗ xl−1 (cid:16)x1 i ⊗ ··· ⊗ xl−1 i (cid:17) ⊗ xl → x1 i=1 i=1 i=1 . . Now, consider the canonical isomorphism and denote by E the ellipsoid in (cid:0)Rd1 ⊗ ··· ⊗ Rdl−1(cid:1)⊗ Rdl determined by this isomor- phism and E. Then, for each non-zero xl ∈ Rdl, and u = PN i ⊗ ··· ⊗ xl−1 ∈ i=1 x1 i CONVEX BODIES ASSOCIATED TO TENSOR NORMS 19 Rd1 ⊗ ··· ⊗ Rdl−1 we have (cid:13)(cid:13)(cid:13) u ⊗ xl(cid:13)(cid:13)(cid:13) E And (cid:13)(cid:13)(cid:13) u ⊗ xl(cid:13)(cid:13)(cid:13) E◦ i=1 i=1 = sup x1 i ⊗ ··· ⊗ xl−1 i ⊗ ··· ⊗ xl−1 x1 =(cid:13)(cid:13)(cid:13)(cid:13)XN i ⊗ xl(cid:13)(cid:13)(cid:13)(cid:13)E i (cid:13)(cid:13)(cid:13)(cid:13)H(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)XN xl(cid:13)(cid:13)(cid:13)2 a∈ E(cid:12)(cid:12)(cid:12)Da, u ⊗ xlEH(cid:12)(cid:12)(cid:12) kψ(a)kE≤1(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)ψ (a) ,XN ≤(cid:13)(cid:13)(cid:13)(cid:13)XN i ⊗ ··· ⊗ xl−1 x1 =(cid:13)(cid:13)(cid:13)(cid:13)XN i ⊗ ··· ⊗ xl−1 x1 i ⊗ xl(cid:13)(cid:13)(cid:13)(cid:13)E◦ i (cid:13)(cid:13)(cid:13)(cid:13)H(cid:13)(cid:13)(cid:13) xl(cid:13)(cid:13)(cid:13)2 = sup i=1 i=1 i=1 . = kukH(cid:13)(cid:13)(cid:13) xl(cid:13)(cid:13)(cid:13)2 i ⊗ xl(cid:29)H(cid:12)(cid:12)(cid:12)(cid:12) = kukH(cid:13)(cid:13)(cid:13) xl(cid:13)(cid:13)(cid:13)2 . x1 i ⊗ ··· ⊗ xl−1 2 B = Bd d1,...,dl−1 2 d1,...,dl−1 2 ⊗π Bdl desired result. ⊗ǫ Bdl 2 . 2 ⊂ E ⊂ B Thus, by Proposition 3.1, E is a tensorial ellipsoid in (cid:0)Rd1 ⊗ ··· ⊗ Rdl−1(cid:1) ⊗ Rdl, and d1,...,dl−1 2 Now, let d = d1 ··· dl−1. Then, by the identification B 2.1) and the case of two factors proved in Step 1., we know that E = Bd,dl since E = ψ(cid:16) E(cid:17) and ψ is an orthogonal map, we have that E = Bd1,...,dl The next lemma is used in the proof of Theorem 4.2. For a given n ∈ N, In ∈ 2 (see Subsection . Finally, Mn×n (R) will denote the identity matrix of dimension n. Lemma 4.6. Let m, n ∈ N and d = mn. If S ∈ Md×d (R) is a positive definite matrix and there exist antisymmetric matrices Aki, Bki ∈ Mn,n (R) , k = 1, . . . , m− 1, i = k + 1, . . . , m such that (4.6) S =  . . . B1,m . . . B2,m ... . . . −B1,m −B2,m . . . In . . . A1,m . . . A2,m ... . . . −A1,m −A2,m . . . In , S−1 =  2 which is the (cid:3) Then S = Id. Proof. Let n ≥ 1 be fixed. We will prove the result by induction on m. Assume now that m = 2. In this case, Step 1. If m = 1, then d = n. By definition of S, S = Id and the result is proved. In −B12 In −A12 B12 In ... A12 In ...     ... ... . Then A12 In (cid:21) and S−1 =(cid:20) In −B12 S =(cid:20) In −A12 SS−1 =(cid:20)In − A12B12 A12 + B12 In − A12B12(cid:21) =(cid:20)In In.(cid:21) In(cid:21) . B12 0 −A12 − B12 12 = 0. Since A12 is antisymmetric, the latter equality 0 Thus, B12 = −A12 and A2 implies that At 12A12 = 0 so A12 = 0 which completes the proof. CONVEX BODIES ASSOCIATED TO TENSOR NORMS 20 matrices: −A1,m−1 −A2,m−1 In −A12 ... In −B12 ... A1,m A2,m ... Am−1,m A12 In ... B12 In ... . . . A1,m−1 ··· A2,m−1 . . . ··· ... In Step 2. Assume the result is valid for m− 1. By E, F, G, H we denote the following E :=  G :=   n(m−1),n(m−1)  n(m−1),n(m−1) In(cid:21) , S−1 =(cid:20) G H In(cid:21) F tG + InH t F tH + In(cid:21) =(cid:20) In(m−1) 0n,n(m−1) Clearly, since S is a positive definite matrix then S−1, E, G are positive definite ma- trices. Also,  n(m−1),n  n(m−1),n , F :=  , H :=  SS−1 =(cid:20) EG + F H t EH + F In . . . B1,m−1 ··· B2,m−1 . . . ··· ... In S =(cid:20) E F F t 0n(m−1),n In (cid:21) . B1,m B2,m ... Bm−1,m −B1,m−1 −B2,m−1 , , H t and (4.8) Therefore, F tH +In = In and F tH = 0n,n. Since we also have F tG+InH t = 0n,n(m−1) then H t = −F tG. This yields to (4.7) From the previous equations we get 0 = F tH = −F tGF and H = −GF. F tGF = 0n,n. Now, if we write Fi, i = 1, 2, . . . , n for the columns of F, then from (4.8) we have (4.9) F t i GFi = 0. Since G are positive definite matrix, from (4.9) we know that each Fi = 0, i = 1, 2, . . . , n and F = 0. This and (4.7) imply H = 0. Finally, we are in position to apply our inductive hypothesis to E and E−1 = G. (cid:3) Then, E = In(m−1) which implies S = Id. References [1] A. Acín, N. Gisin, and B. Toner. Grothendieck's constant and local models for noisy entangled quantum states. Phys. Rev. A, 73:062105, 2006. [2] A. Arias and J. Farmer. On the structure of tensor products of lp-spaces. Pacific Journal of Mathematics, 175(1):13–37, 1996. [3] G. Aubrun and S. Szarek. Tensor products of convex sets and the volume of separable states on n qudits. Physical Review A, 73(2):022109, 2006. [4] G. Aubrun and S. Szarek. Alice and Bob Meet Banach: The Interface of Asymptotic Geometric Analysis and Quantum Information Theory, volume 223. American Mathematical Soc., 2017. [5] J. Briët, A. Naor, and O. Regev. Locally decodable codes and the failure of cotype for projective tensor products. Electron. Res. Announc. Math. Sci., 19:120–130, 2012. [6] V. De Silva and L-H. Lim. Tensor rank and the ill-posedness of the best low-rank approximation problem. SIAM Journal on Matrix Analysis and Applications, 30(3):1084–1127, 2008. [7] K. Defant and K. Floret. Tensor Norms and Operator Ideals, volume 176. North Holland Math- ematics Studies, 1st Edition edition, 1992. CONVEX BODIES ASSOCIATED TO TENSOR NORMS 21 [8] J. Diestel, A. Grothendieck, J. Fourie, and J. Swart. The metric theory of tensor products: Grothendieck's résumé revisited. American Mathematical Soc., 2008. [9] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. [10] K. Efremenko. 3-query locally decodable codes of subexponential length. In STOC'09- Proceedings of the 2009 ACM International Symposium on Theory of Computing, pages 39–44. 2009. [11] M. Fernández-Unzueta. The Segre cone of Banach spaces and multilinear mappings. arXiv:1804.10641. [12] O. Giladi, J. Prochno, C. Schütt, N. Tomczak Jaegermann, and E. Werner. On the Geometry of Projective Tensor Products. J. Funct. Anal., 273(2):471–495, 2017. [13] A. Grothendieck. Résumé de la théorie métrique des produits tensoriels topologiques. Soc. de Matemática de São Paulo, 1956. [14] W. Hackbusch. Numerical tensor calculus. Acta numerica, 23:651–742, 2014. [15] K. Życzkowski, P. Horodecki, A. Sanpera, and M. Lewenstein. Volume of the set of separable states. Phys. Rev. A, 58:883–892, 1998. [16] R. Kadison and J. Ringrose. Fundamentals of the theory of operator algebras. Academic Press, 1983. [17] S. Khot and A. Naor. Grothendieck-type inequalities in combinatorial optimization. Comm. Pure Appl. Math., 65(7):992–1035, 2012. [18] T. Kolda and B. Bader. Tensor decompositions and applications. SIAM review, 51(3):455–500, 2009. [19] J. Landsberg. Tensors: geometry and applications. Graduate Studies in Mathematics, 128, 2011. [20] M-H. Lim. Additive preservers of non-zero decomposable tensors. Linear Algebra and its Appli- cations, 428:239–253, 2008. [21] M. Marcus and B. Moyls. Transformation on tensor product spaces. Pacific Journal of Mathe- matics, 9(4):1215–1221, 1959. [22] H. Minkowski. Geometrie der Zahlen. 1910. Teubner, Leipzig, 1927. [23] G. Pisier. Grothendieck's theorem, past and present. Bull. Amer. Math. Soc. (N.S.), 49(2):237– 323, 2012. [24] I. Pitowsky. New Bell inequalities for the singlet state: going beyond the Grothendieck bound. J. Math. Phys., 49(1):012101, 11, 2008. [25] R. Ryan. Introduction to tensor products of Banach spaces. Springer monographs in mathematics, 2002. [26] R. Schneider. Convex bodies: the Brunn-Minkowski theory, volume 44 of Encyclopedia of Math- ematics and Applications, 1993. [27] S. Szarek, E. Werner, and K. Życzkowski. How often is a random quantum state k-entangled? Journal of Physics A. Mathematical and Theoretical, 44(4):045303, 15, 2011. [28] N. Tomczak-Jaegermann. Banach-Mazur distances and finite-dimensional operator ideals, vol- ume 38. Longman Sc & Tech, 1989. [29] N. Vannieuwenhoven, J. Nicaise, R. Vandebril, and K. Meerbergen. On generic nonexistence of the schmidt–eckart–young decomposition for complex tensors. SIAM Journal on Matrix Analysis and Applications, 35(3):886–903, 2014. [30] R. Westwick. Transformation on tensor spaces. Pacific Journal of Mathematics, 23(3), 1967. [31] R. Westwick. Transformation on tensor spaces II. Linear and Multilinear Algebra, 40(1):81–92, 1995. Centro de Investigación en Matemáticas (Cimat), A.P. 402 Guanajuato, Gto., Méx- ico E-mail address: [email protected] , [email protected]
1508.01961
1
1508
2015-08-08T21:59:01
On the complexity of some inevitable classes of separable Banach spaces
[ "math.FA" ]
In this paper, we study the descriptive complexity of some inevitable classes of Banach spaces. Precisely, as shown in [Go], every Banach space either contains a hereditarily indecomposable subspace or an unconditional basis, and, as shown in [FR], every Banach space either contains a minimal subspace or a continuously tight subspace. In these notes, we study the complexity of those inevitable classes as well as the complexity of containing a subspace in any of those classes.
math.FA
math
ON THE COMPLEXITY OF SOME INEVITABLE CLASSES OF SEPARABLE BANACH SPACES. B. M. BRAGA. Abstract. In this paper, we study the descriptive complexity of some in- evitable classes of Banach spaces. Precisely, as shown in [Go], every Banach space either contains a hereditarily indecomposable subspace or an uncondi- tional basis, and, as shown in [FR], every Banach space either contains a min- imal subspace or a continuously tight subspace. In these notes, we study the complexity of those inevitable classes as well as the complexity of containing a subspace in any of those classes. . A F h t a m [ 1 v 1 6 9 1 0 . 8 0 5 1 : v i X r a 1. Introduction. Let SB = {X ⊂ C(∆) X is linear and closed} be our coding for the separable Banach spaces, and endow SB with the Effros-Borel structure (for definitions, see Section 2). Given a separable Banach space X ∈ SB, it is natural to ask about the complexity of the isomorphism class hXi = {Y ∈ SB Y ∼= X}. Is hXi Borel, analytic, coanalytic? This question has only been solved for some very specific Banach spaces. For example, S. Kwapien showed (see [Kw], Proposition 3.1) that a Banach space X is isomorphic to a Hilbert space if, and only if, X has both type and cotype equal 2. As ℓ2 is the only separable Hilbert space (up to isomorphism), this gives us a characterization of separable Banach spaces which are isomorphic to ℓ2 in terms of type and cotype. Using this characterization, it is not hard to show that hℓ2i is Borel (see [B], page 130). In fact, hℓ2i is the only known example of a Borel isomorphism class. We recall the following problem (see [B], Problem 2.9). Problem 1. Let X ∈ SB be a separable Banach space whose isomorphism class hXi is Borel. Is X isomorphic to ℓ2? For some classical spaces, such as Lp[0, 1], with p 6= 2, Pelczynski's universal space U , and C(∆), it had been shown that their isomorphism classes are analytic non Borel (see [B], page 130 and Theorem 2.3, and [K], Theorem 33.24, respec- tively). The problem of classifying Banach spaces up to isomorphism is extremely com- plex. Indeed, considering the isomorphism relation ∼= as a subset of SB × SB, we have that ∼= is a complete analytic equivalence relation (see [FLR], Theorem 31). In a more precise sense, separable Banach spaces up to isomorphism may serve as a complete invariant for any other reasonable class of mathematical objects. There- fore, it makes sense to study easier problems such as classifying Banach spaces up to its subspaces. We have then another natural problem: given a Banach space 2010 Mathematics Subject Classification. Primary: 46B20. Key words and phrases. Minimal Banach spaces, continuously tight Banach spaces, HI spaces, descriptive set theory, trees, Banach spaces dichotomies. 1 2 B. M. BRAGA. X ∈ SB, what can we say about the complexity of CX = {Y ∈ SB X ֒→ Y }? This question had been solved by B. Bossard (see [B], Corollary 3.3). Theorem 2. (Bossard) Let X ∈ SB. Borel. If X is infinite dimensional, then CX is complete analytic. If X is finite dimensional, then CX is In [Go2], W. T. Gowers had started with a new classification theory for Banach spaces. Indeed, after Gowers and Maurey had solved the unconditional basis prob- lem, and proved the existence of hereditarily indecomposable Banach spaces (see [GM]), Gowers proved that every Banach space either contains a subspace with an unconditional basis, or a hereditarily indecomposable subspace (see [Go]). Later, in [Go2], Gowers used Ramsey methods in order to refine his dichotomy. Many other dichotomies had been proved by V. Ferenczi and C. Rosendal in [FR], for exam- ple, their main dichotomy says that every Banach space either contains a minimal subspace or a continuously tight subspace (see [FR], Theorem 3.13). In these notes, we study the descriptive complexity of the inevitable classes above as well as the complexity of containing a subspace in any of those inevitable classes. The table below summarizes the lower and upper bounds known for the complexity of those classes. In the table below, an asterisk before the complexity indicates that this bound was already known before this paper. All the other bounds are either computed in this paper or are given by trivial computations. Classes of Banach spaces Minimal spaces Containing a minimal subspace Continuously tight spaces Containing a continuously tight subspace HI spaces Containing a HI subspace Spaces with unconditional basis Containing unconditional basis Lower bound Upper bound Σ1 Π1 Σ1 Π1 Σ1 1-hard 1-hard 1-hard 1-hard 1-hard Σ1 1-hard ∆1 2 (∗) Σ1 2 Σ1 2 Σ1 2 Π1 1 Σ1 2 Σ1 1 Σ1 1 In particular, we completely compute the descriptive complexity of the class of hereditarily indecomposable spaces, and the class of spaces containing an un- conditional basis, by showing those classes are complete coanalytic and complete analytic, respectively. The properties of being a continuously tight space and of having an unconditional basis are more properties about basis then about the spaces themselves. Therefore, it is more natural to ask what is the complexity of the set of continuously tight basis, and the complexity of the set of unconditional basis. We have the following. Classes of basis Continuously tight basis Unconditional basis Lower bound 1-hard Π1 Upper bound Σ1 2 Borel This paper is organized in the following way. In Section 2, we give all the basic background on both descriptive set theory and the theory of Banach spaces necessary for these notes. In Section 3, we prove a general lemma for Banach spaces with an unconditional basis (xs)s∈θ indexed by a tree θ ∈ WF. In Section 4, we work with ℓp-Baire sums of basic sequences as a tool to compute lower bounds for COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 3 the complexity of classes of Banach spaces. Then, Theorem 9 gives a sufficient condition for the set of separable Banach spaces containing a subspace in a given class to be Σ1 1-hard (in particular, non Borel). Section 5 is dedicated to the computation of the complexities in the tables above. We apply Theorem 9 to the class of spaces containing a hereditarily indecomposable subspace (Subsection 5.1, Corollary 10), and to the class of spaces containing a continuously tight subspace (Subsection 5.2, Corollary 11), obtaining a lower bound for the complexity of those classes. In order to obtain the results above, we rely on the method of ℓp-Baire sums of separable Banach spaces (see Lemma 8). However, this method does not allow us to obtain any information about the complexity of the class of spaces (i) containing an unconditional basis, (ii) containing a minimal subspace, or (iii) containing a continuously tight subspace. Hence, in Subsection 5.3, we define a parameterized version of Tsirelson space in order to show that the class of spaces containing a minimal subspace is Σ1 1-hard. Such parametrization will show that it is impossible to Borel separate the set of separable Banach spaces containing c0 from the set of Tsirelson-saturated Banach spaces (see Theorem 12). In Subsection 5.4, we use the methods of Subsection 5.3 in order to show that the class of continuously tight spaces is Π1 1-hard (Theorem 17). At last, in Subsection 5.6, we follow Argyros presentation of how to construct HI extensions of ground norms ([AT], Chapters II and III) in order to show that the set of hereditarily indecomposable separable Banach spaces is complete coanalytic (Theorem 27). We also obtain that the set of hereditary indecomposable Banach spaces cannot be Borel separated from the set of Banach spaces containing ℓ1. By the same methods, in Subsection 5.7, we show that the class of Banach spaces containing an unconditional basis is complete analytic (Theorem 29). Problem 3. For all the classes in the tables above, we have only completely com- puted the descriptive complexity of the set of HI spaces, and the set of spaces containing an unconditional basis. What are the exact complexities of the remain- ing classes? At last, we would like to make a small remark. In [FR], the authors actually present Theorem 1.1 as their main dichotomy, which says that every Banach space contains either a minimal subspace or a tight subspace. This dichotomy does not explicitly say anything about continuously tight spaces. However, their proof for Theorem 1.1 shows that any Banach space X must contain either a minimal Banach subspace or a continuously tight Banach subspace (see [FR], Theorem 3.13). We have two reasons for choosing to deal with continuously tight spaces instead of tight spaces, (i) we easily get better estimates for the upper bound of continuously tight spaces and spaces containing a continuously tight subspace, (ii) Rosendal had shown (see [R], Appendix) that, for minimal Banach spaces not containing c0, there exists a notion of being continuously minimal, and this notion coincides with being minimal (see Subsection 5.5). 2. Background. A separable metric space is said to be a Polish space if there exists an equiv- alent metric in which it is complete. A continuous image of a Polish space into another Polish space is called an analytic set and a set whose complement is an- alytic is called coanalytic. A measure space (X, A), where X is a set and A is a 4 B. M. BRAGA. σ-algebra of subsets of X, is called a standard Borel space if there exists a Polish topology on this set whose Borel σ-algebra coincides with A. We define Borel, analytic and coanalytic sets in standard Borel spaces by saying that these are the sets that, by considering a compatible Polish topology, are Borel, analytic, and coanalytic, respectively. Observe that this is well defined, i.e., this definition does not depend on the Polish topology itself but on its Borel structure. A function between two standard Borel spaces is called Borel measurable if the inverse image of each Borel subset of its codomain is Borel in its domain. We usually refer to Borel measurable functions just as Borel functions. Notice that, if you consider a Borel function between two standard Borel spaces, its inverse image of analytic sets (resp. coanalytic) is analytic (resp. coanalytic) (see [S], Proposition 1.3, page 50). Given a Polish space X, the set of analytic (resp. coanalytic) subsets of X is 1, or Π1 denoted by Σ1 1. Hence, the terminology Σ1 1-set) is used to refer to analytic sets (resp. coanalytic sets). We define the projective hierarchy by induction of n. We say a subset of a standard Borel space is Π1 n-set, and we say that a subset is Σ1 n-set (see [D], Chapter 1). For each n ∈ N, we denote by ∆1 n the subsets of a standard Borel space which are both Σ1 1(X)). We usually omit X, and simply write Σ1 n+1 if it is a Borel image of a Π1 n if it is the complement of a Σ1 1(X) (resp. Π1 1-set (resp. Π1 n = Σ1 n ∩ Π1 n. n, i.e., ∆1 n and Π1 Let X be a standard Borel space. An analytic (resp. coanalytic) subset A ⊂ X is said to be complete analytic (resp. complete coanalytic) if for every standard Borel space Y and every analytic subset B ⊂ Y (resp. coanalytic), there exists a Borel function f : Y → X such that f −1(A) = B. This function is called a Borel reduction from B to A, and B is said to be Borel reducible to A. n-set B ⊂ Y (resp. Π1 n-hard) means that A is at least as complex as any Σ1 Let X be a standard Borel space. We call a subset A ⊂ X Σ1 n-hard) if for all standard Borel space Y and all Σ1 n-hard (resp. Π1 n-set) there exists a Borel reduction from B to A. Therefore, saying that a set A ⊂ X is n-hard (resp. Π1 Σ1 n-set (resp. Π1 n-set) in the projective hierarchy. With this terminology we have that A ⊂ X is n-complete (resp. Π1 Σ1 n-hard) and Σ1 n (resp. Π1 As there exist analytic non Borel (resp. coanalytic non Borel) sets we have that 1-hard (resp. Π1 1-hard) sets are non Borel. Also, if X is a standard Borel space, 1-hard) subset 1-hard). If A is Σ1 A ⊂ X, and there exists a Borel reduction from a Σ1 B of a standard Borel space Y to A, then A is Σ1 also analytic (resp. coanalytic), then A is Σ1 n-complete) if, and only if, A is Σ1 1-hard (resp. Π1 1-hard (resp. Π1 n-hard (resp. Π1 1-complete (resp. Π1 n). 1-complete). Consider a Polish space X and let F (X) be the set of all its non empty closed sets. We endow F (X) with the Effros-Borel structure, i.e., the σ-algebra generated by {F ⊂ X F ∩ U 6= ∅}, where U varies among the open sets of X. It can be shown that F (X), endowed with the Effros-Borel structure, is a standard Borel space (see [K], Theorem 12.6). The following well known lemma (see [K], Theorem 12.13) will be crucial in some of our proofs. COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 5 Lemma 4. (Kuratowski and Ryll-Nardzewski selection principle) Let X be a Polish space. There exists a sequence of Borel functions (Sn)n∈N : F (X) → X such that {Sn(F )}n∈N is dense in F , for all closed F ⊂ X. In these notes we will only work with separable Banach spaces. We denote It is well known that every the closed unit ball of a Banach space X by BX . separable Banach space is isometrically isomorphic to a closed linear subspace of C(∆) (see [K], page 79), where ∆ denotes the Cantor set. Therefore, C(∆) is called universal for the class of separable Banach spaces and we can code the class of separable Banach spaces, denoting it by SB, by SB = {X ⊂ C(∆) X is a closed linear subspace of C(∆)}. As C(∆) is clearly a Polish space we can endow F (C(∆)) with the Effros-Borel structure. It can be shown that SB is a Borel set in F (C(∆)) and hence it is also a standard Borel space (see [D], Theorem 2.2). It now makes sense to wonder if specific classes of separable Banach spaces are Borel or not (in our coding SB). Throughout these notes we will denote by {Sn}n∈N the sequence of Borel func- tions Sn : SB → C(∆) given by Lemma 4, with X = C(∆) (more precisely, the restriction of those functions to SB). Consider the standard Borel space C(∆)N, and let B = {(xj)j∈N ∈ C(∆)N (xj )j∈N is a basic sequence}. It is easy to see that B is a Borel set. Therefore, we have that B is a standard Borel space, and we can code all the basic sequences as elements of B. We can now wonder about the descriptive complexity of some specific classes of basis. Let (en)n∈N be a basic sequence in a Banach space X ∈ SB. By a standard Skolem hull construction, there is a countable subfield F of R containing the ratio- nals such that for any finite linear combination λ0e0 + ... + λnen, with n ∈ N, and λ1, ..., λn ∈ F, we have kλ0e0 + ... + λnenk ∈ F. By working with F instead of Q, we guarantee that any F-linear combination of (en)n∈N can be normalized and remain a F-linear combination of (en)n∈N. Clearly, the F-span of (en)n∈N is dense in span{en}. Let D be the set of normalized blocks of (en)n∈N, i.e., the set of all λ0e0 + ... + λnen, with n ∈ N, and λ1, ..., λn ∈ F. Clearly, D is countable. A block sequence of (en)n∈N is a sequence (yn)n∈N such that, (i) there exist increasing sequences of natural numbers (pn)n∈N and (qn)n∈N such that pi ≤ qi < pi+1, for all i ∈ N, and (ii) for all n ∈ N yn = qn X j=pn ajej, for some sequence (an)n∈N of real numbers. We define a finite block sequence analo- gously. Denote by bb(en) the set of normalized block sequences with the coefficients (an)n∈N in F. Endowing D with the discrete topology, we can see bb(en) as a closed subset of DN, so bb(en) is a standard Borel space. Denote by f bb(en) the set of normalized finite block sequences with the coefficients (an)n∈N in F. Letting 6 B. M. BRAGA. D<N = ∪nDn, we can see f bb(en) as a closed subset of D<N. Hence, f bb(en) is a standard Borel space. Let X and Y be Banach spaces. We write X ֒→ Y if X can be linearly embedded If K > 0, we write X ֒→K Y if X can be K-embedded in Y , i.e., if into Y . there exists an embedding T : X → Y such that kT kkT −1k ≤ K. If (xn)n∈N and (yn)n∈N are two sequences in Banach spaces, we write (xn)n∈N ≈ (yn)n∈N if (xn)n∈N is equivalent to (yn)n∈N, i.e., if the map xn 7→ yn induces a linear isomorphism between span{xn} and span{yn}. Similarly, if K > 0, we write (xn)n∈N ≈K (yn)n∈N if the induced isomorphism if a K-isomorphism. Denote by N<N the set of all finite tuples of natural numbers plus the empty set. Given s = (s0, ..., sn−1), t = (t0, ..., tm−1) ∈ N<N we say that the length of s is s = n, si = (s0, ..., si−1), for all i ∈ {1, ..., n}, s0 = {∅}, s (cid:22) t iff n 6 m and si = ti, for all i ∈ {0, ..., n − 1}, i.e., if t is an extension of s. We define s < t analogously. Define the concatenation of s and t as sat = (s0, ..., sn−1, t0, ..., tm−1). A subset T of N<N is called a tree if t ∈ T implies ti ∈ T , for all i ∈ {0, ..., t}. We denote the set of trees on N by Tr. A subset I of a tree T is called a segment if I is completely ordered and if s, t ∈ I with s (cid:22) t, then l ∈ I, for all l ∈ T such that s (cid:22) l (cid:22) t. Two segments I1, I2 are called completely incomparable if neither s (cid:22) t nor t (cid:22) s hold if s ∈ I1 and t ∈ I2. As N<N is countable, 2N<N . Thus, it is also Borel in 2N<N (the power set of N<N) is Polish with its standard product topology. If we think about Tr as a subset of 2N<N , it is easy to see that Tr is a Gδ set in 2N<N . As Tr is Borel in the Polish space 2N<N , we have that Tr is a standard Borel space. A β ∈ NN is called a branch of a tree T if βi ∈ T , for all i ∈ N, where βi is defined analogously as above. We call a tree T well-founded if T has no branches and ill-founded otherwise, we denote the set of well-founded and ill-founded trees by WF and IF, respectively. It is well known that WF is a complete coanalytic set of Tr, hence IF is complete analytic (see [K], Theorem 27.1). 3. A Lemma. In this section, we prove a basic lemma that will be essential in many of the main results of this paper. Fix a compatible enumeration of N<N, i.e., a sequence (sn)n∈N in N<N such that sn (cid:22) sm implies n 6 m and for all s ∈ N<N there exists n ∈ N such that sn = s. With this enumeration in mind, if θ ∈ Tr, we say that a sequence (xs)s∈θ is a basis for a given Banach space X, if (xsn )n∈N is a basis for X, where N = {n ∈ N sn ∈ θ}. We now show the following. Lemma 5. Let θ ∈ WF, and let X be a Banach space with an unconditional basis (es)s∈θ. Let Y be an infinite dimensional subspace of X. Then Y contains a basic sequence (yk)k∈N equivalent to a semi-normalized block sequence (xk)k∈N of (es)s∈θ with completely incomparable supports. Before we prove this lemma, let's show a simple lemma that will be important in our proof. We say that an operator T : X → Y is strictly singular if for all infinite dimensional subspace Z ⊂ X, TZ : Z → Y is not an embedding. Lemma 6. Let (X1, k · k1), ..., (Xn, k · kn) be Banach spaces, and let Y ⊂ ⊕n i=1Xi be an infinite dimensional subspace. Consider the standard projections Pj : ⊕n i=1Xi → Xj, for all j ∈ {1, ..., n}. Then, there exists j ∈ {1, ..., n} such that Pj : Y → Xj is not strictly singular. COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 7 i=1Xi. As this is a finite sum, we can assume X = (⊕n Proof. Let X = ⊕n j=1Xj)ℓ1, i.e., if (x1, ..., xn) ∈ X, then kxkX = Pj kxjkj. Assume towards a contradiction that Pj is strictly singular, for all j ∈ {1, ..., n}. By a classic property of strictly singular operators (see [D], Proposition B.5), we know that for all ε > 0 there exists an infinite dimensional subspace A ⊂ Y such that kPjAk < ε, for all j ∈ {1, ..., n}. Pick x ∈ A of norm one. Then, as x = (P1(x), ..., Pn(x)), we have kxkX 6 nε. By choosing ε < 1/n we get a contradiction. (cid:3) Proof of Lemma 5. For each s ∈ θ, let Λs = {λ ∈ N sa(λ) ∈ θ}, and enumerate each Λs, say Λs = {λi s i ∈ N}. For each s ∈ θ, let θs = {τ ∈ θ s (cid:22) τ }. For each s ∈ θ, let Ps : X → X be the projection of X onto span{eτ s (cid:22) τ }. For each s ∈ θ, and each n ∈ N, consider the projections Qs,n : X → ⊕n i=1Psa(λi s)(X) (aτ )τ ∈θ → (aτ )τ ∈∪n i=1θsa(λi s ) . Claim: There exists s ∈ θ such that Ps : Y → X is not strictly singular, but Qs,n : Y → ⊕n i=1Psa(λi s)(X) is strictly singular, for all n ∈ N. Let's assume the claim is true and finish the proof of the lemma. Indeed, if Ps : Y → X is not strictly singular, we can substitute Y by an infinite dimensional Z ⊂ Y such that Ps : Z → X is an isomorphism onto its image. Let E = Ps(Z), and notice that E ⊂ span{eτ s (cid:22) τ }. By Lemma 6, we can actually assume that E ⊂ span{eτ s ≺ τ }. Hence, for all x ∈ E, we have that x = limn Qs,n(x). Claim: There exists a normalized sequence (yj)j∈N in E such that Qs,n(yj) → 0, as j → ∞, ∀n ∈ N. Indeed, for all n ∈ N, there exists a normalized sequence (yn j )j∈N in E such that kQs,n(yn j )k < 1/j, for all j ∈ N. Let (yj)j∈N be the diagonal sequence of the sequences (yn j , for all j ∈ N. Say M is the unconditional constant of (es)s∈θ. Then, m 6 n implies kQs,m(x)k 6 M kQs,n(x)k, for all x ∈ E. Hence, (yj)j∈N has the required property. j )j∈N, i.e., yj = yj Say (εi)i∈N is a sequence of positive real numbers converging to zero. As Qs,n(x) → x, as n → N, for all x ∈ E, we can pick increasing sequences of natural numbers (nk)k∈N, and (lk)k∈N such that (i) kQs,lk(ynk ) − ynk kθ < εk, for all k ∈ N, and (ii) kQs,lk(ynk+1 )kθ < εk, for all k ∈ N. 8 B. M. BRAGA. For each k ∈ N, let xk = Qs,lk (ynk ) − Qs,lk−1(ynk ). Choosing (εk)k∈N converging to zero fast enough, we have that (xk)k∈N is semi- normalized and, by the principle of small perturbations, that (ynk )k∈N is equivalent to (xk)k∈N (see [AK], Theorem 1.3.9). Clearly, (xk)k∈N has completely incompara- ble supports. As Y contains a sequence equivalent to (xk)k∈N, the proof is complete. We now prove our first claim. Suppose the claim is false, i.e., suppose that for all s ∈ θ such that Ps : Y → X is not strictly singular, there exists n ∈ N such that Qs,n : Y → ⊕n s)(X) is not strictly singular. By Lemma 6, if Qs,n : Y → ⊕n s)(X) is not strictly singular, there exists m 6 n such that i=1Psa(λi i=1Psa(λi Psaλm s = Psaλm s ◦ Qs,n : Y → X is not strictly singular. Therefore, for all s ∈ θ such that Ps : Y → X is not strictly singular, there exists s′ ≻ s such that Ps′ : Y → X is not strictly singular. Now notice that P∅ : Y → X is not strictly singular, indeed, P∅ = Id. Therefore, by applying the last paragraph ω times, we get a sequence (sn)n∈N such that Psn : Y → X is not strictly singular, and sn ≺ sn+1, for all n ∈ N. In particular, sn ∈ θ, for all n, absurd, as θ is well-founded. (cid:3) 4. ℓp-Baire sums. We now deal with ℓp-Baire sums of basic sequences, this tool will be crucial in many of our results in these notes. Fix a basic sequence E = (en)n∈N, and p ∈ [1, ∞). Let us define a Borel function ϕ : Tr → SB in the following manner. For each θ ∈ Tr and x = (x(s))s∈θ ∈ c00(θ) we define kxkE,p,θ = supn(cid:16) n X i=1 (cid:13)(cid:13) X s∈Ii 1 p p E(cid:17) x(s)es(cid:13)(cid:13) n ∈ N, I1, ..., In incomparable segments of θo, where k.kE is the norm of span{E}. Define ϕE,p(θ) as the completion of c00(θ) under the norm k.kE,p,θ. The space ϕE,p(θ) is known as the ℓp-Baire sum of span{E} (indexed by θ). Similarly, we define k.kE,0,θ as kxkE,0,θ = supn(cid:13)(cid:13)X s∈I x(s)es(cid:13)(cid:13)E I segment of θo, and let ϕE,0(θ) be the completion of (c00(θ), k.kE,0,θ). We denote by (es)s∈θ the sequence in c00(θ) such that, for each τ ∈ θ, the coordinate es(τ ) equals 1 if s = τ and zero otherwise. Considering a compatible enumeration of N<N, as in Section 3, the sequence (es)s∈θ is clearly a basis for ϕ(θ). Pick Y ⊂ C(∆) such that ϕE,p(N<N) is isometrically isomorphic to Y . If we consider the natural isometries of ϕE,p(θ) into ϕE,p(N<N), we can see ϕE,p as a COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 9 Borel function from Tr into SB. This gives us the following (see [S], Proposition 3.1, page 79). Proposition 7. Let p ∈ [1, ∞) or p = 0. Then, the map ϕE,p : Tr → SB defined above is Borel. The following lemma summarizes the main properties of the ℓp-Baire sum that we will need later in these notes. Lemma 8. Let E be a basic sequence. The Borel function ϕE,p : Tr → SB defined above has the following properties: (i) If θ ∈ IF, then ϕE,p(θ) contains span{E}. (ii) If θ ∈ WF, then ϕE,p(θ) is ℓp-saturated, i.e., every infinite dimensional sub- space of ϕE,p(θ) contains an isomorphic copy of ℓp. The analogous is true for ϕE,0 : Tr → SB, i.e., (i) If θ ∈ IF, then ϕE,0(θ) contains span{E}. (ii) If θ ∈ WF, then ϕE,0(θ) is c0-saturated, i.e., every infinite dimensional sub- space of ϕE,0(θ) contains an isomorphic copy of c0. Proof. If θ ∈ IF, clearly ϕE,p(θ) contains span{E}. Indeed, let β be a branch of θ, then span{E} ∼= ϕE,p(β) ֒→ ϕE,p(θ), where by ϕE,p(β) we mean ϕE,p applied to the tree {s ∈ N<N s < β}. Say θ ∈ WF, and let E be an infinite dimensional subspace of ϕE,p(θ). By Lemma 5, E has a basic sequence equivalent to a semi-normalized block sequence (xk)k∈N with completely incomparable supports. It is trivial to check that a semi- normalized block sequence with completely incomparable supports is equivalent to the ℓp-basis (resp. c0-basis). So we are done. (cid:3) Let P ⊂ SB. We say that P is a class of Banach spaces if P is closed under isomorphism, i.e., for all X, Y ∈ SB, X ∈ P and Y ∼= X imply Y ∈ P. We say that a class of Banach spaces P ⊂ SB is pure if X ∈ P implies Y ∈ P, for all subspace Y ⊂ X. We say a class P ⊂ SB is almost-pure if for all X ∈ P and all infinite dimensional Y ⊂ X there exists an infinite dimensional subspace Z ⊂ Y such that Z ∈ P. Theorem 9. Say P ⊂ SB is almost-pure and that ℓp (reps. c0) does not embed in any Y ∈ P, for some p ∈ [1, ∞). Then CP = {Y ∈ SB ∃Z ∈ P, Z ֒→ Y } is Σ1 1-hard. In particular, the same is true if P is pure and does not contain ℓp (reps. c0), for some p ∈ [1, ∞). Proof. This is a simple application of Lemma 8. Indeed, let E be a basis for C(∆), and consider the restriction of ϕE,p to the set of infinite trees, say ITr. It is easy to see that ITr is Borel (see [S], Proposition 1.6, page 72), so ϕE,pITr is a Borel function. By Lemma 8, ϕE,pITr : ITr → SB is a Borel reduction from IF to CP . Therefore, as IF is Σ1 (cid:3) 1-hard, CP is Σ1 1-hard. 5. Descriptive complexity of the inevitable classes. 5.1. Spaces containing a hereditarily indecomposable subspace. In 1991 W. T Gowers and B. Maurey independently solved the unconditional basic se- quence problem, i.e., they constructed a Banach space with no unconditional basic sequence (see [GM]). It was noticed by W. B. Johnson that the space constructed 10 B. M. BRAGA. by Gowers and B. Maurey not only had no unconditional basic sequence but was also hereditarily indecomposable. We say that an infinite dimensional Banach space X is hereditatily indecom- posable if none of X subspaces can be decomposed as a direct sum of two infi- nite dimensional subspaces. Clearly, the class HI = {X ∈ SB X is hereditarily indecomposable} is a pure class and it contains no ℓp. Hence, Theorem 9 gives us the following. Corollary 10. CHI is Σ1 1-hard. We come back to the class of hereditarily indecomposable spaces in Subsection 5.6, where we show that the set HI is complete coanalytic, and that CHI is at most Σ1 2. 5.2. Spaces containing a continuously tight subspace. In [FR], Ferenczi and Rosendal defined a new class of Banach spaces, the class of continuously tight spaces, and proved many interesting properties about this class. In Theorem 3.13 of [FR], for example, Ferenczi and Rosendal have shown that every Banach space must contain either a minimal subspace or a continuously tight subspace, giving us another dichotomy for Banach spaces. Denote by [N] the set of increasing sequences of natural numbers. We can see [N] as a Borel subset of NN. A basic sequence (en)n∈N is called continuously tight if there exists a continuous function f : bb(en) → [N] such that, for all block basis ¯y = (yn)n∈N ∈ bb(en), if we set Ij = {m ∈ N f (¯y)2j ≤ m ≤ f (¯y)2j+1}, then for all infinite set A ⊂ N, span{¯y} 6֒→ span{en n 6∈ ∪j∈AIj }, i.e., span{¯y} does not embed into span{en} avoiding an infinite number of the intervals Ij . A space with a continuously tight basis is called continuously tight. The Tsirelson space is an example of a continuously tight Banach space (see [FR], Corollary 4.3). For a detailed study of continuously tight spaces and other related properties (e.g., tight spaces, tight with constants, tight by range, etc) see [FR]. Let CT = {X ∈ SB X is continuoulsy tight} be our coding for the class of continuously tight separable Banach spaces. Corollary 11. CCT is Σ1 1-hard. Proof. Proposition 3.3 of [FR] says that continuously tight spaces contain no min- imal subspaces, and Theorem 3.13 of [FR] says that a space with no minimal sub- spaces contains a continuously tight subspace. Therefore, CT is an almost-pure class. Also, again by Proposition 3.3 of [FR], we have that no elements of CT con- tain an isomorphic copy of ℓp. Hence, Theorem 9 gives us that CCT is Σ1 1-hard. (cid:3) We cannot obtain any lower bound for the complexity of the set of continuously tight Banach spaces with the method of ℓp-Baire sums. We come back to the class of continuously tight spaces in Subsection 5.4, where we show that the set CT is Π1 1-hard by using a different method. COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 11 5.3. Spaces containing a minimal subspace. A Banach space X is called min- imal if every infinite dimensional subspace of X contains an isomorphic copy of X. We now turn our attention to the following: Although M = {X ∈ SB X is minimal} is clearly a pure class, M contains c0 and ℓp, for all p ∈ [1, ∞). Therefore, Theorem 9 does not say anything about the complexity of CM. However, we can use the construction of Tsirelson space by T. Figiel and W. B. Johnson (see [FJ], Section 2) in order to construct a ϕ : Tr → SB that will solve our problem. Fix a compatible enumeration of N<N, i.e., (sn)n∈N such that sn (cid:22) sm implies n 6 m and for all s ∈ N<N there exists n ∈ N such that sn = s. This enumeration give us an order on N<N. With this ordering in mind, we say that E0 < E1 if max E0 < min E1, for all finite sets E0, E1 ⊂ N<N. We write k < E if {sk} < E. Given θ ∈ Tr, E ⊂ θ, and x = Pn∈N asn esn ∈ c00(θ) (for some (asn )n∈N ∈ RN), We define a Borel function ϕ : Tr → SB as, for each θ ∈ Tr and each x = we let Ex = Psn∈E asn esn . (x(s))s∈θ ∈ c00(θ), let (k.kθ,m)m∈N be inductively defined by kxkθ,0 = kxk0, and kxkθ,m+1 = max{kxk0, 1 2 max k X i=1 kEixkθ,m}, for all m ∈ N, where the "inner" maximum above is taken over all k ∈ N and all completely incomparable finite sets (Ei)k i=1 (Ei ⊂ θ, for all i ∈ {1, ..., k}) such that k 6 E1 < ... < Ek (for the definition of completely incomparable sets of a tree, see Section 2). Exactly as we have for the standard Tsirelson space, we can define a norm k.kθ as kxkθ = lim m→∞ kxkθ,m, for all x ∈ c00(θ). We define ϕ(θ) to be the completion of c00(θ) under this norm. This norm can be implicitly defined as kxkθ = max{kxk0, 1 2 max k X i=1 kEixkθ}, where the "inner" maximum above is taken over all k ∈ N and all completely incomparable finite sets (Ei)k i=1 (Ei ⊂ θ, for all i ∈ {1, ..., k}) such that k 6 E1 < ... < Ek. By the universality of C(∆) for separable Banach spaces, we can identify ϕ(N<N) with an isometric copy inside of C(∆). As we can identify each ϕ(θ) with a subspace of ϕ(N<N) in a natural fashion, we can see ϕ as a Borel function from Tr to SB (see [S], Proposition 3.1, page 79, for similar arguments). Theorem 12. Let ϕ : Tr → SB be defined as above. Then ϕ is a Borel function with the following properties (i) c0 ֒→ ϕ(θ), for all θ ∈ IF, and (ii) ϕ(θ) has the following property if θ ∈ WF: for every infinite dimensional subspace E ⊂ ϕ(θ), there exists a further subspace F ⊂ E isomorphic to 12 B. M. BRAGA. an infinite dimensional subspace of Tsirelson space, i.e., ϕ(θ) is Tsirelson- saturated. In particular, CM is Σ1 Tsirelson-saturated Banach spaces. 1-hard, and Cc0 cannot be Borel separated from the set of Before proving this theorem, notice the following trivial consequence of Lemma 6. Lemma 13. A finite sum of spaces satisfying property (ii) of Theorem 12 still has property (ii). Proof of Theorem 12. If θ ∈ IF, it is clear that c0 ֒→ ϕ(θ). Say θ ∈ WF, let us show that every infinite dimensional subspace of ϕ(θ) contains a subspace isomorphic to an infinite dimensional subspace of Tsirelson's space. Say E ⊂ ϕ(θ) is an infinite dimensional subspace. As θ ∈ WF, Lemma 5 gives us that E contains a sequence equivalent to a block sequence (yn)n∈N of ϕ(θ) with completely incomparable supports. We will be done once we prove the following claim. Claim: (yn)n∈N is equivalent to a subsequence of the standard basis of Tsirelson space. Let k · kT,θ be the standard Tsirelson norm on c00(θ), i.e., kxkT,θ = max{kxk0, 1 2 max k X i=1 kEixkT,θ}, where the "inner" maximum above is taken over all k ∈ N and all finite sets (Ei)k i=1 (Ei ⊂ θ, for all i ∈ {1, ..., k}) such that k 6 E1 < ... < Ek. The only difference between k · kθ and k · kT,θ is that in k · kT,θ we do not have the restriction of (Ei)k being completely incomparable. i=1 It is clear that the the basis (es)s∈N<N of the completion of (c00(θ), k · kT,θ) is equivalent to the standard basis of Tsirelson space. Moreover, the basis (es)s∈θ of the completion of (c00(θ), k · kT,θ) is equivalent to a subsequence of the standard basis of Tsirelson space, if θ is infinite (this because this norm has no dependance on the structure of the tree T ). Also, we clearly have (5.1) k k X i=1 aiyikθ 6 k k X i=1 aiyikT,θ, for all a1, ..., ak ∈ R. Mimicking the proof of Lemma II.1 of [CS] we have the following lemma. Lemma 14. Let (pn)n∈N be a increasing sequence of natural numbers. Let (esn )n∈N be the standard basis of ϕ(θ). Let yn = Ppn+1 i=pn+1 biesi (for all n ∈ N) be a normal- ized block sequence of (esn )n∈N and assume (yn)n∈N has completely incomparable supports. Then k X n∈N anespn+1kθ 6 k X n∈N anynkθ, COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 13 for any sequence of scalars (an)n∈N. Proof. It is enough to show that, for any sequence (an)n∈N, we have k X n∈N anespn +1kθ,m 6 k X n∈N anynkθ, for all m ∈ N. Let us proceed by induction on m ∈ N. For m = 0 the result is clear. Assume the equation above holds for a fixed m ∈ N. Let x = Pn∈N anepn+1 and y = Pn∈N anyn. Fix k ∈ N, and completely incomparable finite sets (En)k such that k 6 E1 < ... < Ek. Consider the sum n=1 1 2 k X i=1 kEixkθ,m. Since the support of x is contained in {pn + 1 n ∈ N}, we may assume that Ei ⊂ {pn + 1 n ∈ N}, for all i ∈ N. Applying the inductive hypothesis, we have 1 2 k X i=1 kEixkθ,m 6 1 2 k X i=1 anynkθ k X n∈N pn+1∈Ei As pn + 1 ∈ Ei implies k 6 pn + 1, and as (yn)n∈N has completely incomparable supports, the sum on the right hand side of the equation above is allowed as an "inner" sum in the definition of the norm k.kθ. Therefore, we have 1 2 k X i=1 kEixkθ,m 6 kykθ, for all k ∈ N, and all completely incomparable finite sets (En)k E1 < ... < Ek. Hence, for any sequence of scalars (an)n∈N, we have n=1 such that k 6 k X n∈N anespn +1kθ,m+1 6 k X n∈N anynkθ, and we are done. (cid:3) Let (bn)n∈N be the sequence of scalars such that our block sequence (yn)n∈N can be written as yn = Ppn+1 i=pn+1 biesi . Then, Lemma 14 gives us that (5.2) k X n∈N anespn+1kθ 6 k X n∈N anynkθ, for any sequence of scalars (an)n∈N. As the supports of (yn)n∈N are completely incomparable we have that 14 (5.3) B. M. BRAGA. k X n∈N anespn +1kθ = k X n∈N anespn+1kT,θ, for any sequence of scalars (an)n∈N. By Proposition II.4 of [CS], we have (5.4) 1 18 k X n∈N anynkT,θ 6 k X n∈N anespn+1 kT,θ, for all sequence of scalars (an)n∈N. Therefore, putting Equation 5.1, Equation 5.2, Equation 5.3, and Equation 5.4 together, we have that k k X i=1 anespn+1kT,θ 6 k k X i=1 anynkθ 6 18k k X i=1 anespn +1kT,θ, for all a1, ..., ak ∈ R. Hence, the sequence (yn)n∈N as a sequence in ϕ(θ) is equivalent to the sequence (espn+1)n∈N as a sequence in the Tsirelson space (the completion of (c00(θ), k · kT,θ)). To conclude that ϕ(θ) contains no minimal subspaces if θ ∈ WF, recall that Tsirelson space contains no minimal subspaces (see [CS], Corollary VI.b.6), so we are done. (cid:3) It is easy to see, by simply counting quantifiers, that the set CM is at most Σ1 3, i.e., the Borel image of a set that can be written as the complement of the Borel image of a coanalytic set. However, by using Gowers' theorem, Rosendal was able to find a better upper bound for CM (see [R], Appendix). Theorem 15. (C. Rosendal) CM is Σ1 2. Problem 16. What is the exact complexity of CM? Is it Σ1 2-complete? In Subsection 5.5, we talk a little bit about Rosendal's proof for CM being Σ1 2. We will notice that Rosendal's proof also gives us that M is at most ∆1 2. 5.4. Continuously tight Banach spaces. We are now capable of giving a lower bound for the complexity of CT, the set of continuously tight spaces. Recall, a Banach space X with a basis (xk)k∈N is said to be strongly asymptotic ℓp if there exists a function f : N → N and a constant C > 0 such that any set of m unit vectors in span{xk k ≥ f (m)} with disjoint supports is C-equivalent to the basis of ℓm p . The Tsirelson space with its standard basis is an example of an strongly asymptotic ℓ1 space (see [CS], Chapter V). Also, a Banach space X is said to be crudely finitely representable in a Banach space Y if there exists M > 0 such that every finite subspace of X M -embeds into Y . The proof below is an adaptation of the proof of Proposition 4.2, in [FR]. Theorem 17. The set of continuously tight spaces CT is Π1 set of continuously tight basis, say CT , is Π1 1-hard. 1-hard. Moreover, the Proof. Let ϕ : Tr → SB be the map in Lemma 12. As c0 embeds into ϕ(θ) if θ ∈ IF, we only need to show that ϕ(θ) is continuously tight if θ ∈ WF. Indeed, as the map θ ∈ Tr 7→ (es)s∈θ ∈ B is a Borel map, this is enough to prove both assertions of the theorem. COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 15 A Banach space X with a basis is said to be tight with constants if no Banach space embeds uniformly into its tail subspaces (see [FR], Proposition 4.1). Let's show that ϕ(θ) is tight with constants, for all θ ∈ WF. As spaces which are tight with constants are also continuously tight (see Proposition 18 below) we will be done. Assume towards a contradiction that, for some K > 0, there exists a space Y which K-embeds into all tail subspaces of ϕ(θ). By Theorem 12, we can assume, by taking a subspace, that Y is generated by a sequence (yk)k∈N which is equivalent to a subsequence of the basis of Tsirelson space. Therefore, (yk)k∈N is unconditional and strongly asymptotic ℓ1. Let C > 0 and f : N → N be as in the definition of strongly asymptotic ℓ1 spaces. By Proposition 1 of [Jo], we have that for all m ∈ N, there exists N (m) ∈ N such that (y1, ..., ym) is 2K-equivalent to a sequence of vectors in the linear span of N (m) disjointly supported unit vectors in any tail of Y . In particular, in the tail span{yk k ≥ f (N (m))}. Therefore, as Y is strongly asymptotic ℓ1, we have that (y1, ..., ym) 2KC-embeds into ℓ1, for all m ∈ N. So Y is crudely finitely representable in ℓ1, and therefore Y embeds into L1 (see [AK], Theorem 11.1.8). Hence, as (yk)k∈N is unconditional asymptotic ℓ1, we have that Y contains ℓ1 (see [DFKO], Proposition 5), absurd, because Y is a subspace of Tsirelson space. (cid:3) Proposition 18. Let X be a Banach space with basis (en)n∈N. Say (en)n∈N is tight with constants (see definition in the proof above). Then (en)n∈N is continuously tight. Proof. For this, we will use details of the proof of Proposition 4.1 of [FR]. Precisely, let X ∈ SB be a Banach space with a basis (en)n∈N which is tight with constants. For each L ∈ N, let c(L) > 0 be a constant such that if two block sequences of (en)n∈N differ from at most L terms, then they are c(L)-equivalent. For each (yn)n∈N ∈ bb(en), let us define a sequence (Ij)j∈N of finite intervals of natural numbers. By Proposition 4.1 of [FR], for each K, m ∈ N, there exists an l > m such that (5.5) span{yn m ≤ n ≤ l} 6֒→K span{en n ≥ l}. Let l1 ∈ N be the minimal l ∈ N as above, for m = 1, and K = c(1). Let I1 = [1, l1], where if a ≤ b ∈ N, [a, b] = {n ∈ N a ≤ n ≤ b}. Assume we had already defined finite intervals I1 < ... < Ij−1 ⊂ N and numbers l1 < ... < lj−1 ∈ N. Define lj as the minimal l ∈ N as in (5.5) above, for m = max{Ij−1} + 1, and K = j · c(max{Ij−1} + 1). Let Ij = [max{Ij−1} + 1, lj]. By the proof of Proposition 4.1 of [FR], we have that, for all K ∈ N, span{yn n ∈ IK } 6֒→K span{en n 6∈ IK }. In particular, for all infinite set A ⊂ N, span{yn} 6֒→ span{en n 6∈ ∪j∈AIj}. Hence, we had defined a map ¯y = (yn) 7→ (Ij )j∈N, and we will be done if this 16 B. M. BRAGA. assignment is continuous, i.e., if there exists a continuous function f : bb(en) → [N] such that Ij = [f (¯y)2j, f (¯y)2j+1]. For this, we only need to notice that in order to obtain a finite chunk of the sequence of intervals (Ij )j∈N, say I1, ..., IK, we only need to know a finite chunk of (yn)n∈N, precisely, y1, ..., ymax{IK }. So we are done. (cid:3) Proposition 19. CT, CCT and CT are at most Σ1 2. Proof. This is a simple matter of counting quantifiers and the fact that we only quantify over standard Borel spaces in the definition of those three classes. Indeed, for CT , for example, we have (en)n∈N ∈ CT ⇔ ∃ continuous f : bb(en) → [N], ∀(yn)n∈N ∈ bb(en), ∀ infinite A ⊂ N, ∀(xn)n∈N ∈ span{en n 6∈ ∪j∈AIj }N, ∀K ∈ N, (xn)n∈N 6≈K (yn)n∈N. The only quantifier that demands some explanation is "∃ continuous f : bb(en) → [N]". For this, let N[<N ] be the set of finite increasing sequence of natural numbers. Then it is easy to see that a continuous function f : bb(en) → [N] gives us a function g : f bb(en) → N[<N ] such that g(¯y) (cid:22) g(¯x), if ¯y (cid:22) ¯x, and vice versa, where f bb(en) is the set of finite block sequences of (en)n∈N (see Section 2). So, as f bb(en) is countable, the space of functions f bb(en) → N[<N ] is a standard Borel space, so we are done. The same arguments work for CT and CCT. (cid:3) 5.5. Mininal spaces. It follows straight forward from the definition of minimal Banach spaces that M = {X ∈ SB X is minimal} is Π1 2. In this subsection we show that M is also Σ1 2. Hence, M is at most ∆1 2. Using Gowers' theorem (see [Go2]), and a corollary of the solution of the distor- tio problem (see [OS], and [OS2]), Rosendal had shown (see [R], Appendix) that if a Banach space X not containing c0 contains a minimal subspace then there exists a basic sequence (en)n∈N in X, a block subsequence (yn)n∈N ∈ bb(en), and a con- tinuous function f : bb(yn) → span{en}N such that, for all ¯z = (zn)n∈N ∈ bb(yn), we have span{f (¯z)} ⊂ span{zn} and (en)n∈N ≈ f (¯z). By counting quantifiers, the set of Banach spaces satisfying the property above is at most Σ1 2. Clearly, if a Banach space satisfies the property above, then it contains a minimal subspace, indeed, span{yn} is minimal. As the set of Banach spaces containing c0 is Σ1 1-complete actually), this gives us that CM is at most Σ1 2. 1 (Σ1 We now notice that this also gives us an equivalent characterization of minimality. Indeed, Rosendal's result clearly implies that if X is a minimal Banach space not containing c0 then there exists a basic sequence (en)n∈N in X, a block subsequence (yn)n∈N ∈ bb(en), such that X ֒→ span{yn}, and there exists a continuous function f : bb(yn) → span{en}N such that, for all ¯z = (zn)n∈N ∈ bb(yn), we have COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 17 span{f (¯z)} ⊂ span{zn} and (en)n∈N ≈ f (¯z). By counting quantifiers, the set of Banach spaces satisfying the property above is at most Σ1 2. Notice that if a Banach space satisfies the property above, then it is minimal. Therefore, as the set of minimal Banach spaces containing c0 is the set of spaces that embed into c0, and as this set is easily seen to be analytic, we have that M is at most Σ1 2, we have the following. 2. As M is also Π1 Proposition 20. The class of minimal Banach spaces M is at most ∆1 2. 5.6. Hereditarily indecomposable spaces. Let HI = {X ∈ SB X is hereditarily indecomposable}. In Subsection 5.1, we proved, using the method of ℓp-Baire sums (Corollary 10), that CHI is Σ1 1-hard. However, this method does not allow us to obtain any information about the complexity of HI. In this subsection, we will use a more complex construction in order to compute the complexity of HI. Precisely, we follow Argyros presentation (see [AT], Chapters II and III) of how to construct HI extensions of ground norms in order to define a Borel function ϕ : Tr → SB such that ϕ−1(HI) = WF∞, where WF∞ denotes the subset of infinite well-founded trees. This will show that HI is Π1 1-hard. As a Banach space X is hereditarily indecomposable if, and only if, for all subspaces Z, W ⊂ X and all ε > 0, there exist z ∈ Z and w ∈ W such that kz − wk < εkz + wk, we can easily show that HI is coanalytic. Therefore, we will show that HI is complete coanalytic. Theorem 21. HI is coanalytic. Proof. This is a simple consequence of the fact that X ∈ HI if, and only if, for all subspaces Z, W ⊂ X and all ε > 0, there exists z ∈ Z and w ∈ W such that kz − wk < εkz + wk. Indeed, notice that X ∈ HI ⇔ ∀Z, W ⊂ X, ∀ε ∈ Q+, ∃n, m ∈ N kSn(Z) − Sm(W )k < εkSn(Z) + Sm(W )k, where "∀Z, W ⊂ X" means "for all subspaces Z, W ⊂ X". As {(Z, W, X) ∈ SB3Z, W ⊂ X} is well known to be Borel ([S], see Lemma 1.9, page 73), we are done. (cid:3) This trivially gives us the following upper bound for the complexity of CHI. Corollary 22. CHI is Σ1 2. The only thing left to show is that HI is Π1 1-hard. For this, let us define a special Borel map ϕ : Tr → SB such that ϕ−1(HI) = WF∞. As the construction of such ϕ will heavily rely on the construction of HI extensions of a ground norm, we tried to be consistent with Argyros notation ([AT], Chapters II and III). We believe this will make the presentation more clear for the reader which is familiar with HI spaces. Therefore, the notation used to denote some spaces in this subsection will be slightly different from the notation chosen in the rest of these notes. We will make sure to point out the differences as they appear though. To start with, we will denote by X(DG(θ)) the resulting space ϕ(θ). 18 B. M. BRAGA. First, we fix a compatible enumeration for N<N, say (si)i∈N. Let (esi)i∈N be the standard unit basis of c00(N<N). Let x = (x(s))s∈N<N ∈ c00(N<N), and x∗ = (x∗(s))s∈N<N ∈ c00(N<N), we define x∗(x) = X s∈N<N x(s)x∗(s), i.e., the notation "∗" means that we will consider x∗ as a functional on c00(N<N). For x∗ = (x∗(s))s∈N<N ∈ c00(N<N), we let supp(x∗) = {s ∈ N<N x∗(s) 6= 0}. Let G = (cid:8) n X i=1 aiesji n ∈ N, sji ∈ N<N, sj1 ≺ ... ≺ sjn , ai = 1(cid:9). The set G is called a ground set (for a definition of ground sets see [AT], Definition II.1, page 21). We define YG as the completion of c00(N<N) under the norm kxkG = sup{g(x) g ∈ G}, for all x ∈ c00(N<N). For each θ ∈ Tr, we let YG(θ) denote the subspace of YG generated by {es s ∈ θ}. So YG(N<N) = YG. We will define X(DG(θ)) as a ground norm extension of YG(θ). Remark: Notice that, according to the notation of Lemma 8, it is clear that YG(θ) ∼= ϕE,0(θ), if E is the standard ℓ1-basis. For each j ∈ N, let Aj = {F ⊂ N F 6 j}, where F is the cardinality of F ⊂ N. As in Subsection 5.3, for finite sets E1, E2 ⊂ N, we write E1 < E2 if max E1 < min E2. Let g1, ..., gn ∈ c00(N<N). We write g1 < ... < gn if supp(g1) < l=1 be a finite sequence in c00(N<N) such that g1 < ... < gn, ... < supp(gn). Let (gl)n and m ∈ N, we define the (An, 1 l=1 as the functional g = 1 m (g1 + ... + gn). m )-operation on (gl)n Fix two sequences of natural numbers (mj)j∈N and (nj)j∈N such that m1 = 2, mj+1 = m5 j , n1 = 4, and nj+1 = (5nj)s, where sj = log2 m3 j+1. We now define a norming set DG ⊂ c00(N<N) that will give us the norm we will use to define X(DG(θ)), i.e., ϕ(θ). In the definition below we use the term "n2j−1- special sequence". As this definition will play no role in our proof of the theorem and as it is a really technical definition, we chose to omit it here. The interested reader can find the precise definition of an n2j−1-special sequence in [AT], Chapter III, page 40. Let E ⊂ N<N, and x∗ = (x∗(s))s∈N<N ∈ c00(N<N). We define Ex∗ = (x∗(s))s∈E. Definition 23. We define DG as the minimal subset of c00(N<N) satisfying (i) G ⊂ DG. (ii) DG is symmetric, i.e., g ∈ DG implies −g ∈ DG. (iii) DG is closed under the restriction of its elements to intervals of N<N, i.e., if E ⊂ N<N is an interval and g ∈ DG, then Eg ∈ DG. (iv) DG is closed under the (An2j , 1 m2j )-operations, i.e., if (gl)n2j l=1 is a sequence in DG such that g1 < ... < gn2j , then g = 1 m2j (g1 + ... + gn2j ) belongs to DG. COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 19 (v) DG is closed under (An2j−1 , 1 m2j−1 )-operations on special sequences, i.e., for every n2j−1-special sequence (g1, ..., gn2j−1 ) in DG, the functional g = 1 ... + gn2j−1 ) belongs to DG. m2j−1 (g1+ (vi) DG is rationally convex. We define a norm on c00(N<N), as kxkDG = sup{g(x) g ∈ DG}, for all x ∈ c00(N<N). Let X(DG) be the completion of c00(N<N) under this norm. For each θ ∈ Tr, let X(DG(θ)) be the subspace of X(DG) generated by {es s ∈ θ}. Therefore, for each θ ∈ Tr, we assign a space ϕ(θ) = X(DG(θ)). Identify X(DG(N<N)) = X(DG) with one of its isometric copies in SB. It is clear that the map ϕ : Tr → SB such that ϕ(θ) = X(DG(θ)), is Borel (see [S], Proposition 3.1, page 79, for similar arguments). Theorem 24. Let ϕ : Tr → SB be the function defined above. Then (i) ϕ(θ) = X(DG(θ)) contains ℓ1, for all θ ∈ IF. (ii) ϕ(θ) = X(DG(θ)) is hereditarily indecomposable, for all θ ∈ WF∞. In particular, HI is Π1 1-hard, and Cℓ1 cannot be Borel separated from HI. Proof. First, notice that if θ ∈ IF then ℓ1 ֒→ ϕ(θ). Indeed, on segments I ⊂ θ, the ℓ1-norm given by the ground set G is greater than the norm given by DG. So, if θ has a branch, say β, we have X(DG(β)) ∼= ℓ1. In order to show the second part of the theorem consider the "identity" map Id : X(DG(θ)) → YG(θ). We will show that Id : X(DG(θ)) → YG(θ) is strictly singular, for all θ ∈ WF∞. Once we do that, we will be done by Theorem III.7 of [AT] (page 42). Indeed, it is clear from the proof of Theorem III.7 of [AT], that we have the following. Theorem 25. Let G be a ground set in c00(N<N), and let YG, and X(DG) be the spaces obtained as above. If Z ⊂ X(DG) is an infinite dimensional subspace, and the restriction IdZ : Z → YG is strictly singular, then Z is hereditarily indecomposable. Proposition 26. Let θ ∈ WF∞. Then Id : X(DG(θ)) → YG(θ) is strictly singular. Proof. Suppose not. Then there exists an infinite dimensional subspace Y ⊂ X(DG(θ)) such that IdY is an isomorphism with its image. We now look at Y = Id(Y ) ⊂ YG(θ). By Lemma 8 and the remark following the definition of YG(θ), there is a normalized sequence (yi)i∈N in Y ⊂ YDG(θ) which is equivalent to the standard basis of c0. In particular, there exists C > 0 such that n X i=1 (cid:13)(cid:13) yi(cid:13)(cid:13)G < C, for all n ∈ N. Moreover, we can assume, by Lemma 28 below, that there exists a sequence (gi)i∈N in G such that gi(yi) > 1 2 and gi < gi+1, for all i ∈ N, and gi(yk) < 2−(k+2), for all i 6= k. Therefore, by the definition of the norm of X(DG(θ)), we have that 20 B. M. BRAGA. n2j X i=1 (cid:13)(cid:13) ≥ yi(cid:13)(cid:13)DG 1 m2j n2j X i=1 gi(yi) + − n2j m2j ∞ X k=1 ≥ = n2j 2m2j n2j 4m2j gi(yk) 1 m2j n2j X i,k=1 i6=k 2−(k+2) As n2j m2j → ∞, as j → ∞, and as IdY is an isomorphism, we get a contradiction. (cid:3) The proof of Theorem 24 is now done. (cid:3) Theorem 27. HI is complete coanalytic. In order to prove the result above we made use of the following lemma. Lemma 28. Let ϕE,0 : Tr → SB be as in Lemma 8. If θ ∈ WF, and Y ⊂ ϕE,0(θ) is infinite dimensional, then there exists a normalized sequence (yi)i∈N in Y equivalent to the c0-basis. Moreover, there exists a sequence (gi)i∈N in G(θ) such that gi < gi+1, for all i ∈ N, gi(yi) > 1 2 , for all i ∈ N, and gi(yk) < 2−(k+2), for all i 6= k. Proof. This can be obtained by a simple modification in the proof of Lemma 5. For completeness, we write the modifications here. Assume all the notation in the proof of Lemma 5. So we have X = ϕε,0(θ), and s ∈ θ is such that Ps : Y → X is not strictly singular, but Qs,n : Y → X is strictly singular, for all n ∈ N. Let E = Ps(Z), where Z ⊂ Y is a subspace such that Ps : Z → X is an isomorphism with its image. As in Lemma 5, we can assume that E ⊂ span{eτ s ≺ τ }. For each m ∈ N, let Im : X → X be the standard projection over the first m coordinates of the basis (es)s∈θ, i.e., Im(Ps∈θ ases) = Pm i=1 asi esi. Let (yk)k∈N be a normalized sequence in E such that Qs,n(yk) → 0, as k → ∞, for all n ∈ N (we showed such sequence exists in the proof of Lemma 5). As Qs,n(x) → x, as n → N, for all x ∈ E, we have that given a sequence (εk)k∈N of positive real numbers, we can pick increasing sequences of natural numbers (nk)k∈N, (mk)k∈N, and (lk)k∈N such that (i) kQs,lk(ynk ) − ynk kθ < εk, for all k ∈ N, and (ii) kQs,lk(ynk+1 )kθ < εk, for all k ∈ N. (iii) kImk(cid:0)Qs,lk (ynk ) − Qs,lk−1(ynk )(cid:1) − (cid:0)Qs,lk (ynk ) − Qs,lk−1(ynk )(cid:1)kθ < εk, for all k ∈ N. For each k ∈ N, let xk = Imk(cid:0)Qs,lk (ynk ) − Qs,lk−1(ynk )(cid:1). Choosing (εk)k∈N converging to 0 sufficiently fast, we have that (xk)k∈N is equivalent to (ynk )k∈N, and, as (xk)k∈N has completely incomparable supports, it is easy to see that (xk)k∈N is also equivalent to the c0-basis (as in Lemma 8). Also, by taking COMPLEXITY OF SOME CLASSES OF BANACH SPACES. 21 a subsequence if necessary, we can assume that (xk)k∈N is a block sequence. Hence, as we can assume kxikθ > 1 2 , for all i ∈ N, there exists a sequence (gi)i∈N in G(θ) such that gi < gi+1, for all i ∈ N, and gi(xi) > 1 2 , for all i ∈ N, and gi(xk) = 0, for all i 6= k. Clearly, we can assume supp(gi) ⊂ supp(xi), for all i ∈ N. Hence, if (y′ i)i∈N is a i) = gi(yni) = sequence in Z such that Ps(y′ gi(xi), for all i ∈ N. Therefore, gi(y′ i) = yni, for all i ∈ N, we have that gi(y′ i) > 1 2 , for all i ∈ N. On the other hand, it is easy to see that gi(y′ Hence, as kynk − xkkθ < 2εk + εk−1, we can assume that gi(y′ i 6= k. Although (y′ we can assume (y′ k) = gi(ynk − xk), for all i 6= k. k) < 2−(k+2), for all k)k∈N is only semi-normalized, it is clear from the proof, that (cid:3) k)k∈N is normalized, so we are done. 5.7. Unconditional basis. A basic sequence (xj)j∈N in a Banach space X is un- conditional if, and only if, there exists M > 0 such that, for all n ∈ N, for all a1, ..., an ∈ Q, and all b1, ..., bn ∈ Q such that aj 6 bj, for all j ∈ {1, ..., n}, we have k k X j=1 ajxj k 6 k k X j=1 bjxjk. Therefore, it is clear that the set UB ⊂ B of unconditional basis is Borel, where B is our coding for the basic sequences (see Section 2). We now consider instead of the set of unconditional basis, the set of Banach spaces with an unconditional basis, say UB, and the set of Banach spaces containing an unconditional basis, say CUB. As X ∈ UB ⇔ ∃(xj )j∈N ∈ C(∆)N such that (xj )j∈N is an unconditional basis for X, and the condition "(xj )j∈N is a basis for X" is easily seen to be Borel, we have that UB is analytic. Analogously, CUB is also analytic. Let us now give a lower bound for CUB. As hereditarily indecomposable spaces cannot contain an unconditional basis, Theorem 24 gives us the following. Theorem 29. The set of separable Banach spaces containing an unconditional basis CUB is Σ1 1-hard. Moreover, CUB is Σ1 1-complete. Problem 30. Is UB Borel? Is UB complete analytic? What about S = {X ∈ SB X has a Schauder basis}? Similarly, as we have for UB, we can easily see that S is analytic. Is S Borel? Is S complete analytic? Acknowledgements: The author would like to thank his adviser C. Rosendal for all the help and attention he gave to this paper. Also, the author would like to thank Joe Diestel for reading this manuscript, and Spyros Argyros for his help in the HI part of these notes. [AK] F. Albiac, and N. J. Kalton, Topics in Banach Space Theory, Graduate Texts in Mathe- matics, Vol. 233, Springer, New York, 2006. References 22 B. M. BRAGA. [AT] S. Argyros, S. Todorcevic, Ramsey Methods in Analysis, Advanced Courses in Mathematics - CRM Barcelona, Birkhauser 2005. [B] B. Bossard. A coding of separable Banach spaces. Analytic and coanalytic families of Banach spaces. Fund. Math. 172 (2002), 2, 117 -- 152. [CS] P. G. Casazza, T. Shura, Tsirelson Space, Lecture Notes in Math., Vol. 1363, Springer-Verlag, Berlin, 1989. [D] P. Dodos, Banach Spaces and Descriptive Set Theory: Selected Topics, Lecture Notes in Mathematics, Vol. 1993, Springer-Verlag, Berlin, 2010. [DFKO] S. Dilworth, V. Ferenczi, D. Kutzarova, E. Odell, On strongly asymptotically ℓ p spaces and minimality, J. London Math. Soc. (2) 75 (2007) 409 -- 419. [FLR] V. Ferenczi, A. Louveau, and C. Rosendal, The complexity of classifying separable Banach spaces up to isomorphism, J. London Math. Soc. 79 (2009), 323 -- 345. [FR] V. Ferenczi, C. Rosendal, Banach sppaces without minimal subspaces, J. Functional Analysis 257 (2009), no. 1, 149 -- 193, [FJ] T. Figiel, W. B. Johnson, A uniformly convex Banach space which contains no ℓ p, Compo- sitio Mathematica 29 (1974), 179 -- 190. [Go] W. T. Gowers, A new dichotomy for Banach spaces, Geom. Funct. Anal. 6 (6) (1996), 1083 -- 1093. [Go2] W. T. Gowers, An infinite Ramsey theorem and some Banach space dichotomies, Ann. of Math. (2) 156 (3) (2002), 797 -- 833. [GM] W. T. Gowers, B. Maurey,The unconditional basic sequence problem, Journal of AMS 6, 1993, 851 -- 874. [G] G. Godefroy, Analytic sets of Banach spaces, Rev. R. Acad. Cien. serie A. Mat. (RACSAM), 104 (2) (2010), 365 -- 374. [Jo] W. B. Johnson, A reflexive Banach space which is not sufficiently Euclidean, Studia Math. 55 (1978) 201 -- 205. [K] A. S. Kechris. Classical Descriptive Set Theory. Graduate Texts in Mathematics, Vol. 156, Springer-Verlag, 1995. [Kw] S. Kwapien. Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients. Studia Math. 44 (1972), 583 -- 595. [OS] E. Odel, T. Schlumprecht, The distortionn problem. Acta Mah. 173 (194), 259 -- 281. [OS2] E. Odel, T. Schlumprecht, Distortion and asymptotic structure, in: Handbook of the Ge- ometry of Banach Spaces. Vol. 2, North-Holland, Amsterdam, 2003, 1333 -- 1360. [R] C. Rosendal, Incomparable, non-isomorphic and minimal Banach spaces, Fundamenta Math- ematicae, 183 (2004), no. 3, 253 -- 274. [S] Th. Schlumprecht, Notes on Descriptive Set Theory, and Applications to Banach Spaces. Class notes for Reading Course in Spring/Summer 2008. Department of Mathematics, Statistics, and Computer Science (M/C 249), University of Illinois at Chicago, 851 S. Morgan St., Chicago, IL 60607-7045, USA E-mail address: [email protected]
1302.5598
1
1302
2013-02-22T14:08:06
A Haagerup Inequality for $\tA_1\times\tA_1$ and $\tA_2$ Buildings
[ "math.FA", "math.GR" ]
Haagerup's inequality for convolvers on free groups may be interpreted as a result on $\tA_1$ buildings, i.e. trees. Here are proved analogous inequalities for discrete groups acting freely on the vertices of $\tA_1\times\tA_1$ and $\tA_2$ buildings. The results apply in particular to groups of type-rotating automorphisms acting simply transitively on the vertices of such buildings. These results provide the first examples of higher rank groups with property (RD).
math.FA
math
A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER Abstract. Haagerup's inequality for convolvers on free groups may be interpreted as a result on eA1 buildings, i.e. trees. Here are proved analogous inequalities for discrete groups acting freely on the vertices of eA1 × eA1 and eA2 buildings. The results apply in particular to groups of type- rotating automorphisms acting simply transitively on the vertices of such buildings. These results provide the first examples of higher rank groups with property (RD). 1. Introduction U. Haagerup has given a beautiful and useful estimate for convolvers on a free group [H, Lemma 1.4]. Suppose Γ is the free group on a set of generators N+ and let N consist of the generators from N+ and their inverses. Each c ∈ Γ can be written uniquely as c = a1a2 · · · an with aj ∈ N and ajaj+1 6= 1. This product is called the reduced word for c, and if the reduced word has n factors, we say that c is a word of length n. Haagerup's inequality applies to a function g ∈ ℓ1(Γ) supported on the words of length n and it asserts that kf ∗ gk2 ≤ (n + 1)kfk2kgk2. Denoting by ρ the right regular representation of ℓ1(Γ) on ℓ2(Γ), the inequality reads kρ(g)k ≤ (n + 1)kgk2. In [H] Haagerup's inequality was used in the course of establishing that the reduced C ∗-algebra of a free group on finitely many generators has the metric approximation property. In another development, it follows from Haagerup's inequality that Jolissaint's property (RD) holds for free groups. Haagerup's inequality was extended to word hyperbolic groups in [J, Ha1], proving that they too satisfy property (RD). This is an ingredient in the proof of the Novikov conjecture for word hyperbolic groups [CoM]. An exposition of this last result may be found in [C, Chapter III.5]. Also a recent paper of A. Nevo [N] provides a new application of the Haagerup inequality to ergodic theorems on groups. Restating Haagerup's proof geometrically, we find that it gives a result somewhat more general than was orginally stated. Suppose Γ acts freely on the vertices of a tree, and let d(u, v) denote graph theoretic distance. Fix a vertex v0 and define the length of c ∈ Γ as c = d(v0, cv0). If g ∈ ℓ1(Γ) is supported on elements of length n, then kρ(g)k ≤ (n + 1)kgk2. To recover the original inequality, consider the free group Γ acting simply transitively on its Cayley graph with respect to the generating set N and take v0 to be 1 ∈ Γ. This is the first such generalization to "higher rank" groups of either Haagerup's inequality or of property (RD). Trees are eA1 buildings. This paper generalizes Haagerup's inequality to eA1×eA1 and eA2 buildings. Let ∆ be an eA1 × eA1 or eA2 building. In §1.2 we define the shape, σ(u, v) ∈ N× N between any two vertices u, v ∈ ∆ which essentially gives the dimensions of the convex hull of the two vertices in the building sense. The index of AutS(∆) = {c ∈ Aut(∆) σ(cu, cv) = σ(u, v) for all u, v ∈ V∆} This research was partly funded by University of Newcastle RMC Project Grant number 45/290/501 and by the Italian MURST. The first two authors would like to thank the University of Newcastle study leave program for support during the later stages of this project and the University of Sassari for support and hospitality. 1 2 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER in Aut(∆) is 1 or 2. Define a function in two variables p(m, n) =((m + 1)(n + 1) Our main result is the following. 1/2(m + 1)(n + 1)(m + n + 2)pmax(m, n) + 1 (eA1 × eA1 case) (eA2 case). Theorem 1.1. Suppose ∆ is an eA1 × eA1 or eA2 building and Γ ≤ AutS(∆) acts freely on the vertices of ∆. Fix any vertex v0 ∈ ∆ and define a shape function σ on Γ by σ(c) = σ(v0, cv0). If g ∈ CΓ is supported on words of shape (m, n) and f ∈ ℓ2(Γ), then kf ∗ gk2 ≤ p(m, n) kfk2 kgk2. in [Swa], A. Valette [V] has shown that there is a Haagerup inequality applying to radial functions In the eA1 × eA1 case our bound is optimal. From an analysis of radial functions, we conjecture that the optimal value of p(m, n) in the eA2 case is 1/2(m + 1)(n + 1)(m + n + 2). Using a result on an eAn group. Among buildings of dimension 2 only eA1 × eA1 and eA2 buildings ever admit simply transitive they are also the easiest. Indeed, Haagerup's original method, broadly conceived, handles eA1 ×eA1 buildings, and only one new ingredient is required for the eA2 case. It is natural to conjecture that analogous versions of Haagerup's inequality hold for all types of affine buildings. In fact, Valette has conjectured [FRR, p. 70] that any group acting properly and cocompactly on either a Riemannian symmetric space or an affine building has property (RD). actions on their vertices. It was this that led us to consider these two cases -- by happy coincidence 1.1. Some analytical results used in the proofs. We note the following well-known results which will be used on several occasions. Their proofs are straightforward and so are omitted. Lemma 1.2. Let T : H −→ K be an operator between two Hilbert spaces, and suppose we have orthogonal decompositions H = ⊕jHj and K = ⊕kKk. Expressing T as an operator matrix [Tkj] where Tkj : Hj −→ Kk, one has kTk ≤ k [kTkjk ] k ≤ Xj,k kTkjk2!1/2 where all norms are operator norms and [kTkjk ] is the matrix with scalar entries kTkjk. Lemma 1.3. If T = [tij], S = [sij] with 0 ≤ tij ≤ sij for all i, j, then kTk ≤ kSk. 1.2. The buildings and the shape between a pair of vertices. Given an eAn building ∆, there that τ (av) = τ (v) + i for all vertices v ∈ ∆. More generally, suppose that ∆ is an eAn1 × · · · × eAnk is a type map τ defined on the vertices of ∆ such that τ (v) ∈ Z/(n + 1)Z for each vertex v ∈ ∆. A brief inductive argument shows that every a ∈ Aut(∆) gives rise to a permutation of the set of types. An automorphism a of ∆ is said to be type-rotating if there exists i ∈ {0, 1, . . . , n} such building. There is a type map τ on the vertices of ∆ where τ (v) = (τ (v)1, . . . , τ (v)k) ∈ Z/(n1 + 1)Z × · · · × Z/(nk + 1)Z is a k-tuple. Again, any automorphism a of ∆ gives rise to a permutation on the set of types. We say such an a is type-rotating if there exists a k-tuple (i1, . . . , ik) ∈ {0, 1, . . . , n1} × · · · × {0, 1, . . . , nk} such that τ (av) = (τ (v)1 + i1, . . . , τ (v)k + ik). A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 3 It follows (see §4.1) that a is a type-rotating automorphism of ∆ if and only if it is a Cartesian will then be of index at most 2 in Aut(∆). This follows from the facts that [D8 : C2 × C2] = [D6 : C3] = 2, where Cn is the cyclic group and Dn is the dihedral group of order n. An apartment in ∆ is a chamber subcomplex of ∆ isomorphic to the corresponding Coxeter complex. Thus an product of type-rotating automorphisms of the k factors. Recall that an eA1 building is a tree and note that any automorphism of an eA1 building is type-rotating. Henceforth, let ∆ be an eA1 × eA1 or eA2 building. The subgroup of type-rotating automorphisms apartment in ∆ is a plane tessellated by squares in the eA1 × eA1 case and by equilateral triangles in the eA2 case. Denote by V∆ the vertex set of ∆. Any two vertices u, v ∈ V∆ belong to a common apartment. The convex hull, in the sense of buildings, between two vertices u and v is depicted in Figure 1, with Fundamental properties of buildings imply that any apartment various degeneracies possible. n n ........................... m ........................... ........................... •v ................................................................................................................................ ..................................................................................................................................................................................................................................... ..................................................................................................................................................................................................................................... ................................................................................................................................ ........................... •u eA1 × eA1 case m v • ........................... n m ........................................................................................................................................................................................................................................................................................................................ .......................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................ .......................................................................................................................................................................... eA2 case ........................... ........................... ........................... •u m n Figure 1. Convex hull of two vertices. containing u and v must also contain their convex hull. Define the distance, d(u, v), between u and v to be the graph theoretic distance on the one- skeleton of ∆. Any path from u to v of length d(u, v) lies in their convex hull, and the union of the vertices in such paths is exactly the set of vertices in the convex hull. Note that although this chambers. We define the shape σ(u, v) of the ordered pair of vertices (u, v) ∈ V∆ × V∆ to be the pair in Figure 1 point in the direction of cyclically increasing type, i.e. {. . . , 0, 1, 2, 0, 1, . . .}. In the last statement is true for eA1 × eA1 and eA2 buildings, it does not hold for arbitrary affine buildings. For example in a eG2 building it fails for two vertices whose convex hull contains at least three (m, n) ∈ N × N as indicated in Figure 1. Note that d(u, v) = m + n. In the eA2 case the arrows eA1 × eA1 case the components of σ(u, v) indicate the relative contributions to d(u, v) from the two eA1 factors. More specifically, suppose ∆ = ∆1 × ∆2 where ∆1 and ∆2 are eA1 buildings, i.e. trees. If v = (v1, v2) and w = (w1, w2) are vertices of ∆, the shape from v to w is σ(v, w) = (d(v1, w1), d(v2, w2)) where d denotes the usual graph-theoretic distance on a tree. An edge in ∆ connects the vertices v and w if σ(v, w) = (0, 1) or σ(v, w) = (1, 0). Lemma 1.4. Suppose m1, m2, n1, n2 ∈ N and σ(u, w) = (m1 + m2, n1 + n2) for vertices u, w ∈ V∆. Then there is a unique vertex v ∈ V∆ such that σ(u, v) = (m1, n1) and σ(v, w) = (m2, n2). 4 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER Proof. Such a v ∈ V∆ satisfies d(u, w) = d(u, v) + d(v, w) so it must lie in the convex hull of u and w. Inside the convex hull existence and uniqueness of v are clear. (cid:3) It is a direct consequence of the definitions that every type-rotating automorphism a ∈ Aut(∆) preserves shape in the sense that σ(au, av) = σ(u, v) for all u, v ∈ V∆. Furthermore, if b ∈ Aut(∆) is not type-rotating then σ(bu, bv) = (n, m) whenever σ(u, v) = (m, n). Thus an element a ∈ Aut(∆) preserves shape if and only it is type-rotating. Thus the group AutS(∆) of shape-preserving automorphisms coincides with the group of type-rotating automorphisms. We refer to Section 4 for a more detailed analysis of the interesting special case where the group Γ ≤ AutS(∆) acts simply transitively on V∆. 1.3. Free actions and groupoids. Suppose now that Γ ≤ AutS(∆) acts freely, but not neces- sarily transitively, on the vertices V∆ of ∆. This induces an action of Γ on V∆ × V∆. Denote by [u, v] the orbit of the ordered pair (u, v) ∈ V∆ × V∆ under the left action of Γ. Define the set of Γ-orbits of V∆ × V∆. Given [u, v] and [v, w] ∈ Γ′, define their product as Γ′ = Γ\(V∆ × V∆), [u, v][v, w] = [u, w]. The product of [u, v] and [x, y] is not always defined. It exists only if v and x lie in the same Γ-orbit, and in that case it is well defined because Γ acts freely on V∆. Specifically, if bx = v for b ∈ Γ, then [u, v][x, y] = [u, v][bx, by] = [u, by]. The set Γ′ with this product satisfies the axioms of a groupoid, see [C, Chapter II.5]: Definition 1.5. A groupoid consists of a set G, a distinguished subset G(0) ⊂ G, two maps s, r : G −→ G(0) and a law of composition (α1, α2) ∈ G(2) 7→ α1α2 ∈ G, with domain G(2) = {(α1, α2) ∈ G × G : s(α1) = r(α2)}, such that • s(α1α2) = s(α2), r(α1α2) = r(α1), for all (α1, α2) ∈ G(2); • s(α) = r(α) = α , for all α ∈ G(0); • αs(α) = α , r(α)α = α, for all α ∈ G; • (α1α2)α3 = α1(α2α3); • each α has a two-sided inverse α−1, with αα−1 = r(α), α−1α = s(α). The maps r and s are called the range and source maps. An element ι ∈ G(0) is called a unit. The units in Γ′ are of the form [v, v] for v ∈ V∆. If α = [u, v] ∈ Γ′ then r(α) = [u, u], s(α) = [v, v] and α−1 = [v, u]. Definition 1.6. Let G be a groupoid and X be a set together with a surjection s′ : X → G(0). Form the fibred product X ∗ G = {(v, α) ∈ X × G s′(v) = r(α)} A (right) groupoid action of G on X is a map (v, α) 7→ vα ∈ X from X ∗ G to X such that • s′(vα) = s(α); • v(αβ) = (vα)β, whenever (v, α) ∈ X ∗ G and (α, β) ∈ G(2); • vs′(v) = v for all v ∈ X. We think of s′ as being a 'generalized source' map for the action. A groupoid action is called simply transitive if in addition • given v, w ∈ X there exists a unique α ∈ G such that (v, α) ∈ X ∗ G and vα = w. We refer the reader to [Ren] and [MW, Section 2] for details on groupoid actions. When G acts simply transitively on X and v, w ∈ X, we write v−1w for the unique α ∈ G such that vα = w. This notation is possibly misleading since v−1 has no independent existence. However, if (v, β), (w, γ) ∈ X ∗ G then (vβ)−1(wγ) = β −1(v−1w)γ A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 5 as is easily checked. We define a simply transitive groupoid action of Γ′ on V∆ via v[v, w] = w. Thus s′(v) = [v, v] and (u, [v, w]) ∈ X∗G exactly when u and v lie in the same Γ-orbit. If v, w ∈ V∆ then vα = w if and only if α = [v, w] so that v−1w = [v, w]. Definition 1.7. Let ∆ be an eA1 × eA1 or eA2 building, V∆ its vertex set and σ∆ its shape function. Let G be a groupoid acting on V∆ on the right. A shape function on G compatible with the action is a map σG : G → N × N such that σ∆(v, vα) = σG(α) whenever v ∈ V∆, α ∈ G and vα is defined. Henceforth we will omit the subscripts and denote all shape functions by σ. A shape function on Γ′ compatible with the action of Γ′ on ∆ is given by σ([u, v]) = σ(u, v) for any u, v ∈ V∆. This is well-defined since Γ ≤ AutS(∆) acts on ∆ by shape-preserving auto- morphisms. Lemma 1.8. Let ∆ be an eA1 × eA1 or eA2 building. Let G be a groupoid endowed with a simply transitive action on V∆ and a compatible shape function. Let m1, m2, n1, n2 ∈ N. Whenever α ∈ G has shape (m1 + m2, n1 + n2) there exist unique β, γ ∈ G such that α = βγ, σ(β) = (m1, n1) and σ(γ) = (m2, n2). Proof. Fix u ∈ V∆ such that w = uα exists. A pair (β, γ) satisfies the required conditions if and only if σ(β) = (m1, n1), σ(β −1α) = (m2, n2) and γ = β −1α. This happens if and only if v = uβ satisfies σ(u, v) = (m1, n1), and σ(v, w) = (m2, n2). By Lemma 1.4, there is exactly one such vertex v. Hence the pair (β, γ) = (u−1v, v−1w) uniquely satisfies the required conditions. (cid:3) To summarize, any group Γ ≤ AutS(∆) with a free left action on V∆ gives rise to a groupoid Γ′ with a simply transitive (right) groupoid action on V∆ and a compatible shape function. The actions of Γ and Γ′ commute in the sense that c(vα) = (cv)α for c ∈ Γ, v ∈ V∆ and α ∈ Γ′ provided one side or the other exist. We will call Γ′ the commutant groupoid of Γ. Note that the left action of Γ on V∆ preserves the building structure and that the right action of Γ′ on V∆ does not. If, as in §4, Γ acts simply transitively on V∆, then Γ′ is a group isomorphic to Γ and the groupoid action of Γ′ on V∆ constructed in the above manner is equivalent to a right group action of Γ on V∆. For any fixed v0 ∈ V∆, the map c 7→ [v0, cv0] gives an isomorphism Γ → Γ′. This isomorphism depends on the choice of v0, varying up to inner automorphisms of Γ. Our results will be phrased in the language of groupoid actions with implications for free actions of groups on V∆ via the above construction. 6 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER 1.4. Convolutions. If G is a groupoid, let CG denote the space of finitely supported complex valued functions on G. The convolution product on CG defined by (g1 ∗ g2)(γ) = Xαβ=γ g1(α)g2(β) makes CG into an associative algebra. Suppose that G acts simply transitively on a set X. For g ∈ CG, define ρ(g) : ℓ2(X) → ℓ2(X) by (ρ(g)f )(v) = (f ∗ g)(v) = Xu∈X uα=v f (u)g(α). It follows from the simple transitivity of the action of G that ρ(g)f actually lies in ℓ2(X); indeed that kf ∗ gk2 ≤ kfk2kgk1. The map ρ defines a right action of CG on ℓ2(X). Let ∆ be an eA1 × eA1 or eA2 building and let p(m, n) =((m + 1)(n + 1) (1) In Sections 2 and 3 we prove the following result: 1/2(m + 1)(n + 1)(m + n + 2)pmax(m, n) + 1 (eA1 × eA1 case) (eA2 case). Theorem 1.9. Suppose G is a groupoid acting simply transitively on V∆ and σ : G → N is a shape function compatible with the action. Fix any (m, n) ∈ N × N. If g ∈ CG is supported on elements of shape (m, n), then kρ(g)k ≤ p(m, n)kgk2. Equivalently, the conclusion asserts that We can and will restrict to nonnegative and finitely supported f and g when proving this. kf ∗ gk2 ≤ p(m, n)kfk2kgk2. If Γ is a discrete group, we define convolution between f ∈ ℓ2(Γ) and g ∈ CΓ in the usual way: Theorem 1.1 is a consequence of Theorem 1.9. (f ∗ g)(c) =Xab=c f (a)g(b). Proof of Theorem 1.1. Let Γ′ be the commutant groupoid of Γ, as defined in §1.3. Given f ∈ ℓ2(Γ) and g ∈ CΓ, define f ′ ∈ ℓ2(V∆) and g′ ∈ CΓ′ by f ′(cv0) = f (c) f ′(w) = 0 g′([dv0, cv0]) = g(d−1c) g′([w1, w2]) = 0 for c ∈ Γ unless w ∈ Γv0 for c, d ∈ Γ unless w1, w2 ∈ Γv0. It is then immediate that f ′ ∗ g′ is related to f ∗ g via (f ′ ∗ g′)(cv0) = (f ∗ g)(c) (f ′ ∗ g′)(w) = 0 for c ∈ Γ unless w ∈ Γv0. Moreover, g′ is supported on elements of Γ′ of shape (m, n), kg′k2 = kgk2, and kf ′k2 = kfk2. Now apply Theorem 1.9 to the groupoid Γ′ and the functions f ′ and g′ to complete the proof of Λ Theorem 1.1. Recall from [J] that a group G has property (RD) if there is a length function L on G such that any function on G which is rapidly decreasing relative to L belongs to the reduced C ∗-algebra of G. Corollary 1.10. Any group Γ ≤ Aut(∆) which acts freely on V∆ has property (RD). A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 7 Proof. The index of AutS(∆) in Aut(∆) is at most 2. Hence the index of Γ ∩ AutS(∆) in Γ is at most 2. As property (RD) holds for a group whenever it holds for some subgroup of finite index [J, Proposition 2.1.5], we may assume that Γ ≤ AutS(∆). Fix v0 ∈ V∆ and define σ(c) = σ(v0, cv0) for c ∈ Γ. Likewise, define c = d(v0, cv0). Then c is a length function on Γ, in the sense of [J]. When σ(c) = (m, n) one has c = m + n. To derive property (RD) from Theorem 1.1 we note that a given value of c corresponds to only c+1 different shapes, and that the relevant function p(m, n) from (1) can be bounded by a polynomial in m + n. Now argue as in [C, Chapter III.5, Theorem 5] or [H, Lemma 1.5]. (cid:3) Remark 1.11. Corollary 1.10 implies property (RD) for any discrete subgroup of Aut(∆) which has a torsion-free subgroup of finite index. As a consequence, if F is a finite extension of Qp and Γ is a finitely generated discrete subgroup of SL3(F), then it follows from Selberg's Lemma [Sel] that Γ has a torsion free finite index subgroup, and so satisfies property (RD). 1.5. The retraction centred at a boundary point of the building. The proof of Theorem 1.9 relies heavily on the notion of the retraction of a building centred at a boundary point. We outline a construction based on [Br, page 171]. There is also a concise description in [R, §9.3]. An infinite straight line of edges in an apartment is called a wall. Fix an apartment A of ∆. Consider two half-planes in A bounded by intersecting walls and measure the angle between those two walls through the intersection of the two half-planes. If that angle is nonzero and minimal, we call the intersection of the two half-planes a sector. A sector in ∆ is by definition a sector in some apartment of ∆. The boundary, Ω, of ∆ is the set of sectors in ∆ under the equivalence relation that S ∼ S ′ if S ∩ S ′ contains a sector. (This is the analogue in higher rank of the notion of tail equivalence for semi-infinite paths in trees.) Given a vertex v ∈ V∆ and a boundary point ω ∈ Ω there is a unique sector based at v representing ω, denoted Sv(ω). Thus Ω can be identified with the set of sectors emanating from any fixed vertex in ∆. Fix an apartment A0 of ∆ and a sector S0 in A0 representing a boundary point ω0 ∈ Ω. Then ∆ is a union of apartments A′ each of which contains a subsector of S0. There is a retraction r : ∆ −→ A0 which is a contraction, in the sense that d(r(u), r(v)) ≤ d(u, v) for any u, v ∈ V∆, and whose restriction to any such A′ is an isomorphism from A′ onto A0 which fixes A′ ∩ A0 pointwise. In particular, given any v ∈ V∆, r(v) is independent of the apartment A′ chosen such that v ∈ A′. The boundary point ω0 defines a preferred direction in A0 which we shall call up. Each chamber in A0 may therefore be labelled with an arrow pointing up, as in Figure 2. This labelling of the chambers of A0 induces a labelling of all chambers of ∆ via the retraction r. We say a line segment is horizontal if it is perpendicular to the up direction. In particular, a horizontal line segment can not be part of a sector wall for any sector representing ω0. It is helpful to imagine that ∆ has been folded flat over A0 according to the retraction r so as to hang down from ω0. Note that, as in Figure 2, each edge of the apartment A0 is incident with one chamber which lies above it and one chamber which lies below it. Given any edge (u, v) of ∆, there is a unique chamber of ∆ containing (u, v) which retracts to the chamber lying above (r(u), r(v)) and all other chambers containing (u, v) retract to the chamber of A0 lying below (r(u), r(v)). Suppose that C and C ′ are adjacent chambers of ∆. Then either r(C) is adjacent to but distinct from r(C ′), or r(C) equals r(C ′). In the first case the arrows on the two chambers are parallel and in the second case they mirror each other through the common edge and point in converging directions. If A is an apartment of ∆, then the retraction r : A → A0 need not be injective. The above constraints then lead to overall labellings of its chambers as in Figure 3, with various degenerate cases also possible. This pattern of arrows is called the folding diagram for A, as it indicates how A is folded by r. The folding diagram is a convenient visual convention for In particular, two apartments have the same encoding geometric properties of the retraction. folding diagram if their retractions are equivalent up to translation in A0. One may similarly 8 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER ω0 ω0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ....... ................................................................................................................................................................................................................................................................................................................................................................................................................... ................................................................................................................................................................................................................................................................................................................................................................................................................... .............................................................................................................................................................................................................................................................................................................. .............................................................................................................................................................................................................................................................................................................. .............................................................................................................................................................................................................................................................................................................. .............................................................................................................................................................................................................................................................................................................. ............................. • ............................. ............................. ............................. ....... ....... ....... . . . . . . . . eA1 × eA1 case ....... ............................... .............................................................................................................................................................................................................................................................................................................. .............................................................................................................................................................................................................................................................................................................. ....... ............................... ............................... ....... ............................... ....... eA1 × eA1 case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ........................ ........................ ........................ ........................ ........................ ........................ ........................ ........................ ........................ ........................ ........................ • ........................ ........................ .......................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................... .......................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................................................................................................................................... .............................................................................................................................................................................................................................................................................................. .............................................................................................................................................................................................................................................................................................. eA2 case .................................. ........................... ........................ ........................ ....... ....... ....... ........................... ....... ....... ........................... eA2 case Figure 2. Arrows indicating preferred direction. Figure 3. Generic labelling of apartments. discuss the folding diagram of any connected subset of an apartment A. Note that in certain degenerate cases it becomes necessary for us to consider the folding diagram of a line segment in ∆ where there are no chambers for reference. The folding obtained in such a case will be a piecewise straight line and we trust that the reader will have little difficulty in appreciating the possibilities. For clarity of exposition we will now consider the eA1 × eA1 and eA2 cases separately. Although the ideas used to tackle the eA1 × eA1 case are also used in the proof of the eA2 case, the extra complexity of the eA2 case would make a joint exposition unwieldy. Proof of Theorem 1.9 in the eA1 × eA1 case. Let Wm,n = {γ ∈ G σ(γ) = (m, n)} and fix an element g ∈ CG whose support is contained in Wm,n, i.e. g ∈ CWm,n. The matrix coefficients of ρ(g) are given by hρ(g)δx, δyi = g(x−1y) for x, y ∈ V∆ where x−1y is defined as in §1.3. Thus, 2. The eA1 × eA1 Case hρ(g)δx, δyi =(g(x−1y) 0 if x−1y ∈ Wm,n otherwise. Geometrically, hρ(g)δx, δyi is non-zero only if x and y are opposite vertices of a rectangle Pm,n(x, y) of the sort pictured in Figure 4. This rectangle is the convex hull of x and y, and as such is common to all apartments containing both x and y. Consider the folding diagram of Pm,n(x, y) induced by the retraction r. In general this will be Let z be the apex of the as indicated in Figure 5, although certain degeneracies are possible. A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 9 n m y • ................................................................................................................................ ........................... ..................................................................................................................................................................................................................................... ..................................................................................................................................................................................................................................... ................................................................................................................................ ........................... •x ........................... ........................... m n Figure 4. Convex hull of two vertices in ∆. y • .............................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................. .......................................................................................................................................................................................................... ......................... ..................................................................................................... •z ......................... ...................... ....... ....... ...................... j i ....... ....... •x Figure 5. General labelling of a rectangle in ∆. upward labelled subrectangle based at x such that x−1z ∈ Wi,j with i and j maximal, as in the diagram. We call z the focal point of the folding. Associated with each rectangle Pm,n(x, y) is an abstract diagram D(x, y) = Di,j, which is a copy of Pm,n(x, y) in which the labels of the vertices are forgotten, but the arrows are retained. Denote by D(m, n) the set of all possible diagrams Di,j for fixed m, n ∈ N. Lemma 2.1. #D(m, n) = (m + 1)(n + 1). Proof. In order to determine the number of possible diagrams it is sufficient to enumerate the possible locations of the focal point of the folding. Hence, there are (m + 1)(n + 1) possible diagrams. (cid:3) For each D ∈ D(m, n), define an operator TD on ℓ2(V∆) by hTDδx, δyi =(g(x−1y) 0 if D(x, y) = D otherwise. Then ρ(g) = XD∈D(m,n) (2) TD and so kρ(g)k ≤ XD∈D(m,n) kTDk. We now fix D = Di,j ∈ D(m, n) and proceed to estimate kTDk. For each z ∈ V∆ define Hz = hδx x−1z ∈ Wi,j and Sz(ω0) ⊆ Sx(ω0)i and Kz = hδy z−1y ∈ Wm−i,n−j and Sz(ω0) ⊆ Sy(ω0)i. These give rise to two decompositions; ℓ2(V∆) = ⊕z∈V∆Hz = ⊕z∈V∆Kz. Define an operator Tz,z : Hz −→ Kz by hTz,zδx, δyi = g(x−1y) for x ∈ Hz and y ∈ Kz, and define Tz ′,z : Hz −→ Kz ′ to be zero for z′ 6= z. Then TD can be expressed as a block diagonal operator matrix TD = [Tz ′,z] = ⊕z∈V∆Tz,z and it is sufficient to show that for fixed z ∈ V∆, kTz,zk ≤ kgk2. 10 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER So fix z and suppose that δx ∈ Hz and δy ∈ Kz. According to Lemma 1.8, x and y are uniquely determined by x−1y. So each γ ∈ Wm,n contributes at most one matrix coefficient of Tz,z in the form g(γ). Therefore we have kTz,zk ≤ kTz,zkHS ≤ kgk2. Λ This concludes our proof of Theorem 1.9 in the eA1 × eA1 case. Proof of Theorem 1.9 in the eA2 case. Let Wm,n = {γ ∈ G : σ(γ) = (m, n)} and fix an element g ∈ CG whose support is contained in Wm,n, i.e. g ∈ CWm,n. The matrix coefficients of ρ(g) are given by hρ(g)δx, δyi = g(x−1y) for x, y ∈ V∆. Thus, 3. The eA2 Case hρ(g)δx, δyi =(g(x−1y) 0 if x−1y ∈ Wm,n otherwise. Geometrically, hρ(g)δx, δyi is non-zero only if x and y are opposite vertices of a parallelogram Pm,n(x, y) of the sort pictured in Figure 6. This parallelogram is the convex hull of x and y, and y • ........................... n m .......................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................................................ ........................................................................................................................................................................................................................................................................................................................ .......................................................................................................................................................................... ........................... ........................... m n ........................... •x Figure 6. Convex hull of two vertices in ∆. as such is common to all apartments containing both x and y. Note that x and y are at the acute angles of the parallelogram. Consider the folding of Pm,n(x, y) induced by the retraction r. This will be one of the possibilities depicted in Figure 7, with possible degeneracies. Let zx be the apex of the upward labelled region y • ....... ........................... ....... ........................... .......................... ........................................................................ ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ............................................................................................................................................................................................................................................................................................. ........................................................................ .......................... ........................... .................................. ....... ....... ....... •x y • ....... .......................... ............................................................................................................................. ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ................................................................................................................................................................................... ........................... ....... ...................................................... ................. ....... ................. ....... .................................. .......................... ....... •x Figure 7. Possible labellings of parallelograms. based at x such that x−1zx ∈ Wi,j with i and i + j maximal. Similarly, let zy be the apex of the A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 11 upward labelled region based at y such that y−1zy ∈ Wk,l with k and k + l maximal. In terms of our diagrams, zx and zy would be as labelled in Figure 8. y • ....... l k ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ............................................................................................................................................................................................................................................................................................. .......................... ........................................................................ • zy •zx ........................................................................ .......................... j i ....... •x y • ....... l k .......................... ............................................................................................................................. ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ................................................................................................................................................................................... zy • ...................................................... • zx .......................... j i ....... •x Figure 8. Positions of the vertices zx and zy. We shall refer to zx and zy as the focal points of the folding. If zx 6= zy, the convex hull of zx and zy is a line segment which we denote zx ↔ zy and the retraction r maps zx ↔ zy injectively to a horizontal line segment r(zx ↔ zy) in A0. Note that the positions of the focal points completely determine the folding. Among possible degeneracies, we could have zx = x and zy = y, or zx = zy. i,j , which is a copy of Pm,n(x, y) in which the labels of the vertices are forgotten, but the arrows are retained as in Figure 9. Associated to each parallelogram Pm,n(x, y) is an abstract diagram D(x, y) = Dk,l ....... l k ........................... ....... ........................... .......................... ........................................................................ ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ............................................................................................................................................................................................................................................................................................. ........................................................................ .......................... ........................... .................................. j i ....... ....... ....... Figure 9. The diagram Dk,l i,j . Denote by D(m, n) the set of all possible diagrams Dk,l i,j for fixed m, n ∈ N. We divide the family D(m, n) into two classes as follows. Let D′(m, n) be the set of diagrams associated to parallelograms Pm,n(x, y) in which x−1zx ∈ Wp,q and x−1zy ∈ Ws,t where p + q = s + t. Thus a diagram D ∈ D(m, n) is in D′(m, n) if it is of the form depicted in Figure 10 or if the focal points are coincident. Let D′′(m, n) = D(m, n) \ D′(m, n). So D ∈ D(m, n) is in D′′(m, n) if it is of one of the forms depicted in Figure 11 with distinct focal points. Lemma 3.1. #D′′(m, n) = 1/2(m + 1)(n + 1)(m + n). Proof. In order to determine the number of diagrams in D′′(m, n) it is sufficient to enumerate the possible positions of the focal points in the foldings. There are (m + 1)(n + 1) choices for the first focal point. Given one focal point, the possible positions of the other are indicated in Figure 12. So there are m + n positions the second focal point could occupy. Since the order of choice is irrelevant we must divide the total number of possibilities by a factor of 2. Hence the result. (cid:3) 12 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER Figure 10. Parallelograms in D′(m, n). ....... .......................... ....... ....... .......................... ........................... ....... .......................... ........................................................................ ........................... ....... ...................................................... ....... ................. ................. ....... .................................. ................................................................................................................................................................................... ............................................................................................................................. ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................ ............................................................................................................................................................................................................................................................................................. ............................................................................................................................................................................................................................................................................................. .......................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................................................................ .......................... .......................... ........................................................................ ........................................................................ .......................... ........................... ....... ....... ........................... ........................... ....... .................................. ....... ....... ........................... ........................... ....... .................................. • ....... ....... Figure 11. Parallelograms in D′′(m, n). Figure 12. Possible positions of second focal point. For each D ∈ D(m, n), define an operator TD on ℓ2(V∆) by hTDδx, δyi =(g(x−1y) TD + XD∈D′′(m,n) 0 TD = XD∈D′(m,n) if D(x, y) = D otherwise. TD. Thus ρ(g) = XD∈D(m,n) For each diagram D ∈ D′(m, n) choose one of the focal points and call it a distinguished focal i,j(m, n) point. There are (m + 1)(n + 1) possible positions for the distinguished focal point. Let D′ be the set of diagrams in D′(m, n) whose distinguished focal point is in position (i, j), so that XD∈D′(m,n) TD = mXi=0 nXj=0 XD∈D′ i,j(m,n) TD. Hence and so (3) ≤ kgk2. i,j(m,n) TD(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) XD∈D′ TD(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) XD∈D′(m,n) ≤ (m + 1)(n + 1)kgk2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kρ(g)k ≤ (m + 1)(n + 1)kgk2 + XD∈D′′(m,n) kTDk. A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 13 An argument analogous to that used in §2 shows that We proceed to show that, for D ∈ D′′(m, n), kTDk ≤pmax(m, n) + 1kgk2. Since (m + 1)(n + 1) + (m + 1)(n + 1)(m + n) = (m + 1)(n + 1)(m + n + 2) 1 2 1 2 we will then have established that kρ(g)k ≤ 1/2(m + 1)(n + 1)(m + n + 2)pmax(m, n) + 1 kgk2. We now fix D = Dk,l x zy ∈ Ws,t where (s, t) = (p, 0) or (0, q), and (s, t) is determined by D. Thus i, j, k, l, s, t are all now fixed. i,j ∈ D′′(m, n) and estimate kTDk. Note that, if D(x, y) = D, then z−1 Given z ∈ V∆, define and Hz = hδx x−1z ∈ Wi,j and Sz(ω0) ⊆ Sx(ω0)i Kz = hδy y−1z ∈ Wk,l and Sz(ω0) ⊆ Sy(ω0)i. These give rise to two decompositions; ℓ2(V∆) = ⊕z∈V∆Hz = ⊕z∈V∆Kz. define Tz2,z1 : Hz1 −→ Kz2 by If (z1, z2) ∈ V∆ × V∆ with z−1 hTz2,z1δx, δyi = g(x−1y) for δx ∈ Hz1 and δy ∈ Kz2, 1 z2 ∈ Ws,t and r maps z1 ↔ z2 injectively to a horizontal line, and define Tz2,z1 : Hz1 −→ Kz2 to be zero for other pairs (z1, z2) ∈ V∆ × V∆. Then TD can be expressed as an operator matrix TD = [Tz2,z1] and therefore kTDk ≤ k [kTz2,z1k ] k (4) by Lemma 1.2. We proceed to estimate kTz2,z1k. 1 z2 ∈ Ws,t and such that r maps z1 ↔ z2 injectively to a horizontal line. Thus r(z1 ↔ z2) is a horizontal line of shape (s, t). The vectors δx ∈ Hz1, δy ∈ Kz2 are uniquely determined by x−1y ∈ G according to Lemma 1.8. Thus each γ ∈ Wm,n contributes Fix (z1, z2) ∈ V∆ × V∆ with z−1 at most one matrix coefficient of Tz2,z1 in the form g(γ). Defineeg ∈ CWs,t by 1/2 eg(ζ) = Note that kegk2 = kgk2 and (5) αζβ∈Wm,n  g(αζβ)2  Xα∈Wi,j,β∈Wk,l kTz2,z1k ≤ kTz2,z1kHS ≤eg(z−1 0 1 z2). if ζ ∈ Ws,t otherwise. 14 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER 1 z2) Define a new operator eTD on ℓ2(V∆) by heTDδz1, δz2i =(eg(z−1 By (4), (5) and Lemma 1.3 we have kTDk ≤ keTDk. Since we also have kegk2 = kgk2, it will be 1 z2 ∈ Ws,t and r(z1 ↔ z2) is a horizontal line of shape (s, t), if z−1 otherwise. sufficient to prove 0 Recall that (s, t) = (p, 0) with 0 ≤ p ≤ m or (0, q) with 0 ≤ p ≤ n. We suppose the former and keTDk ≤pmax(m, n) + 1 kegk2. keTDk ≤pp + 1 kegk2. (6) (7) prove that We must prove (8) In order to prove (7), we simplify our notation. Let g ∈ CWp,0 where 0 ≤ p ≤ m and define an A similar argument in the case (s, t) = (0, q) gives keTDk ≤ √q + 1 kegk2 thus establishing (6). hT δx, δyi =(g(x−1y) if x−1y ∈ Wp,0 and r(x ↔ y) is a horizontal line of shape (p, 0), otherwise. operator T on ℓ2(V∆) by 0 kTk ≤pp + 1 kgk2. Given (x, y) ∈ V∆ × V∆ with x−1y ∈ Wp,0 and such that r(x ↔ y) is a horizontal line of shape (p, 0), there is a unique z = z(x, y) ∈ V∆ such that the convex hull of x, y and z is labelled as in Figure 13 and such that z−1x, x−1y, y−1z ∈ Wp,0. That is xyz is an equilateral triangle, with side of length p, base xy and pointing to ω0. To see this, suppose that x1 is the vertex adjacent to, z • ........................... ....... .............................................................................................................................................................................................................................................................................................................................................................................................................................................•x •y Figure 13. Labelling of convex hull of x, y and z = z(x, y). but distinct from, x in x ↔ y. Let w be the third vertex of the unique chamber containing (x, x1) and retracting above (r(x), r(x1)). The convex hull of x, y and w is a trapezoidal strip in ∆ which has x ↔ y as one of its bases, contains the chamber with vertices (x, x1, w) and whose other base, w ↔ w′, has length p− 1 as in Figure 14. Since the retraction is contractive, r must map w ↔ w′ w′ w • • ........................................................................ ................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................... ...................... ....... •x •x1 •y ...................... ....... Figure 14. Trapezoidal strip retracting above x ↔ y. injectively to a horizontal line lying above r(x) ↔ r(y). By induction, this argument demonstrates the existence of the triangle depicted in Figure 13. Note that Sz(ω0) = Sx(ω0) ∩ Sy(ω0). For later convenience we define a set C = {(x, y, z) ∈ V∆ × V∆ × V∆ x−1y ∈ Wp,0, r (x ↔ y) is a horizontal line of shape (p, 0), and Sz(ω0) = Sx(ω0) ∩ Sy(ω0)} . A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 15 For each z ∈ V∆, define and Hz = hδx z−1x ∈ Wp,0 and Sz(ω0) ⊆ Sx(ω0)i Kz = hδy y−1z ∈ Wp,0 and Sz(ω0) ⊆ Sy(ω0)i. Thus δx ∈ Hz if z = z(x, y) for some y ∈ V∆ and a similar condition characterizes the elements δy ∈ Kz. Let Tz ′,z : Hz −→ Kz ′ be defined by hTz ′,zδx, δyi = g(x−1y) if δx ∈ Hz and δy ∈ Kz ′. Note that ℓ2(V∆) = ⊕z∈V∆Hz = ⊕z ′∈V∆Kz ′. Since Tz ′,z = 0 unless z′ = z, T can be expressed as a block diagonal matrix T = [Tz ′,z] = ⊕z∈V∆Tz,z. Thus it is sufficient to show that, for fixed z ∈ V∆, (9) We now fix z and show that for f1 ∈ Hz, and f2 ∈ Kz. We may assume f1, f2, g ≥ 0. Then kTz,zk ≤pp + 1 kgk2. hTz,zf1, f2i ≤pp + 1 kf1k2 kf2k2 kgk2 hTz,zf1, f2i = Xx,y∈V∆ f1(x)f2(y)g(x−1y). (x,y,z)∈C We define g1, g2 ∈ CG via g1(α) =(f1(x) 0 if α = z−1x otherwise and g2(β) =(f2(y) 0 if β = y−1z otherwise and change our emphasis to obtain hTz,zf1, f2i = Xx,y∈V∆ (x,y,z)∈C g1(z−1x)g2(y−1z)g(x−1y). (10) We define a set Tp = {(α, β, γ) ∈ G × G × G α, β, γ ∈ Wp,0 and γβα is a unit} It can be shown that if x, y, z are vertices in an eA2 building and σ(x, y) = σ(y, z) = σ(z, x) = (p, 0), the convex hull of {x, y, z} is necessarily an equilateral triangle of side length p in some apartment. Hence, given any (α, β, γ) ∈ Tp and any v ∈ V∆ such that vγ exists, the vertices v, vγ, vα−1 lie in a common apartment and their convex hull is an equilateral triangle of side length p. Use Tp to rewrite equation (10) as where efi ∈ CWp,0 for i = 1, 2, 3. This is equivalent to having changed the emphasis z • ................ from β α to ............................................................................................................................................................... ............................................................................................................................................................... hTz,zf1, f2i = X(α,β,γ)∈Tp ef1(α)ef2(β)ef3(γ), .............................................................................................................................................................................................................................................................................................................................................................................................................................................•x f1(α)f2(β)f3(γ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ................ ................ ............................................................................................................................................................... •y . γ X(α,β,γ)∈Tp (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤pp + 1 kf1k2 kf2k2 kf3k2. We now complete the proof of (9) by proving the following result. Lemma 3.2. If f1, f2, f3 ∈ CWp,0 then 16 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER Proof. Assume that f1, f2, f3 ≥ 0. For a fixed f3, define T3 : CWp,0 −→ CWp,0 by hT3δα, δβi = Xγ:(α,β,γ)∈Tp f3(γ) where the right hand side has at most one non-zero term, and we adopt the convention that it is zero if the set {γ ∈ Wp,0 (α, β, γ) ∈ Tp} is empty. We will use this convention again later. Thus f1(α)f2(β)f3(γ) = hT3f1, f2i. We must show that kT3k ≤ √p + 1 kf3k2, or equivalently that kT ∗ X(α,β,γ)∈Tp 3 T3k ≤ (p + 1) kf3k2 2. Now, hT ∗ 3 T3δα1, δα2i = hT3δα1, T3δα2i = Xβ∈G =  Xγ1:(α1,β,γ1)∈Tp Xβ,γ1,γ2∈G (α1,β,γ1),(α2,β,γ2)∈Tp f3(γ1) Xγ2:(α2,β,γ2)∈Tp f3(γ2) f3(γ1)f3(γ2). We say that α1, α2 ∈ G share precisely j initial letters if • there exists a z ∈ V∆ such that zα1 and zα2 are both defined, and • there exist ζ ∈ Wj,0, and eα1,eα2 ∈ Wp−j,0 such that α1 = ζeα1, α2 = ζeα2, and σ(eα−1 1 eα2) = (p − j, p − j). Note that if α1, α2 ∈ Wp,0, and if α−1 for some 0 ≤ j ≤ p. We have a decomposition T ∗ 3 T3 = S0 + · · · + Sp where 1 α2 is defined, then α1 and α2 must share precisely j letters hSjδα1, δα2i =(hT ∗ 0 3 T3δα1, δα2i if α1, α2 share precisely j initial letters, otherwise. It is therefore sufficient to prove that kSjk ≤ kf3k2 hSjδα1, δα2i 6= 0 only if there are diagrams 2 for 0 ≤ j ≤ p. Diagrammatically speaking, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ................ . ξ ................ ............................................................................................................................................................... ............................................................................................................................................................... ............................................................................................................................................................... eγ1 eβ ............................................................................................................................................................................... eα1 ............................................................................................................................................................................... ............................................................................................................................................................................................................................................................................................................. ζ ................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ................ β and β . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ξ ............................................................................................................................................................................................................................................................................................................. ............................................................................................................................................................................... ................ ............................................................................................................................................................................... ............................................................................................................................................................... eγ2 eβ ................ ............................................................................................................................................................... eα2 ............................................................................................................................................................... ζ ................ ................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . whose common (shaded) sections are identical. For a fixed ζ ∈ Wj,0, let Hζ = hδα α ∈ Wp,0, α = ζeα,eα ∈ Wp−j,0i. If δα ∈ Hζ then hSjδα, δα′i = 0 unless δα′ ∈ Hζ. Therefore SjHζ ⊆ Hζ and it follows that Sj = ⊕ζ∈Wj,0Sζ j : Hζ −→ Hζ is the restriction of Sj to Hζ. So it is enough to bound Sζ j , where Sζ j A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 17 for each ζ ∈ Wj,0. However, in this case eα−1 1 2 2 kSζ f3(γ1)f3(γ2)  α1=ζ eα1 α2=ζ eα2 eα2∈Wp−j,p−j Xβ,γ1,γ2∈Wp,0 (α1,β,γ1)∈Tp (α2,β,γ2)∈Tp HS = Xα1,α2∈Γ jk2 ≤ kSζ jk2   eαi, eβ, andeγi to enumerate the larger ones formed by αi, β, and γi. Thus  f3(eγ1ξ)f3(eγ2ξ) j k2 ≤ Xeα1, eα2∈Wp−j,0 kSζ  Xξ∈Wj,0 Xξ∈Wj,0, eβ,eγ1,eγ2∈Wp−j,0 f3(eγ1ξ)f3(eγ2ξ) Xeβ,eγ1,eγ2∈Wp−j,0 Xeα1, eα2∈Wp−j,0 eγ1ξ,eγ2ξ,ξ−1 eβζ −1∈Wp,0 (eα1, eβ,eγ1),(eα2, eβ,eγ2)∈Tp−j (eα1, eβ,eγ1)∈Tp−j (eα2, eβ,eγ2)∈Tp−j   eα2∈Wp−j,p−j eα2∈Wp−j,p−j eγ1ξ,eγ2ξ∈Wp,0 (11) eα−1 1 = eα−1 1 , 2 where the last term is obtained from the previous one by using the smaller triangles formed by where the middle sum is over all possible diagrams of the form depicted in Figure 15. Since eα−1 1 eα2 ∈ Wp−j,p−j, an application of Lemma 1.8 shows that this rhombus is entirely determined by eα1 and eα2 and so that the middle sum contains only one term. The rhombus in Figure 15 is 2 ∈ Wp−j,p−j, so we in fact have Figure 15. Diagram indexing the middle sum in (11). ................ ............................................................................................................................................................... ................ ............................................................................................................................................................... ................ ............................................................................................................................................................................... eγ1 eγ2 eβ ............................................................................................................................................................................... ............................................................................................................................................................... eα1 eα2 also uniquely determined byeγ1 andeγ2 whenevereγ1eγ−1 f3(eγ1ξ)f3(eγ2ξ)  Xξ∈Wj,0 jk2 ≤ Xeγ1,eγ2∈Wp−j,0 kSζ  Xξ2∈Wj,0 f3(eγ1ξ1)2  Xξ1∈Wj,0 f3(γ2)2! f3(γ1)2! Xγ2∈G jk2 ≤ Xeγ1,eγ2∈Wp−j,0 kSζ ≤ Xγ1∈G Applying the Cauchy-Schwarz inequality we thus obtain 2 ∈Wp−j,p−j 2 ∈Wp−j,p−j eγ1ξ,eγ2ξ∈Wp,0 ξ1eγ1∈Wp,0 eγ2ξ2∈Wp,0 eγ1eγ−1 eγ1eγ−1 . 2 f3(eγ2ξ2)2 = kf3k4 2 18 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER as desired. This concludes our proof of Theorem 1.9 in the eA2 case. 4. Simply transitive actions (cid:3) Λ Suppose that Γ ≤ AutS(∆) acts simply transitively on V∆. Fix any vertex v0 ∈ V∆ and let N = {a ∈ Γ d(v0, av0) = 1}. The Cayley graph of Γ constructed via right multiplication with respect to the set N has Γ itself as its vertex set and has {(c, ca) c ∈ Γ, a ∈ N} as its edge set. There is a natural action of Γ on its Cayley graph via left multiplication. Using the convention that an undirected edge between vertices u and v in a graph represents the pair of directed edges (u, v) and (v, u), it is immediate that the Γ-map c 7→ cv0 from Γ to ∆ is an isomorphism between the Cayley graph of Γ and the one-skeleton of ∆. In this way we identify Γ with the vertex set V∆ of the building. Connectivity of the building implies that N is a generating set for Γ. It is traditional to label the directed edge (c, ca) with the generator a ∈ N. More generally, to the pair (c, d) ∈ Γ × Γ we assign the label c−1d. Equivalently, to the pair (c, cd) we assign the label d ∈ Γ. Suppose this label is written as a product of generators; d = a1 · · · aj. Then there is a path (c, ca1, ca1a2, . . . , cd) from c to cd whose successive edges are labelled a1, . . . , aj. The left translate of (c, cd) by b ∈ Γ is (bc, bcd) and also carries the label d. Conversely, any pair (c′, c′d) which carries the label d is the left translate by b = c′c−1 of (c, cd). Thus two pairs carry the same label if and only if one is the left translate of the other. We define a shape function on Γ by σ(b) = σ(v0, bv0) for b ∈ Γ. The pair (c, cb) has label b and its shape, defined via the identification of the Cayley graph and the one-skeleton of ∆, is σ(c, cb) = σ(cv0, cbv0) = σ(v0, bv0) = σ(b). A different choice of v0 leads to a shape function on Γ which differs from the first by precomposition with an inner automorphism of Γ. Suppose ∆ is an eAn1 × · · · × eAnk building and that Γ ≤ Aut(∆) consists of type-rotating × · · · × eAnk group. automorphisms and acts simply transitively on V∆. Then Γ is called an eAn1 4.1. The case of simply transitive group actions on the vertices of eA1 × eA1 buildings. Suppose that ∆ = ∆1 × ∆2 is an eA1 × eA1 building and a ∈ AutS(∆). If we write (u, v) for a generic vertex of ∆ = ∆1 × ∆2 we have a(u, v) = (a1u, a2v) Indeed, suppose that (u, v) and (u, v′) are for some type rotating automorphisms ai of ∆i. neighbouring vertices in ∆. Then a(u, v) = (x, y) and a(u, v′) = (x′, y′) are neighbouring ver- tices in ∆ and the type-rotating assumption on a means that τ (x) = τ (x′). Since neighbour- ing vertices in ∆1 have distinct types we must have x = x′. By induction on d(v, v′), we see that the first coordinate of a(u, v) is independent of v ∈ ∆2. Similarly, the second coordi- nate of a(u, v) is independent of u ∈ ∆1. Thus there exist maps a1 of ∆1 and a2 of ∆2 such that a(u, v) = (a1u, a2v). Since a ∈ AutS(∆) it follows that ai is a (type-rotating) automorphism of ∆i. Thus each a ∈ AutS(∆) acts as a1 × a2 for some (type-rotating) automorphisms ai of ∆i. A simple example of such a group is Fp× Fq where Fj denotes the free group on j generators. More Consider an eA1 × eA1 group, that is a group Γ ≤ AutS(∆) which acts simply transitively on V∆. generally, Γ = Γ1 × Γ2 is an eA1 × eA1 group whenever Γ1 and Γ2 are eA1 groups. However there are also many examples of eA1 × eA1 groups generated by eA1 groups Γ1 and Γ2 where the elements has given examples of eA1 groups Γ1, Γ2 acting on trees T1, T2 which are embedded "diagonally" of Γ1 do not all commute with those of Γ2. In fact S. Mozes [M, Theorem 3.2] (see also [BM]) A HAAGERUP INEQUALITY FOR eA1 × eA1 AND eA2 BUILDINGS 19 in Aut(T1 × T2), so that the embeddings do not commute but the group generated by Γ1 ∪ Γ2 is of automorphisms, Γ is not a direct product of the groups Γ1 and Γ2 . an eA1 × eA1 group acting on T1 × T2. In other words, even though each a ∈ Γ is a direct product As above, fix a vertex v0 = (v1, v2) ∈ V∆ and suppose that b = b1 × b2 ∈ Γ. Recall that σ(b) = σ(v0, bv0) = (d(v1, b1v1), d(v2, b2v2)) , so that where σi is the shape function on Aut(∆i). Consider the generating set σ(b1 × b2) = (σ1(b1), σ2(b2)) of Γ. Let N = {a ∈ Γ d(v0, av0) = 1} N1 = {a ∈ Γ σ(a) = (1, 0)} and N2 = {a ∈ Γ σ(a) = (0, 1)}. Then each element c ∈ Γ has a unique reduced expression of the form c = a1 · · · ambm+1 · · · bm+n for some ai ∈ N1 and bi ∈ N2 and the shape function satisfies σ(a1 · · · ambm+1 · · · bm+n) = (m, n). Corollary 4.1. If Γ is an eA1 × eA1 group, ρ is the right regular representation of ℓ1(Γ) on ℓ2(Γ) defined by ρ(g)f = f ∗ g, and g ∈ CΓ is supported on words of shape (m, n) then kρ(g)k ≤ (m + 1)(n + 1) kgk2. Consequently Γ has property (RD). Proof. These results follow from Theorem 1.1 and Corollary 1.10. (cid:3) our approach provides a unified method to establish property (RD) for these groups. This answers a question in [Ha2, Section 6]. Remark 4.2. Since Z2 and the free group F2 can both be realized as subgroups of eA1×eA1 groups, 4.2. The case of simply transitive group actions on the vertices of eA2 buildings. Suppose now that ∆ is an eA2 building and Γ ≤ AutS(∆) acts simply transitively on V∆. Recall that such a group Γ is called an eA2 group. A detailed account of eA2 groups may be found in [CMSZ], while [RR] and [RS] have quick introductions to eA2 buildings and groups. Some, but not all, eA2 groups For an eA2 group Γ, the shape function σ defined in §4 has an illuminating algebraic interpreta- can be embedded as lattices in matrix groups of the form P GL3(K) where K is a local field [CMSZ, II §8]. tion. Recall that for any fixed vertex v0 ∈ V∆ the set N = {a ∈ Γ σ(v0, av0) = (1, 0)} is a set of generators of Γ. In terms of these generators Γ has a presentation of the form where T is a so-called triangle presentation. Each element of Γ has a unique shortest expression of the form Γ = hai axayaz = 1 for (x, y, z) ∈ T i im+1 · · · a−1 where the aij ∈ N and the shape function σ on Γ is given by x = ai1 · · · aima−1 im+n σ(ai1 · · · aima−1 im+1 · · · a−1 im+n) = (m, n). 20 JACQUI RAMAGGE, GUYAN ROBERTSON, AND TIM STEGER Corollary 4.3. Let Γ be an eA2 group and ρ be the right regular representation of ℓ1(Γ) on ℓ2(Γ) defined by ρ(g)f = f ∗ g. If g ∈ CΓ is supported on words of shape (m, n) then kρ(g)k ≤ 1/2(m + 1)(n + 1)(m + n + 2)pmax(m, n) + 1 kgk2. Moreover, Γ has property (RD). Proof. These results follow from Theorem 1.1 and Corollary 1.10. (cid:3) [BM] M. Burger and S. Mozes, Finitely presented groups and products of trees, C.R. Acad. Sci. Paris, S´er. 1, 324 [Br] (1997), 747 -- 752. K. Brown, Buildings, Springer-Verlag, New York 1989. References [CMS] D. I. Cartwright, W. M lotkowski, and T. Steger, Property (T) and eA2 groups, Ann. Inst. Fourier, 44 (1994) [CMSZ] D. I. Cartwright, A. M. Mantero, T. Steger, and A. Zappa, Groups acting simply transitively on the vertices pp. 213 -- 248. of a building of type eA2, I & II, Geom. Ded. 47 (1993) pp. 143 -- 166 and 167 -- 226. [C] [CoM] A. Connes and H. Moscovici, Cyclic cohomology, the Novikov conjecture and hyperbolic groups, Topology, A. Connes, Noncommutative Geometry, Academic Press, New York, 1994. 29 (1990), pp. 345 -- 388. S.C. Ferry, A. Ranicki and J. Rosenberg (eds), Novikov conjectures, index theorems and rigidity, volume I, L.M.S. Lecture Note Series, 226, C.U.P. 1995. U. Haagerup, An example of a nonnuclear C ∗-algebra which has the metric approximation property, Inv. Math., 50 (1979), pp. 279 -- 293. P. de la Harpe, Groupes hyperboliques, alg`ebres d'op´erateurs et un th´eoreme de Jolissaint, C. R. Acad. Sc. Paris, 307, S´erie I (1988), pp. 771 -- 774. P. de la Harpe, Operator algebras, free groups and other groups, Recent Advances in Operator Algebras, Ast´erisque No. 232, Soci´et´e Math´ematique de France, 1995. P. Jolissaint, Rapidly decreasing functions in reduced C ∗-algebras of groups, Trans. Amer. Math. Soc., 317 (1990), pp. 167 -- 196. S. Mozes, Actions of Cartan Subgroups, Israel J. Math., 90 (1995), pp. 253 -- 294. A. Nevo, Spectral transfer and pointwise ergodic theorems for Kazhdan groups, preprint, Technion, Haifa, 1997. P.S. Muhly and D. Williams, Groupoid cohomology and the Dixmier-Douady class, Proc. London Math. Soc., 71 (1995), pp. 109 -- 134. M. Ronan, Lectures on Buildings, Academic Press, San Diego, 1989. J. Renault, C ∗-algebras of groupoids and foliations, Operator algebras and applications, part 1, Proc. Symp. Pure. Math. AMS, 38 (1982), pp. 339 -- 350. J. Ramagge and G. Robertson, Triangle buildings and actions of type III1/q2 , J. Func. Anal., 140 (1996), 472 -- 504. G. Robertson and T. Steger, C ∗-algebras arising from group actions on the boundary of a triangle building, Proc. London Math. Soc., 72 (1996), pp. 613 -- 637. A Selberg, On discontinuous groups in higher dimensional symmetric spaces, Contributions to Function Theory, Tata Instituta, Bombay, 1960. J. ´Swia tkowski, On the loop inequality for Euclidean buildings, preprint, Wroc law, 1995. ' [FRR] [H] [Ha1] [Ha2] [J] [M] [N] [MW] [R] [Ren] [RR] [RS] [Sel] [Swa] [V] A. Valette, On the Haagerup inequality and groups acting on eAn buildings, preprint, Neuchatel, 1996. Mathematics Department, University of Newcastle, Callaghan, NSW 2308, Australia E-mail address: [email protected] Mathematics Department, University of Newcastle, Callaghan, NSW 2308, Australia E-mail address: [email protected] Istituto Di Matematica e Fisica, Universit`a degli Studi di Sassari, Via Vienna 2, 07100 Sassari, Italia E-mail address: [email protected]
1601.05761
3
1601
2016-08-14T01:54:09
Super-resolution by means of Beurling minimal extrapolation
[ "math.FA" ]
Let $M(\mathbb{T}^d)$ be the space of complex bounded Radon measures defined on the $d$-dimensional torus group $(\mathbb{R}/\mathbb{Z})^d=\mathbb{T}^d$, equipped with the total variation norm $\|\cdot\|$; and let $\hat\mu$ denote the Fourier transform of $\mu\in M(\mathbb{T}^d)$. We address the super-resolution problem: For given spectral (Fourier transform) data defined on a finite set $\Lambda\subset\mathbb{Z}^d$, determine if there is a unique $\mu\in M(\mathbb{T}^d)$ of minimal norm for which $\hat\mu$ equals this data on $\Lambda$. Without additional assumptions on $\mu$ and $\Lambda$, our main theorem shows that the solutions to the super-resolution problem, which we call minimal extrapolations, depend crucially on the set $\Gamma\subset\Lambda$, defined in terms of $\mu$ and $\Lambda$. For example, when $\Gamma=0$, the minimal extrapolations are singular measures supported in the zero set of an analytic function, and when $\Gamma\geq 2$, the minimal extrapolations are singular measures supported in the intersection of $\Gamma\choose 2$ hyperplanes. By theory and example, we show that the case $\Gamma=1$ is different from other cases and is deeply connected with the existence of positive minimal extrapolations. This theorem has implications to the possibility and impossibility of uniquely recovering $\mu$ from $\Lambda$. We illustrate how to apply our theory to both directions, by computing pertinent analytical examples. These examples are of interest in both super-resolution and deterministic compressed sensing. Our concept of an admissibility range fundamentally connects Beurling's theory of minimal extrapolation with Candes and Fernandez-Granda's work on super-resolution. This connection is exploited to address situations where current algorithms fail to compute a numerical solution to the super-resolution problem.
math.FA
math
SUPER-RESOLUTION BY MEANS OF BEURLING MINIMAL EXTRAPOLATION JOHN J. BENEDETTO AND WEILIN LI Abstract. Let M (Td) be the space of complex bounded Radon measures defined on the d-dimensional torus group (R/Z)d = Td, equipped with the total variation norm k · k; and let bµ denote the Fourier transform of µ ∈ M (Td). We address the super-resolution prob- lem: For given spectral (Fourier transform) data defined on a finite set Λ ⊆ Zd, determine if there is a unique µ ∈ M (Td) of minimal norm for which bµ equals this data on Λ. With- out additional assumptions on µ and Λ, our main theorem shows that the solutions to the super-resolution problem, which we call minimal extrapolations, depend crucially on the set Γ ⊆ Λ, defined in terms of µ and Λ. For example, when #Γ = 0, the minimal extrapo- lations are singular measures supported in the zero set of an analytic function, and when #Γ ≥ 2, the minimal extrapolations are singular measures supported in the intersection of (cid:0)#Γ 2 (cid:1) hyperplanes. By theory and example, we show that the case #Γ = 1 is different from other cases and is deeply connected with the existence of positive minimal extrapolations. This theorem has implications to the possibility and impossibility of uniquely recovering µ from Λ. We illustrate how to apply our theory to both directions, by computing per- tinent analytical examples. These examples are of interest in both super-resolution and deterministic compressed sensing. Our concept of an admissibility range fundamentally connects Beurling's theory of minimal extrapolation [Beu89a, Beu89b] with Cand`es and Fernandez-Granda's work on super-resolution [CFG14]. This connection is exploited to address situations where current algorithms fail to compute a numerical solution to the super-resolution problem. 1. Introduction 1.1. Motivation. The term super-resolution varies depending on the field, and, conse- quently, there are various types of super-resolution problems. In some situations [PPK03], super-resolution refers to the process of up-sampling an image onto a finer grid, which is a spatial interpolation procedure. In other situations [Lin12], super-resolution refers to the process of recovering the object's high frequency information from its low frequency infor- mation, which is a spectral extrapolation procedure. In both situations, the super-resolution problem is ill-posed because the missing information can be arbitrary. However, it is possi- ble to provide meaningful super-resolution algorithms by using prior knowledge of the data. We develop a mathematical theory of the spectral extrapolation version, which we simply refer to as the super-resolution problem, see Problem (SR) for a precise statement. To be concrete, our exposition focuses on imaging applications, but super-resolution ideas are of interest in other fields, e.g., [Rie99, KLM04]. In such applications, an image is obtained by convolving the object with the point spread function of an optical lens. Alternatively, the Fourier transform of the object is multiplied by a modulation transfer function. The resulting image's resolution is inherently limited by the Abbe diffraction 2010 Mathematics Subject Classification. 43A25, 46E27, 46N10, 94A08, 94A12. Key words and phrases. Super-resolution, minimal extrapolation, total variation optimization. 1 2 JOHN J. BENEDETTO AND WEILIN LI limit, which depends on the illumination light's wavelength and on the diameter of the optical lens. Thus, the optical lens acts as a low-pass filter, see [Lin12]. The purpose of super-resolution is to use prior knowledge about the object to obtain an accurate image whose resolution is higher than what can be measured by the optical lens. We mention two specific imaging problems, and they motivate the theory that we develop. (a) In astronomy [PK05], each star can be modeled as a complex number times a Dirac δ-measure, and the Fourier transform of each star encodes important information about that star. However, an image of two stars that are close in distance resembles an image of a single star. In this context, the super-resolution problem is to determine the number of stars and their locations, using the prior information that the actual object is a linear combination of Dirac δ-measures. (b) In medical imaging [Gre09], machines capture the structure of the patient's body tissues, in order to detect for anomalies in the patient. Their shapes and locations are the most important features, so each tissue can be modeled as the characteristic function of a closed set, or as a surface measure supported on the boundary of a set. The super- resolution problem is to capture the fine structures of the tissues, given that the actual object is a linear combination of singular measures. 1.2. Problem statement. We model objects as elements of M (Td), the space of complex bounded Radon measures on Td = (R/Z)d, the d-dimensional torus group. M (Td) equipped with the total variation norm k · k is a Banach algebra with unit, the Dirac δ-measure, where multiplication is defined by convolution ∗. The Fourier transform of µ ∈ M (Td) is the function bµ : Zd → C, whose m-th Fourier coefficient is defined as e−2πim·x dµ(x). bµ(m) =ZTd See [BC09] for further details. (SR) arg min We consider the super-resolution problem: For a given finite subset Λ ⊆ Zd and given spectral data F of the form bµ Λ for some unknown µ ∈ M (Td), find ν∈M (Td)kνk subject to bν = bµ on Λ. This is a is a convex minimization problem and we interpret a solution as a simple or least complicated high resolution extrapolation of F. Remark 1.1 (Image processing). (a) A primary objective is to determine if µ ∈ M (Td) is a solution to the super-resolution problem given its spectral data bµ Λ. The current literature has focused on discrete µ ∈ M (Td), e.g., see [CFG14, TBSR13, DCG12]. However, in the context of image processing, µ is the unknown high resolution image, and a typical image cannot be approximated by a discrete measure. Hence, it is impor- tant to determine if non-discrete measures are solutions to the super-resolution problem. For this reason, we have formulated and shall study Problem (SR) in an abstract way. (b) Further, in the context of image processing, we can think of being given information that represents an image µ ∈ M (Td), and this information is in the form, F (x) = (µ ∗ ψ)(x) =ZTd ψ(x − y) dµ(y), BEURLING SUPER-RESOLUTION 3 for some ψ : Td → C. For simplicity, we assume that bψ = 1Λ, the characteristic function of some finite set Λ ⊆ Zd. Then, we have Thus, we interpret bµ Λ as the given low frequency information of µ, and (bµ Λ)∨ as the given low resolution image, even though in applications, we do not know the desired object µ. This is the background for our formulation of Problem (SR). F = (µ ∗ ψ) = (bµ Λ)∨. With regard to Problem (SR), a fundamental question of uniqueness must be addressed. To see why this is so, if µ is not the unique solution to the super-resolution problem, then the output of a numerical algorithm is not guaranteed to approximate µ. Thus, it is important to determine sufficient conditions such that µ is the unique solution. For this reason, we say that super-resolution reconstruction of µ from Λ is possible, or it is possible to super-resolve µ from Λ, if and only if µ is the unique solution to Problem (SR). Of course, it could be theoretically be possible to reconstruct µ by other means. 1.3. Background. Problem (SR) was first studied by Cand`es and Fernandez-Granda [CFG14] and was inspired by results from compressed sensing [Don06, CRT06]. By considering dif- ferent models of images and/or alternative measurement processes, it is possible to derive related (noiseless) "super-resolution" problems, e.g., see [Don92, DCG12, TBSR13, ASB14, ASB15]. There are several papers that address "super-resolution" in the presence of noise, e.g., see [CFG13, BTR13, ADCG15, TBR15, DP15]. We briefly discuss the important results of Cand´es and Fernandez-Granda [CFG14]. k=1 akδxk ∈ M (Td) satisfies Given Λ = {−M,−M + 1, . . . , M}d ⊆ Zd, we say µ = PK the minimum separation condition with constant C > 0 if (1.1) inf 1≤j,k≤K j6=k kxj − xkkℓ∞(Td) ≥ C M . Theorem 1.2 (Cand`es and Fernandez-Granda, Theorems 1.2-1.3, [CFG14]. Fernandez -- k=1 akδxk ∈ M (Td) and Λ = {−M,−M + Granda, Theorem 2.2, [FG16]). Let µ = PK 1, . . . , M}d ⊆ Zd. (a) If d = 1, M ≥ 128, and µ satisfies the minimum separation condition with C = 2, then µ is the unique solution to Problem (SR). If additionally, µ is real valued, then the constant can be reduced to C = 1.87. (b) If d = 1, M ≥ 103, and µ satisfies the minimum separation condition with C = 1.26, (c) If d = 2, µ is real valued, M ≥ 512, and µ satisfies the minimum separation condition then µ is the unique solution to Problem (SR). for C = 2.38, then µ is the unique solution to Problem (SR). The theorem proves that, in theory, it is possible to uniquely recover discrete measures satisfying a sufficiently strong minimum separation condition. To compute a numerical approximation of such a measure, they proposed a convex-optimization based algorithm which relies on [Dum07, Theorem 4.24]. They stated their algorithm for d = 1, but it is possible an analogous algorithm for higher dimensions. Algorithm 1.3 (Cand´es and Fernandez-Granda, Section 4, [CFG14]). Suppose µ ∈ M (T) and Λ = {−M,−M + 1, . . . , M} ⊆ Z satisfy the hypotheses in any of the three situations described in Theorem 1.2. 4 JOHN J. BENEDETTO AND WEILIN LI (1) Solve the convex optimization problem described in [CFG14, Corollary 4.1] to obtain m=−M cme2πimx such that kϕk∞ ≤ 1 and hϕ, µi = kµk. Then, a function ϕ(x) = PM µ is supported in the set, S = {x ∈ T : ϕ(x) = 1}. (2) Assume that S 6= T. Then, S is a discrete set that consists of at most 2M points, and it is possible to determine µ by solving a system of over-determined equations. Cand`es and Fernandez-Granda pointed out that their algorithm for recovering µ is in- complete because it fails precisely when {x ∈ T : ϕ(x) = 1} = T, i.e., when ϕ ≡ 1, since in this case, ϕ does not provide information about the support of µ. We emphasize that this scenario can occur even if µ ∈ M (T) and Λ ⊆ Z satisfy the hypotheses of Theorem 1.2. 1.4. Our approach. This paper connects the Cand´es and Fernandez-Granda theory on super-resolution [CFG14] with Beurling's work on minimal extrapolation [Beu89a, Beu89b]. This connection is exploited to obtain new theoretical and computational results on super- resolution. We first introduce some additional notation, since referring to Problem (SR) can be ambiguous when working with several different measures. We adopt Beurling's language, and we say that ν ∈ M (Td) is a minimal extrapolation of µ ∈ M (Td) from Λ ⊆ Zd if it is a solution to Problem (SR). If ν ∈ M (Td) and bµ = bν on Λ, then we say ν is an extrapolation of µ from Λ. When there is no ambiguity, we shall say that "ν is an extrapolation (respectively, a minimal extrapolation)" as a shortened version of "ν is an extrapolation (respectively, a minimal extrapolation) of µ from Λ". Let E = E(µ, Λ) be the set of all minimal extrapolations. For any ν ∈ E, let ε = ε(µ, Λ) = kνk, which is the optimal value attained by Problem (SR). It is possible to compute a numerical approximation of ε by solving the convex optimization problem described in [CFG14, Coroallary 4.1], but its exact value is unknown. Finally, we define the set, (1.2) Γ = Γ(µ, Λ) = {m ∈ Λ : bµ(m) = ε(µ, Λ)}. Although Γ is defined in terms of the unknown quantity ε, we shall see that, for many important applications, it is possible to deduce ε and Γ. Our theorem shows that the minimal extrapolations intimately depend on Γ. Theorem 1.4. Let µ ∈ M (Td) and let Λ ⊆ Zd be a finite subset. (a) Suppose #Γ = 0. Then, there exists a closed set S of d-dimensional Lebesgue measure zero such that each minimal extrapolation is a singular measure supported in S. In particular, if d = 1, then S is a finite number of points and each minimal extrapolation is a discrete measure supported in S. (b) Suppose #Γ ≥ 2. For each distinct pair m, n ∈ Γ, define αm,n ∈ R/Z by e2πiαm,n = (1.3) bµ(m)/bµ(n) and the closed set, S = \m,n∈Γ which is an intersection of(cid:0)#Γ m6=n {x ∈ Td : x · (m − n) + αm,n ∈ Z}, 2 (cid:1) periodic hyperplanes. Then, each minimal extrapolation is a singular measure supported in S. In particular, if d = 1, then S is a finite number of points and each minimal extrapolation is a discrete measure supported in S. BEURLING SUPER-RESOLUTION 5 If d ≥ 2 and there exist d linearly independent vectors, p1, p2, . . . , pd ∈ Zd, such that then S is a lattice on Td. {p1, p2, . . . , pd} ⊆ {m − n : m, n ∈ Γ}, The following topics encapsulate the contributions of this paper. (a) Qualitative behavior of minimal extrapolations. It is interesting to note that Γ provides significant information about the minimal extrapolations. Theorem 1.4 shows that, regardless of the dimension, when #Γ = 0 or #Γ ≥ 2, the minimal extrapolations are always singular measures. However, when #Γ = 1, the minimal extrapolations can be extremely different from each other. Proposition 2.8 shows that the case #Γ = 1 is connected with the existence of positive absolutely continuous minimal extrapolations. In Example 3.3, we provide an example for which there are uncountably many discrete minimal extrapolations as well as uncountably many positive absolutely continuous minimal extrapolations. (b) Computational consequences. One important aspect of Theorem 1.4 is its relationship with Algorithm 1.3. Proposition 2.6 shows that, if the algorithm fails, then we nec- essarily have ε = kbµkℓ∞(Λ) and Γ 6= ∅. Since we are given the values of bµ on Λ, this immediately tells us what ε and Γ are. If #Γ = 1, then we cannot deduce anything about the minimal extrapolations. As previously discussed, this is inherent with the super-problem and is not an artifact of our analysis. If #Γ ≥ 2, then the theorem is applicable and the minimal extrapolations are singular measures supported in the set defined in (1.3), which can be explicitly computed. We show how to apply our theorem to compute pertinent analytical examples in Section 3. Hence, our theorem is applicable even when Algorithm 1.3 fails. (c) Super-resolution of singular measures. When d ≥ 2, Theorem 1.4 suggests that some singular continuous measures could be solutions to the super-resolution problem. Some- what surprisingly, this is indeed the case because Example 3.7 provides an example of a singular continuous minimal extrapolation. This also demonstrates that the conclusion of Theorem 1.4b is optimal. Determining what types of singular continuous measures can be uniquely recovered by Problem (SR) is an interesting mathematical problem, and we believe such a result would be useful in applications, e.g., [Gre09]. We shall carefully address the super-resolution of singular continuous measures in the sequel [BL16]. (d) Impossibility of super-resolution. Theorem 1.4 does not require additional assumptions on µ ∈ M (Td) or on the finite subset Λ ⊆ Zd. Since the theorem also describes the support set of the minimal extrapolations of µ from Λ, it is useful for determining whether it is possible for a given µ to be a minimal extrapolation. To illustrate this point, in Example 3.6, we provide a simple proof that shows, in general, a minimal separation condition is necessary to super-resolve a discrete measure. (e) Discrete-continuous correspondence. The super-resolution problem can be viewed as a continuous analogue of the discrete Fourier compressed sensing problem that was studied in [CRT06]. Let FN ∈ CN ×N be the DFT matrix, x ∈ CN be an unknown vector, and Λ ⊆ {0, 1, . . . , N − 1}. The goal is to recover x from FN x Λ by solving (CS) y∈CN kykℓ1 arg min subject to FN x = FN y on Λ. By considering the case that µ = PN n=1 anδn/N and x(n) = an, it is straightforward to see that Problem (SR) is a generalization of Problem (CS), see [CFG14] for further 6 JOHN J. BENEDETTO AND WEILIN LI details. From this point of view, Theorem 1.4 is a discrete-continuous correspondence result. Indeed, if #Γ ≥ 2, and either d = 1 or the vectors {pj} exist, then the minimal extrapolations are necessarily discrete measures whose support lie on a lattice; as we just discussed, such measures can be identified with vectors that are solutions to the discrete problem. (f) Mathematical connections. Our results are closely related to Beurling's work on minimal extrapolation. He dealt with R1 instead of Td, so Theorem 1.4 is an adaptation to the torus and a generalization to higher dimensions of [Beu89b, Theorem 2, page 362]. We remark that there are non-trivial technical differences between working with R1 and Td. There are also other classical papers related to minimal extrapolation, e.g., [Ess45, Ess54, HR55]. Propositions 2.4 connects our results to the Cand`es and Fernandez-Granda theory [CFG14] by introducing a concept called an admissibility range for ε. This connection is exploited to prove Proposition 2.6, which in turn, shows that our theorem is applicable to situations where Algorithm 1.3 fails. Proposition 2.1 contains basic duality results for Problem (SR) that can be derived from abstract convex analysis, e.g., see [DP15]. None of our proofs require knowledge of convex analysis, so we believe our paper is accessible to non-specialists. 1.5. Outline. Section 2 contains the basis of our mathematical theory. Section 2.1 proves well-known results on the super-resolution problem using only basic functional analysis. Section 2.2 provides a characterization of Γ and illustrates its importance in our analysis. The proof of Theorem 1.4 is given in Section 2.3. Section 2.4 contains the results that fully connect the Beurling and Cand`es and Fernandez-Granda theories, as well as the the rela- tionship between Theorem 1.4 and Algorithm 1.3. Section 2.5 includes a basic uniqueness result, and the mentioned non-uniqueness results for the case #Γ = 1. Section 2.6 exam- ines whether it is possible to relate the minimal extrapolations of µ with those of T µ for various operators T . Section 3 contains all of the examples that we have previously men- tioned, including the minimal extrapolations of discrete measures and of singular continuous measures. 2. Mathematical Theory 2.1. Preliminary results. Let C(Td) be the space of complex-valued continuous functions on Td equipped with the sup-norm k · k∞. Then, C(Td) is a Banach space, and let C(Td)′ be its dual space of continuous linear functionals with the usual norm k · k. The celebrated Riesz representation theorem, e.g., see [BC09, Theorem 7.2.7, page 334] states that: (a) For each µ ∈ M (Td), there exists a bounded linear functional ℓµ ∈ C(Td)′ such that kµk = kℓµk and ∀f ∈ C(Td), ℓµ(f ) = hf, µi =ZTd f (x) dµ(x). (b) For each bounded linear functional ℓ ∈ C(Td)′, there exists a unique µ ∈ M (Td) such that kµk = kℓk and ∀f ∈ C(Td), ℓ(f ) = hf, µi =ZTd f (x) dµ(x). BEURLING SUPER-RESOLUTION 7 Proposition 2.1a shows that the super-resolution problem is well-posed, by using stan- dard functional analysis arguments. However, this type of argument does not yield useful statements about the minimal extrapolations. Instead of working with C(Td), we shall work with a subspace. Since only bµ Λ is important to the super-resolution problem, we consider the subspace C(Td; Λ) =nf ∈ C(Td) : f (x) = Xm∈Λ ame2πim·x, am ∈ Co. Proposition 2.1b shows that C(Td; Λ) is a closed subspace of C(Td), and that implies C(Td; Λ) is a Banach space. Further, Proposition 2.1c demonstrates that the closed unit ball of C(Td; Λ), is compact. U = U (Td; Λ) = {f ∈ C(Td; Λ) : kfk∞ ≤ 1}, The purpose of restricting to the subspace, C(Td; Λ), is to identify µ ∈ M (Td) with the bounded linear functional, Lµ ∈ C(Td; Λ)′, defined as ∀f ∈ C(Td; Λ), Lµ(f ) =ZTd f (x) dµ(x). Although, by definition, kLµk = supf ∈U Lµ(f ), Proposition 2.1d,e show that we have the stronger statement, kLµk = max f ∈U Lµ(f ) = ε. The purpose of studying Lµ is to deduce information about the minimal extrapolations. As a consequence of the Radon-Nikodym theorem, for each µ ∈ M (Td), there exists a µ-measurable function sign(µ) such that sign(µ) = 1 µ-a.e. and satisfies the identity ∀f ∈ L1 µ(Td), ZTd f dµ =ZTd f sign(µ) dµ. See [BC09, Theorem 5.3.2, page 242, and Theorem 5.3.5, page 244] for further details. The support of µ ∈ M (Td), denoted supp(µ), is the complement of all open sets A ⊆ Td such that µ(A) = 0. Proposition 2.1g shows that it is possible to interpolate the sign of the minimal extrapolations by a function belonging to U . Proposition 2.1. Let µ ∈ M (Td) and let Λ ⊆ Zd be a finite subset. (a) E ⊆ M (Td) is non-empty, weak-∗ compact, and convex. (b) C(Td; Λ) is a closed subspace of C(Td). (c) U is a compact subset of C(Td; Λ). (d) ε = kLµk. (e) ε = maxf ∈U hf, µi. (f ) There exists ϕ ∈ U such that hϕ, µi = ε. (g) If ϕ ∈ U and hϕ, µi = ε, then ∀ν ∈ E, ϕ = sign(ν) supp(ν) ⊆ {x ∈ Td : ϕ(x) = 1}. ν-a.e., and Proof. (a) By definition of ε, there exists a sequence {νj} ⊆ {ν : bν = bµ on Λ} such that kνjk → ε. Then, this sequence is bounded. By Banach-Alaoglu, after passing to a subsequence, we can assume there exists ν ∈ M (Td) such that νj → ν in the weak-∗ topology. 8 JOHN J. BENEDETTO AND WEILIN LI Let V be the closed unit ball of C(Td). We have kνk ≤ ε because kνk = sup f ∈V hf, νi = sup f ∈V j→∞hf, νji ≤ sup lim f ∈V j→∞kfk∞kνjk ≤ ε. lim e−2πim·x dνj(x) =ZTd Moreover, for m ∈ Λ, we have j→∞ZTd j→∞ bνj(m) = lim e−2πim·x dν(x) = bν(m). bµ(m) = lim This shows that ν is an extrapolation, and thus, kνk ≥ ε. Therefore, ν ∈ E. The proof that E is weak-∗ compact is similar. Pick any sequence {νj} ⊆ E and after passing to a subsequence, we can assume νj → ν in the weak-∗ topology for some ν ∈ M (Td). By the same argument, we see that ν ∈ E. If E contains exactly one measure, then E is trivially convex. Otherwise, let t ∈ [0, 1], ν0, ν1 ∈ E, and νt = (1− t)ν0 + tν1. Then, νt is an extrapolation and thus, kνtk ≥ ε. By the triangle inequality, we have kνtk ≤ (1 − t)kν0k + tkν1k = ε. Thus, νt ∈ E for each t ∈ [0, 1]. (b) Suppose {fj} ⊆ C(Td) and that there exists f ∈ C(Td) such that fj → f uniformly. Then, bfj(m) → bf (m) for all m ∈ Zd. Since bfj(m) = 0 if m 6∈ Λ, we deduce that bf (m) = 0 if m 6∈ Λ. This shows that f ∈ C(Td; Λ) and thus, C(Td; Λ) is a closed (c) Let {fj} ⊆ U . We first show that {fj} is a uniformly bounded equicontinuous family. By definition, kfjk∞ ≤ 1, and there exist aj,m ∈ C such that fj(x) =Pm∈Λ aj,me2πim·x. Note that aj,m ≤ kfjk∞ ≤ 1. Then, for any x, y ∈ Td, we have subspace of C(Td). fj(x) − fj(y) =(cid:12)(cid:12)(cid:12) Xm∈Λ aj,m(e2πim·x − e2πim·y)(cid:12)(cid:12)(cid:12) ≤ Xm∈Λ e2πim·x − e2πim·y. Let ε > 0 and m ∈ Λ. There exists δm > 0 such that e2πim·x − e2πim·y < ε whenever x − y < δm. Let δ = minm∈Λ δm. Combining this with the previous inequalities, we have fj(x) − fj(y) ≤ Xm∈Λ e2πim·x − e2πim·y ≤ #Λε, whenever x − y < δ. This shows that {fj} is a uniformly bounded equicontinuous family. By the Arzel`a-Ascoli theorem, there exists f ∈ C(Td), with kfk∞ ≤ 1, and a subse- quence {fjk} such that fjk → f uniformly. Thus, cfjk(m) → bf (m) for all m ∈ Zd, which shows that f ∈ C(Td; Λ). (d) Let ν ∈ E. Then, ∀f ∈ U, Lµ(f ) = hf, µi = hf, νi ≤ kfk∞kνk ≤ ε, which proves the upper bound, kLµk ≤ ε. For the lower bound, we use the Hahn-Banach theorem to extend Lµ ∈ C(Td; Λ)′ to ℓ ∈ C(Td)′, where kLµk = kℓk. By the Riesz representation theorem, there exists a unique σ ∈ M (Td) such ℓ(f ) = hf, σi for all f ∈ C(Td) and kσk = kℓk. In particular, ∀f ∈ C(Td; Λ), hf, σi = ℓ(f ) = Lµ(f ) = hf, µi. BEURLING SUPER-RESOLUTION 9 Set f (x) = e−2πim·x, where m ∈ Λ, to deduce that bµ = bσ on Λ. This implies kσk ≥ ε. Combining these facts, we have This proves the lower bound. ε ≤ kσk = kℓk = kLµk. (e) We know that kLµk = ε. By definition, there exists {fj} ⊆ U such that Lµ(fj) ≥ ε − 1/j. By compactness of U , there exists a subsequence {fjk} ⊆ U and f ∈ U such that fjk → f uniformly. We immediately have Lµ(f ) ≤ kfk∞ε ≤ ε. For the reverse inequality, as k → ∞, Lµ(f ) ≥ Lµ(fjk) − Lµ(f − fjk ) ≥ ε − 1 jk − kµkkf − fjkk∞ → ε. This proves that Lµ(f ) = ε. (f) There exists f ∈ U such that Lµ(f ) = ε. This implies eiθLµ(f ) = ε for some θ ∈ R. Hence, let ϕ = eiθf ∈ U , and thus, Lµ(ϕ) = ε. (g) Let ν ∈ E. Since ϕ ∈ U ⊆ C(Td; Λ), we have hϕ, νi = hϕ, µi = Lµ(ϕ) = ε = kνk. Since kϕk∞ ≤ 1 and hϕ, νi = kνk, we must have ϕ = sign(ν) ν-a.e. Since ϕ = sign(ν) = 1 ν-a.e. and ν is a Radon measure, supp(ν) ⊆ {x ∈ Td : ϕ(x) = 1} = {x ∈ Td : ϕ(x) = 1}. The last equality holds because the inverse image of the closed set {1} under the con- tinuous function ϕ is closed. (cid:3) 2.2. A characterization of Γ. Proposition 2.1f,g show that there exists ϕ ∈ U such that the minimal extrapolations are supported in the closed set S = {x ∈ Td : ϕ(x) = 1}. By itself, this statement is not useful because if ϕ ≡ 1, then S = Td. However, Proposition 2.2 characterizes the case that ϕ ≡ 1 in terms of the set Γ. Intuitively, #Γ represents the number of "bad" functions in U that interpolate the signs of the minimal extrapolations. While it is desirable to have Γ = ∅, perhaps surprisingly, we can make strong statements when #Γ is large. Proposition 2.2. Let µ ∈ M (Td) and Λ ⊆ Zd be a finite subset. (a) Suppose ϕ ∈ U , hϕ, µi = ε, and ϕ ≡ 1. Then, Γ 6= ∅. (b) For each m ∈ Zd, define αm ∈ R/Z by the formula e−2πiαmbµ(m) = bµ(m). If m ∈ Γ, then sign(ν)(x) = e2πiαm e2πim·x ν-a.e. ∀ν ∈ E, Proof. (a) Since ϕ ∈ U and ϕ ≡ 1, we must have ϕ = e2πiθe2πim·x for some m ∈ Λ and θ ∈ R. Then, we have which shows that m ∈ Γ. ε = hϕ, µi = e−2πiθ bµ(m), 10 JOHN J. BENEDETTO AND WEILIN LI (b) Suppose m ∈ Γ and ν ∈ E. Then ZTd e−2πiαme−2πim·x dν(x) = e−2πiαmbν(m) = e−2πiαmbµ(m) = bµ(m) = ε = kνk. This shows sign(ν)(x) = e2πiαme2πim·x ν-a.e. 2.3. An analogue of Beurling's theorem. We are ready to prove Theorem 1.4. Proof. (a) By Proposition 2.1, there exists ϕ ∈ U such that hϕ, µi = ε, and each minimal extrap- olation is supported in the closed set (cid:3) S = {x ∈ Td : ϕ(x) = 1}. Note that ϕ 6≡ 1. In fact, if ϕ ≡ 1, then Proposition 2.2a implies Γ 6= ∅, which is a contradiction. We have ϕ(x) = Pm∈Λ ame2πim·x for some am ∈ C. Consider the function (2.1) Φ(x) = 1 − ϕ(x)2 = 1 − Xm∈ΛXn∈Λ amane2πi(m−n)·x. Then, the minimal extrapolations are supported in the closed set S = {x ∈ Td : Φ(x) = 0}. Note that Φ 6≡ 0 because ϕ 6≡ 1. Since Φ is a non-trivial real-analytic function, S is a set of d-dimensional Lebesgue measure zero. In particular, if d = 1, then S is a finite set of points. (b) Let m, n ∈ Γ. There exist αm, αn ∈ R/Z defined in Proposition 2.2b, such that sign(ν)(x) = e2πiαm e2πim·x = e2πiαn e2πin·x ν-a.e. ∀ν ∈ E, Set αm,n = αm − αn ∈ R/Z. Then each minimal extrapolation is supported in Thus, each minimal extrapolation is supported in the set Sm,n = {x ∈ Td : x · (m − n) + αm,n ∈ Z}. Sm,n = \m,n∈Γ {x ∈ Td : x · (m − n) + αm,n ∈ Z}. S = \m,n∈Γ m6=n m6=n Note that e2πiαm,n = bµ(m)/bµ(n) because Suppose {p1, . . . , pd} satisfies the hypothesis. By the support assertion that we just proved, there exists β = (β1, β2, . . . , βd) ∈ Td such that every minimal extrapolation is supported in e−2πiαmbµ(m) = bµ(m) = ε = bµ(n) = e−2πiαnbµ(n). S = d\j=1 {x ∈ Td : x · pj + βj ∈ Z}. Let us explain the geometry of the situation before we proceed with the proof that S is a lattice. Note that {x ∈ Td : x·pj+βj ∈ Z} is a family of parallel and periodically spaced hyperplanes. Since the vectors, p1, p2, . . . , pd, are assumed to be linearly independent, one family of hyperplanes is not parallel to any other family of hyperplanes. Hence, BEURLING SUPER-RESOLUTION 11 the intersection of d non-parallel and periodically spaced hyperplanes is a lattice, see Figure 2.1 for an illustration. For the rigorous proof of this observation, first note that S = {x ∈ Td : P x + β ∈ Zd}, where P = (p1, p2, . . . , pd)t ∈ Zd×d is invertible because its rows are linearly indepen- dent. Let x0 ∈ Rd be the solution to P x + β = 0, and let qj ∈ Qd be the solution to P x = ej, where ej is the standard basis vector for Rd. Then pj · qk = δj,k, and S is generated by the point x0 and the lattice vectors, q1, q2, . . . , qd. p2 (cid:3) p1 q1 x0 q2 Figure 2.1. An illustration of Theorem 1.4b, where d = 2, p1 = (1, 2), p2 = (−3, 2), β1 = 1/2, β2 = −1/2, q1 = (1/4, 3/8), and q2 = (−1/4, 1/8). The family of hyperplanes are the dashed lines, the lattice S is the black dots, and the shaded region is [0, 1)2. Remark 2.3. Example 3.7 provides an example where #Γ = 3 and there exists a singular continuous minimal extrapolation. This demonstrates that the conclusion of Theorem 1.4b is optimal. 2.4. The admissibility range for ε. While the mathematical theory we have developed connects the set Γ with the support of the minimal extrapolations, the difficulty of applying the theory is that in general, ε is unknown. However, in many important situations, it is possible to deduce the value of ε. We say [A, B] ⊆ R+ is an admissiblity range for ε provided that 0 ≤ A ≤ ε ≤ B. The following proposition shows that we have A = kbµkℓ∞(Λ) and B = kµk as an admissibility range, and B can be improved in certain situations. Proposition 2.4. Let µ ∈ M (Td) and let Λ ⊆ Zd be a finite subset. We have the lower and upper bounds, (2.2) kbµkℓ∞(Λ) ≤ ε ≤ kµk. 12 JOHN J. BENEDETTO AND WEILIN LI Further, if there exists an extrapolation ν ∈ M (Td) and kνk < kµk, then (2.3) kbµkℓ∞(Λ) ≤ ε ≤ kνk < kµk. Proof. To see the lower bound for ε in (2.2) and (2.3), let σ be a minimal extrapolation. Then, The upper bounds, ε ≤ kµk and ε ≤ kνk in (2.2) and (2.3), follow by definition of ε. sup m∈Λbµ(m) = sup m∈Λ bσ(m) ≤ kσk = ε. (cid:3) kµk kνk kbµkℓ∞(Λ) -3 -2 -1 1 2 3 0 Λ whereas Figure 2.2. The black points represent the values of bµ on Λ. The union of the light and dark gray regions is the admissibility range [kbµkℓ∞(Λ),kµk], [kbµkℓ∞(Λ),kνk]. the dark gray region is an improved admissibility range Remark 2.5. By tightening the admissibility range for ε, we can deduce information about Example 3.5. 3.2 and 3.3. The next simplest case is when there exists an extrapolation ν, such that better suited for deducing the impossibility of super-resolution reconstruction. Remarkably the minimal extrapolations. The simplest case is when ε = kbµkℓ∞(Λ) = kµk, see Examples ε = kbµkℓ∞(Λ) = kνk, see Example 3.4. A more complicated case is when kbµkℓ∞(Λ) < ε, see Cand`es and Fernandez-Granda [CFG14] focused entirely on the case that kµk = ε because this is a necessary condition to super-resolve µ. In contrast to their result, our theory is better suited for situations when kbµkℓ∞ = ε. This is because Theorem 1.4b is strongest when bµ(m) = ε for many points m ∈ Λ, i.e., when #Γ is large. Hence, our theory is though, the following proposition shows that if Algorithm 1.3 fails, then ε = kbµkℓ∞ , and since we are given bµ Λ, we immediately get the exact value of ε. Recall that the algorithm fails if the function ϕ ∈ U computed in the first step of the algorithm satisfies hϕ, µi = ε and ϕ ≡ 1. Proposition 2.6. Let µ ∈ M (Td) and let Λ ⊆ Zd be a finite subset. Suppose there exists ϕ ∈ U such that ϕ ≡ 1 and hϕ, µi = ε. Then, Γ 6= ∅ and ε = kbµkℓ∞(Λ). Proof. By Proposition 2.2, we must have Γ 6= ∅. Let m ∈ Γ and by definition of Γ, we have bµ(m) = ε. This combined with the lower bound in Proposition 2.4 proves that ε = kbµkℓ∞(Λ). (cid:3) BEURLING SUPER-RESOLUTION 13 2.5. On uniqueness and non-uniqueness of minimal extrapolations. Theorem 1.4 provides sufficient conditions for the minimal extrapolations to be supported in a discrete set, but it does not provide sufficient conditions for uniqueness. Since a family of discrete measures supported on a common set behaves essentially like a vector, we use basic linear al- gebra to address the question of uniqueness when the minimal extrapolations are supported in a discrete set. Proposition 2.7. Let µ ∈ M (Td) and let Λ = {m1, m2, . . . , mJ} ⊆ Zd be a finite sub- set. Suppose there exists a finite set, {xk ∈ Td : k = 1, 2, . . . , K}, such that each minimal extrapolation is supported in this set. Suppose that the matrix,   e−2πim1·x1 e−2πim2·x1 e−2πim1·x2 e−2πim2·x2 ... e−2πimJ ·x1 ... e−2πimJ ·x2 e−2πim1·xK e−2πim2·xK ... e−2πimJ ·xK   , ··· ··· ··· (2.4) E(m1, . . . , mJ ; x1, . . . , xK ) = has full column rank (this can only occur if J ≥ K). Then, the minimal extrapolation is unique. Proof. Let ν be the difference of any two minimal extrapolations. Then ν is also supported in {xk : k = 1, 2, . . . , K} and it is of the form ν = PK k=1 akδxk . Since bν = 0 on Λ, we have 0 = bν(mj) =PK k=1 ake−2πimj ·xk for j = 1, 2, . . . , J, which is equivalent to the linear system Ea = 0, where a = (a1, . . . , aK ) ∈ CK. By assumption, E has full column rank. This implies a = 0. (cid:3) To finish this subsection, we address the situation when #Γ = 1, i.e., the missing case of Theorem 1.4. We say a measure µ ∈ M (Td) is a positive measure if µ(A) ≥ 0 for all Borel sets A ⊆ Td. A sequence {am ∈ C : m ∈ Zd} is a positive-definite sequence if for all sequences {bm ∈ C : m ∈ Zd} of finite support, we have am−nbmbn ≥ 0. Xm,n∈Zd For y ∈ Td, let Ty : M (Td) → M (Td) be the translation operator defined by Tyµ(x) = µ(x − y). For n ∈ Zd, let Mn : M (Td) → M (Td) be the modulation operator defined by Mnµ(x) = e2πin·xµ(x). Proposition 2.8. Let µ ∈ M (Td) and let Λ ⊆ Zd be a finite subset. Suppose there exists n ∈ Λ such that (M−nµ)∧ Λ−n extends to a positive-definite sequence on Zd. Then, n ∈ Γ, and each positive-definite extension of (M−nµ)∧ Λ−n corresponds to a positive measure ν, such that Mnν is a minimal extrapolation of µ from Λ. Proof. Extend (M−nµ)∧ Λ−n to a positive-definite sequence. By Herglotz' theorem, there exists a positive measure ν ∈ M (Td) such that bν = (M−nµ)∧ = T−nbµ on Λ− n. Then, Mnν is an extrapolation of µ from Λ, which shows that For the reverse inequality, since ν is a positive measure, we have kνk = bν(0). Then, kMnνk = kνk = bν(0) = bν(0) = bµ(n) ≤ kbµkℓ∞(Λ) ≤ ε(µ, Λ), where the last inequality follows by Proposition 2.4. This shows that bµ(n) = kνk = ε, which proves that n ∈ Γ and Mnν is a minimal extrapolation of µ from Λ. kνk = kMnνk ≥ ε(µ, Λ). (cid:3) 14 JOHN J. BENEDETTO AND WEILIN LI Remark 2.9. Beurling [Beu89b] and Esseen [Ess45] essentially proved the analogue of the Proposition for the special case that n = 0, for R instead of Td. Proposition 2.8 generalizes their result to handle situations when 0 6∈ Λ. This is important because from the viewpoint of Proposition 2.12c, the super-resolution problem is invariant under simultaneous translations of bµ and Λ, which means that 0 ∈ Zd is no more special than any other point n ∈ Zd. Remark 2.10. Proposition 2.8 suggests that the case #Γ = 1 is special compared to the cases #Γ = 0 or #Γ ≥ 2 because, when #Γ = 1, there may exist absolutely continuous minimal extrapolations. In Example 3.2, #Γ = 1, and there exist uncountably many discrete and positive absolutely continuous minimal extrapolations. sequence on Z. In theory, there are an infinite number of such extensions, and one particu- lar method of choosing such an extension is called the Maximum Entropy Method (MEM). Remark 2.11. Suppose that µ ∈ M (T) and bµ Λ can be extended to a positive-definite According to MEM, one extends bµ Λ to the positive-definite sequence {am ∈ C : m ∈ Z} whose corresponding density function f ∈ L1(T) is the unique maximizer of a specific loga- rithmic integral associated with the physical notion of entropy, e.g., see [Ben96, Theorems 3.6.3 and 3.6.6]. MEM is related to spectral estimation methods [Chi78], the maximum likelihood method [Chi78], and moment problems [Lan87]. 2.6. Basic properties of minimal extrapolation. Our next goal is to examine the symmetries of the minimal extrapolations. We are interested in the vector space operations, namely, addition of measures and the multiplication of measures by complex constants. We are also interested in the operations that are well-behaved under the Fourier transform on Td, namely, translation on the torus, modulation by integers, convolution of measures, and product of measures. Proposition 2.12. Let µ ∈ M (Td), Λ ⊆ Zd be a finite subset, c ∈ C non-zero, and y ∈ Td. (a) Multiplication by constants is bijective: c ε(µ, Λ) = ε(cµ, Λ), and ν ∈ E(µ, Λ) if and only if cν ∈ E(cµ, Λ). (b) Translation is bijective: ε(µ, Λ) = ε(Tyµ, Λ), and ν ∈ E(µ, Λ) if and only if Tyν ∈ E(Tyµ, Λ). (c) Minimal extrapolation is invariant under simultaneous shifts of bµ and Λ: ε(µ, Λ) = ε(Mnµ, Λ + n), and E(µ, Λ) = E(Mnµ, Λ + n). (d) The product of minimal extrapolations is a minimal extrapolation for the product: For j = 1, 2, let µj ∈ M (Tdj ), let Λj ⊆ Zdj be a finite subset, and let νj ∈ E(µj, Λj). Then ε(µ1, Λ1)ε(µ2, Λ2) = ε(µ1 × µ2, Λ1 × Λ2), and ν1 × ν2 ∈ E(µ1 × µ2, Λ1 × Λ2). Proof. (a) If ν ∈ E(µ, Λ), then cν is an extrapolation of cµ from Λ. Suppose cν 6∈ E(cµ, Λ). Then there exists σ such that bσ = cbµ on Λ and kσk < kcνk. Thus, bσ/c = bµ on Λ and kσ/ck < kνk, and this contradicts the assumption that ν ∈ E(µ, Λ). The converse follows by a similar argument. (b) If ν ∈ E(µ, Λ), then Tyν is an extrapolation of Tyµ from Λ. Suppose Tyν 6∈ E(Tyµ, Λ). Then there exists σ such that bσ = (Tyµ)∧ on Λ and kσk < kTyνk = kνk. Then (T−yσ)∧ = bµ on Λ and kT−yσk = kσk < kνk, which contradicts the assumption that ν ∈ E(µ, Λ). The converse follows by a similar argument. BEURLING SUPER-RESOLUTION 15 (c) If f ∈ U (Td; Λ), then Mnf ∈ U (Td; Λ + n). By Parseval's theorem, hf, µi = Xm∈Λ bf (m)bµ(m) = Xm∈Λ+n (Mnf )∧(m)(Mnµ)∧(m) = hMnf, Mnµi. Using Proposition 2.1e, we see that ε(µ, Λ) = ε(Mnµ, Λ + n). ε(µ, Λ) = ε(Mnµ, Λ + n). The converse follows similarly. If ν ∈ E(µ, Λ), then Mnν is an extrapolation of Mnµ from Λ + n, and kMnk = kνk = (d) For convenience, let µ = µ1 × µ2, ν = ν1 × ν2, Λ = Λ1 × Λ2, εj = ε(µj, Λj), and ε = ε(µ, Λ). Since ν is an extrapolation of µ, by Proposition 2.4, we have ε ≤ kνk = kν1kkν2k = ε1ε2. To see the reverse inequality, we use Proposition 2.1f, and see that there exist ϕj ∈ U (Tdj ; Λj) such that εj = hϕj , µji, for j = 1, 2. Let ϕ = ϕ1 ⊗ ϕ2 and observe that ϕ ∈ U (Td; Λ). By Proposition 2.1e, we have ε = max f ∈U (Td;Λ)hf, µi ≥ hϕ, µi ≥ hϕ, µi = hϕ1, µ1ihϕ2, µ2i = ε1ε2. This shows that ε = ε1ε2, and, since kνk = ε1ε2, we conclude that ν ∈ E(µ, Λ). (cid:3) While minimal extrapolation is well-behaved under translation, it not well behaved under modulation. This is because the Fourier transform of modulation is translation, and so bµ Λ and (Mnµ)∧ Λ are, in general, not equal. In contrast, the Fourier transform of translation is modulation, and so bµ Λ and (Tyµ)∧ Λ only differ by a phase factor. We shall prove these statements in Proposition 2.13. Proposition 2.13. (a) For j = 1, 2, there exist µj ∈ M (T), a finite subset Λ ⊆ Z, and νj ∈ E(µj, Λ), such that (b) There exist µ ∈ M (T), a finite subset Λ ⊆ Z, ν ∈ E(µ, Λ), and n ∈ Z, such that (c) For j = 1, 2, there exist µj ∈ M (T), a finite subset Λ ⊆ Z, and νj ∈ E(µj, Λ), such that ν1 + ν2 6∈ E(µ1 + µ2, Λ). Mnν 6∈ E(Mnµ, Λ). ν1 ∗ ν2 6∈ E(µ1 ∗ µ2, Λ). Proof. (a) Let µ1 = δ0 + δ1/2, µ2 = −δ0 − δ1/2, and Λ = {−1, 0, 1}. By Example 3.2, we have ν1 = δ0 + δ1/2 ∈ E(µ1, Λ), and ν2 = −δ1/4 − δ3/4 ∈ E(µ2, Λ). Then, µ1 + µ2 = 0, and so ε(µ1 + µ2, Λ) = 0. However, ν1 + ν2 6∈ E(µ1 + µ2, Λ) because kν1 + ν2k = kδ0 − δ1/4 + δ1/2 − δ3/4k > 0. (b) Let µ = δ0+δ1/2, Λ = {−1, 0, 1}, and n = −1. By Example 3.2, we have ν = δ1/4+δ3/2 ∈ E(µ, Λ). However, M−1µ = δ0 − δ1/2, and by Example 3.3, E(M−1µ, Λ) = {δ0 − δ1/2}. Thus, M−1ν 6∈ E(M−1µ, Λ). (c) Let µ1 = δ0 + δ1/2, µ2 = δ0 − δ1/2, and Λ = {−1, 0, 1}. Then (µ1 ∗ µ2)∧ = 0 on Λ, which implies ε(µ1 ∗ µ2, Λ) = 0. By Examples 3.2 and 3.3, ν1 = µ1 ∈ E(µ1, Λ) and ν2 = µ2 ∈ E(µ2, Λ). However, ν1 ∗ ν2 6∈ E(µ1 ∗ µ2, Λ) because kν1 ∗ ν2k = kδ0 − δ1k > 0. (cid:3) 16 JOHN J. BENEDETTO AND WEILIN LI 3. Examples 3.1. Discrete measures. There are several reasons why we are interested in computing the minimal extrapolations of discrete measures. They are the simplest types of measures, and so, their minimal extrapolations can be computed rather easily. By Theorem 1.4, the minimal extrapolations of a non-discrete measure are sometimes discrete measures, so they appear naturally in our analysis. As discussed in Section 1.4, examples of µ that have minimal extrapolations supported in a lattice can be interpreted in the context of deterministic compressed sensing. Examples 3.2-3.5 and Example 3.7 can be written in the context of compressed sensing. kδx′ k=2 a′ k=2 a′ k < ∞. k ∈ M (Td), where P∞ k=1 akδxk ∈ M (Td), where P∞ Remark 3.1. In view of Proposition 2.12 a,b, and without loss of generality, we can assume k=1 ak < ∞, can be written as any discrete measure µ = P∞ µ = δ0 +P∞ Example 3.2. Let µ = δ0 + δ1/2, and Λ = {−1, 0, 1}. We have bµ(0) = 2, and bµ(±1) = 0. Clearly kbµkℓ∞(Λ) = kµk = 2. By Proposition 2.4, ε = kbµkℓ∞(Λ) = kµk = 2, which implies µ ∈ E. for each y ∈ T and any integer K ≥ 2, define 2 K Further, there is an uncountable number of discrete minimal extrapolations. To see this, νy,K = δy+ k K . K−1Xk=0 A straightforward calculation shows that νy,K is an extrapolation and that kνy,Kk = ε. Also, we can construct positive absolutely continuous minimal extrapolations. One exam- ple is the constant function f ≡ 2 on T. For other examples, let N ≥ 2 and let FN ∈ C ∞(T) be the Fej´er kernel, FN (x) = NXn=−N(cid:16)1 − n N + 1(cid:17)e2πinx. For any c > 0 such that c ≤ (2N + 2)/(3N + 1), extend bµ Λ to the sequence {(aN,c)m : m ∈ Z}, where m = 0, N +1(cid:17) 2 ≤ m ≤ N, otherwise. NXn=2 (aN,c)ne2πinx N + 1(cid:17)e2πinx + c NXn=2(cid:16)1 − n N + 1(cid:17)e2πinx. Consider the real-valued function 2 (aN,c)m = c(cid:16)1 − m  0 (aN,c)ne2πinx + −2Xn=−N −2Xn=−N(cid:16)1 − n fN,c(x) = 2 + = 2 + c We check that dfN,c(m) = (aN,c)m for all m ∈ Z, which implies fN,c is an extrapolation of µ. Using the upper bound on c, we have, for all x ∈ T, that N + 1(cid:17)e−2πix. N + 1(cid:17) cos(2πx) = c + c(cid:16)1 − N + 1(cid:17)e2πix + c(cid:16)1 − 2 ≥ c + 2c(cid:16)1 − 1 1 1 Using this inequality, and the definitions of FN and fN,c, we have BEURLING SUPER-RESOLUTION 17 fN,c(x) ≥ cFN ≥ 0. fN,c(x) dx = 2 +ZT(cid:16) −2Xn=−N Since fN,c ≥ 0, we also have kfN,ck1 =ZT (aN,c)ne2πinx(cid:17) dx = 2 = ε. NXn=2 Thus, for any N ≥ 2 and c ≤ (2N + 2)/(3N + 1), fN,c is a positive absolutely continuous minimal extrapolation. Hence, we have constructed an uncountable number of positive absolutely continuous minimal extrapolations. (aN,c)ne2πinx + Example 3.3. Let µ = δ0 − δ1/2, and Λ = {−1, 0, 1}. We have bµ(0) = 0, and bµ(±1) = 2. Further, we have kbµkℓ∞(Λ) = kµk = 2, so that by Proposition 2.4, we have ε = kbµkℓ∞(Λ) = kµk = 2 and µ ∈ E. Consequently, Γ = {−1, 1}. By Theorem 1.4b, there exists α−1,1 ∈ R/Z satisfying e2πiα−1,1 = bµ(−1) bµ(1) = 1, and the minimal extrapolations are supported in the set {x ∈ T : 2x ∈ Z} = {0, 1/2}. This implies each ν ∈ E is discrete and can be written as ν = a1δ0 + a2δ1/2. In theory, a1, a2 depend on ν, so we cannot conclude uniqueness yet. The matrix E from (2.4) is E(−1, 0, 1; 0, 1/2) =    . 1 −1 1 1 1 −1 Clearly, E has full column rank, so, by Proposition 2.7, µ is the unique minimal extrapola- tion. Thus, super-resolution reconstruction of µ from Λ is possible. Example 3.4. Let µ = δ0−δ1/4, and let Λ = {−1, 0, 1}. We have bµ(±1) = 1±i = √2e±πi/4, and bµ(0) = 0. Note that kbµkℓ∞(Λ) = √2 < 2 = kµk, which shows that √2 ≤ ε ≤ 2. We claim that ε = √2. To see this, consider ν = (−δ3/8 + δ7/8)/√2. We verify that kνk = √2 and that ν is an extrapolation. By Proposition 2.4, ε = √2 and ν ∈ E. This also implies µ 6∈ E, and so super-resolution reconstruction of µ from Λ is impossible. We claim that ν is the unique minimal extrapolation. The matrix E from (2.4) is E(−1, 0, 1; 3/8, 7/8) =  e2πi3/8 e2πi7/8 1 1 e−2πi3/8 e−2πi7/8   , which we observe to have full column rank. By Proposition 2.7, we conclude that ν is the unique minimal extrapolation. Thus, super-resolution reconstruction of ν from Λ is possible. We explain the derivation of ν. We guess that ε = √2 and see what Theorem 1.4b implies. Under this assumption that ε = √2, we have Γ = {−1, 1}. By Theorem 1.4b, there exists α−1,1 ∈ R/Z satisfying e2πiα1,−1 = bµ(1) bµ(−1) = eπi/2, 18 JOHN J. BENEDETTO AND WEILIN LI fact, a minimal extrapolation. and the minimal extrapolations are supported in {x ∈ T : 2x + 1/4 ∈ Z} = {3/8, 7/8}. Hence, if ε = √2, then every σ ∈ E is of the form σ = a1δ3/8 + a2δ7/8. Thus, by definition of a minimal extrapolation, kσk = √2 and bµ = bσ on Λ. Using this information, we solve for the coefficients a1, a2, and compute that a1 = a2 = √2, a1 = −a2, and a1 = −√2/2. Thus, we obtain that σ = (−δ3/8 + δ7/8)/√2. From here, we simply check that ν = σ is, in Example 3.5. Let µ = δ0 + eπi/3δ1/3 and let Λ = {−1, 0, 1}. We have bµ(−1) = 0, bµ(0) = 1 + eπi/3 = √3 eπi/6, and bµ(1) = 1 + e−πi/3 = √3 e−πi/6. Suppose, for the purpose of obtaining a contradiction, that ε = kbµkℓ∞(Λ) = √3. Then Γ = {0, 1}. By Theorem 1.4b, there is α0,1 ∈ R/Z such that = eπi/3, e2πiα0,1 = bµ(0) bµ(1) and each ν ∈ E is of the form ν = aδ1/6 for some a ∈ C. Then, bν = a on Z and, in particular, bµ 6= bν on Λ, which is a contradiction. Thus, ε > √3, i.e., Γ = ∅. Therefore, Theorem 1.4a applies, and so there is a finite set S such that supp(ν) ⊆ S for each ν ∈ E. In particular, each ν ∈ E is discrete. Hence, we have to solve the optimization problem given in Proposition 2.1e, which is ε = maxn(cid:12)(cid:12)(cid:12)a√3 eπi/6 + b√3 e−πi/6(cid:12)(cid:12)(cid:12) : ∀x ∈ T, ae2πix + b + ce−2πix ≤ 1, a, b, c ∈ Co. This optimization problem can be written as a semi-definite program, see [CFG14, Corollary 4.1, page 936]. After obtaining numerical approximations to the optimizers of this problem, we guess that the the exact optimizers are a = e−πi/6, b = 2 3√3 4 3√3 eπi/6, c = − i 3√3 . These values of a, b, c are, in fact, the optimizers because a√3 eπi/6 + b√3 e−πi/6 = 2 and ae2πix +b+ce−2πix ≤ 1 for all x ∈ T. Thus, ε = 2 and µ ∈ E. Since ae2πix +b+ce−2πix = 1 precisely on S = {0, 1/3}, by Theorem 1.4a, the minimal extrapolations are supported in S. The matrix E from (2.4) is E(−1, 0, 1; 0, 1/3) =  1 e−2πi/3 1 1 e2πi/3 1   . Since E has full rank, by Proposition 2.7, µ is the unique minimal extrapolation. Thus, super-resolution reconstruction of µ from Λ is possible. The following example illustrates that if µ is a sum of two Dirac measures, then their supports have to be sufficiently spaced apart in order for super-resolution of µ to be possible. This shows that, in general, a minimum separation condition is necessary to super-resolve a sum of two Dirac measures, see [CFG14]. Example 3.6. Let µy = δ0 − δy for some non-zero y ∈ Td and let Λ ⊆ Zd be a finite subset. We claim that if y is sufficiently small depending on Λ, then µy 6∈ E(µy, Λ). Note that kµyk = 2 for any y ∈ Td. Let η denote the normalized Lebesgue measure on Td and define the measures νy by the formula BEURLING SUPER-RESOLUTION 19 νy(x) = Xm∈Λcµy(m)e2πim·x η(x). By construction, νy is an extrapolation of µy because, for each n ∈ Λ, e−2πin·x dνy(x) = Xm∈Λcµy(m)ZTd bνy(n) =ZTd kνyk =ZTd(cid:12)(cid:12)(cid:12) Xm∈Λcµy(m)e2πim·x(cid:12)(cid:12)(cid:12) dx =ZTd(cid:12)(cid:12)(cid:12) Xm∈Λ e−2πi(n−m)·x dx = cµy(n). (1 − e−2πim·y)e2πim·x(cid:12)(cid:12)(cid:12) dx → 0. Then, as y → 0, Thus, we take y sufficiently small so that kνyk < 2 = kµyk, and, hence, µy 6∈ E(µy, Λ). Note that this argument does not contradict Proposition 2.4 because kcµykℓ∞(Λ) → 0 as y → 0. Thus, for y sufficiently small, super-resolution reconstruction of µy from Λ is impossible. 3.2. Singular continuous measures. The following example is an analogue of Example 3.2 for higher dimensions. Example 3.7. Let µ = δ(0,0) + δ(1/2,1/2), and Λ = {−1, 0, 1}2 \ {(1,−1), (−1, 1)}. Then, bµ(m) = 1 + e−πi(m1+m2), and, in particular, bµ(1, 1) = bµ(−1,−1) = bµ(0, 0) = 2 and bµ(±1, 0) = bµ(0,±1) = 0. We deduce that ε = kµk = kbµkℓ∞(Λ) = 2 from Proposition 2.4, and so µ ∈ E. Further, Γ = {(0, 0), (1, 1), (−1,−1)}, and so #Γ = 3. According to the definition of αm,n in Theorem 1.4b, set α(−1,−1),(0,0) = α(0,0),(1,1) = α(−1,−1),(1,1) = 0. By the conclusion of Theorem 1.4b, the minimal extrapolations are supported in the set S = S(−1,−1),(0,0) ∩ S(0,0),(1,1) ∩ S(−1,−1),(1,1), where S(−1,−1),(0,0) = {x ∈ T2 : x · (−1,−1) ∈ Z} = {x ∈ T2 : x1 + x2 ∈ Z}, S(0,0),(1,1) = {x ∈ T2 : x · (−1,−1) ∈ Z} = {x ∈ T2 : x1 + x2 ∈ Z}, S(−1,−1),(1,1) = {x ∈ T2 : x · (−2,−2) ∈ Z} = {x ∈ T2 : 2x1 + 2x2 ∈ Z}. It follows that the minimal extrapolations are supported in S = S(−1,−1),(0,0) ∩ S(0,0),(1,1) ∩ S(−1,−1),(1,1) = {x ∈ T2 : x1 + x2 = 1}. We can construct other discrete minimal extrapolations besides µ. For each y ∈ T and for each integer K ≥ 2, define the measure K−1Xk=0 νy,K = 2 K δ(cid:0)y+ k K ,1−y− k K(cid:1). We claim νy,K is a minimal extrapolation. We have kνy,Kk = ε, and dνy,K(m) = e−2πi(m1y−m2y)e−2πim2 2 We see that dνy,K = bµ on Λ, which proves the claim. K K−1Xk=0 e−2πi(m1−m2)k/K. 20 JOHN J. BENEDETTO AND WEILIN LI We can also construct continuous singular minimal extrapolations. Let σ = √2σS, where σS is the surface measure of the Borel set S. We readily verify that kσk = ε and e−2πi(m1−m2)t dt = 2δm1,m2, e−2πim·x dσS = 2e−2πim2Z 1 0 bσ(m) = √2ZT2 which proves that σ ∈ E. In particular, S is the smallest set that contains the support of all the minimal extrapolations. Since µ is not the unique minimal extrapolation, super-resolution reconstruction of µ from Λ is impossible. Example 3.8. For an integer q ≥ 3, let Cq be the middle 1/q-Cantor set, which is defined by Cq =T∞ k=0 Cq,k, where Cq,0 = [0, 1] and Cq,k Cq,k Cq,k+1 = q ∪(cid:16)(1 − q) + q (cid:17). Let Fq : [0, 1] → [0, 1] be the Cantor-Lebesgue function on Cq, which is defined by the point-wise limit of the sequence {Fq,k}, where Fq,0(x) = x and Fq,k+1(x) =  1 2 Fq,k(qx) 1 2 1 2 Fq,k(qx − (q − 1)) + 1 2 1 0 ≤ x ≤ 1 q , q ≤ x ≤ q−1 q , q ≤ x ≤ 1. q−1 By construction, Fq(0) = 0, Fq(1) = 1, and Fq is non-decreasing and uniformly continuous on [0, 1]. Thus, Fq can be uniquely identified with the measure σq ∈ M (T), and kσqk = 1. Since F ′ q = 0 a.e. and Fq does not have any jump discontinuities, σq is a continuous singular measure, with zero discrete part. The Fourier coefficients of σq are see [Zyg59, pages 195-196]. In particular, for any integer n ≥ 1, we have bσq(m) = (−1)m ∞Yk=1 cos(πmq−k(1 − q)), bσq(qn) = (−1)qn ∞Yk=1 cos(πq−k(1 − q)), which is convergent and independent of n. Since bσq(0) = kσqk = 1, we immediately see that ε = 1 and σq ∈ E. Again, we cannot determine whether σq is the unique minimal extrapolation because Theorem 1.4 cannot handle the case #Γ = 1, see Remark 2.10. Example 3.9. Let σA, σB ∈ M (Td) be the surface measures of the Borel sets A = {x ∈ T2 : x2 = 0} and B = {x ∈ T2 : x2 = 1/2}, respectively. Let µ = σA + σB, and Λ = {−2,−1, . . . , 1}2. Then, bµ(m) =Z 1 0 e2πim1t dt +Z 1 0 e2πi(m1t+m2/2) dt = δm1,0 + (−1)m2 δm1,0. We immediately see that ε = kbµkℓ∞(Λ) = kµk = 2, which implies µ ∈ E. Then, Γ = {(0, 0), (0, 2), (0, −2)}, and, by Theorem 1.4b, the minimal extrapolations are supported in {x ∈ T2 : x2 = 0} ∪ {x ∈ T2 : x2 = 1/2}. Determining whether µ is the unique minimal extrapolation is beyond the theory we have developed herein, and we shall examine this uniqueness problem in [BL16]. BEURLING SUPER-RESOLUTION 21 4. Acknowledgements The first named author gratefully acknowledges the support of ARO Grant W911 NF- 15-1-0112, ARO Grant W911 NF-16-1-0008, and DTRA Grant 1-13-1-0015. [ADCG15] Jean-Marc Azaıs, Yohann De Castro, and Fabrice Gamboa. Spike detection from inaccurate References [ASB14] [ASB15] [AYB15] [BC09] samplings. Applied and Computational Harmonic Analysis, 38(2):177 -- 195, 2015. C´eline Aubel, David Stotz, and Helmut Bolcskei. Super-resolution from short-time Fourier trans- form measurements. In IEEE International Conference on Acoustics, Speech and Signal Process- ing, pages 36 -- 40. IEEE, 2014. C´eline Aubel, David Stotz, and Helmut Bolcskei. A theory of super-resolution from short-time Fourier transform measurements. arXiv preprint arXiv:1509.01047, 2015. Enrico Au-Yeung and John J. Benedetto. Generalized Fourier frames in terms of balayage. Journal of Fourier Analysis and Applications, 21(3):472 -- 508, 2015. John J. Benedetto and Wojciech Czaja. Integration and Modern Analysis. Birkhauser Boston, Inc., Boston, 2009. John J. Benedetto. Spectral Synthesis. Academic Press, Inc., New York-London, 1975. John J. Benedetto. Harmonic Analysis and Applications. CRC Press Inc., 1996. [Ben75] [Ben96] [Beu89a] Arne Beurling. Balayage of Fourier-Stieltjes transforms. The Collected Works of Arne Beurling, 2:341 -- 350, 1989. [Beu89b] Arne Beurling. Interpolation for an interval in R1. The Collected Works of Arne Beurling, 2:351 -- [BL16] [BTR13] [CFG13] [CFG14] 365, 1989. John J. Benedetto and Weilin Li. Super-resolution of singular continuous measures on Td. In Preparation, 2016. Badri Narayan Bhaskar, Gongguo Tang, and Benjamin Recht. Atomic norm denoising with applications to line spectral estimation. IEEE Transactions on Signal Processing, 61(23):5987 -- 5999, 2013. Emmanuel J. Cand`es and Carlos Fernandez-Granda. Super-resolution from noisy data. Journal of Fourier Analysis and Applications, 19(6):1229 -- 1254, 2013. Emmanuel J. Cand`es and Carlos Fernandez-Granda. Towards a mathematical theory of super- resolution. Communications on Pure and Applied Mathematics, 67(6):906 -- 956, 2014. Donald G. Childers. Modern Spectrum Analysis. IEEE Press, Piscataway, New Jersey, 1978. [Chi78] [CRPW12] Venkat Chandrasekaran, Benjamin Recht, Pablo A. Parrilo, and Alan S. Willsky. The convex geometry of linear inverse problems. Foundations of Computational mathematics, 12(6):805 -- 849, 2012. Emmanuel J. Cand`es, Justin K. Romberg, and Terence Tao. Robust uncertainty principles: Exact signal reconstruction from highly incomplete frequency information. IEEE Trans. Infor- mation Theory, 52(2):489 -- 509, 2006. [CRT06] [DCG12] Yohann De Castro and Fabrice Gamboa. Exact reconstruction using Beurling minimal extrapo- [Don92] [Don06] [DP15] [Dum07] [Ess45] [Ess54] [FG16] lation. Journal of Mathematical Analysis and Applications, 395(1):336 -- 354, 2012. David L. Donoho. Superresolution via sparsity constraints. SIAM Journal on Mathematical Anal- ysis, 23(5):1309 -- 1331, 1992. David L. Donoho. Compressed sensing. IEEE Trans. Information Theory, 52(4):1289 -- 1306, 2006. Vincent Duval and Gabriel Peyr´e. Exact support recovery for sparse spikes deconvolution. Foun- dations of Computational Mathematics, 15(5):1315 -- 1355, 2015. Bogdan Dumitrescu. Positive Trigonometric Polynomials and Signal Processing Applications. Springer, Dordrecht, 2007. Carl-Gustav Esseen. Fourier analysis of distribution functions. a mathematical study of the Laplace-Gaussian law. Acta Mathematica, 77(1):1 -- 125, 1945. Carl-Gustav Esseen. A note on Fourier-Stieltjes transforms and absolutely continuous functions. Math. Scand, 2:153 -- 157, 1954. Carlos Fernandez-Granda. Super-resolution of point sources via convex programming. Informa- tion and Inference, pages 1 -- 53, 2016. 22 [Gre09] [HR55] [HT29] [Kah70] JOHN J. BENEDETTO AND WEILIN LI Hayit Greenspan. Super-resolution in medical imaging. The Computer Journal, 52(1):43 -- 63, 2009. Edwin Hewitt and Herman Rubin. The maximum value of a Fourier-Stieltjes transform. Math. Scand, 3:97 -- 102, 1955. Einar Hille and J.D. Tamarkin. Remarks on a known example of a monotone continuous function. American Mathematical Monthly, 36(5):255 -- 264, 1929. Jean-Pierre Kahane. S´eries de Fourier Absolument Convergentes. Springer-Verlag, Berlin-New York, 1970. [KLM04] Valery Khaidukov, Evgeny Landa, and Tijmen Jan Moser. Diffraction imaging by focusing- defocusing: An outlook on seismic superresolution. Geophysics, 69(6):1478 -- 1490, 2004. [KMM07] Aggelos K. Katsaggelos, Rafael Molina, and Javier Mateos. Super Resolution of Images and [Lan87] [Lin12] [PK05] [PPK03] Video. Morgan & Claypool Publishers, 2007. Henry J. Landau. Maximum entropy and the moment problem. Bulletin of the American Math- ematical Society, 16(1):47 -- 77, 1987. Jari Lindberg. Mathematical 14(8):083001, 2012. K.G. Puschmann and F. Kneer. On super-resolution in astronomical imaging. Astronomy & Astrophysics, 436(1):373 -- 378, 2005. Sung Cheol Park, Min Kyu Park, and Moon Gi Kang. Super-resolution image reconstruction: a technical overview. Signal Processing Magazine, IEEE, 20(3):21 -- 36, 2003. Fred Rieke. Spikes: Exploring the Neural code. MIT press, 1999. superresolution. Journal of Optics, concepts of optical [Rie99] [Rud62] Walter Rudin. Fourier Analysis on Groups. John Wiley & Sons, Inc., New York, 1962. [Sal42] Raphael Salem. On singular monotonic functions of the Cantor type. J. Math. Phys, 21:69 -- 82, 1942. David Slepian. Prolate spheroidal wave functions, Fourier analysis, and uncertainty - V: The discrete case. Bell System Technical Journal, 57(5):1371 -- 1430, 1978. [Sle78] [TBR15] Gongguo Tang, Badri Narayan Bhaskar, and Benjamin Recht. Near minimax line spectral esti- mation. IEEE Transactions on Information Theory, 61(1):499 -- 512, 2015. [TBSR13] Gongguo Tang, Badri Narayan Bhaskar, Parikshit Shah, and Benjamin Recht. Compressed sens- [Zyg59] ing off the grid. IEEE Trans. Information Theory, 59(11):7465 -- 7490, 2013. Antoni Zygmund. Trigonometric Series, volume 1. Cambridge University Press, 2002 (1959). Norbert Wiener Center, Department of Mathematics, University of Maryland, College Park, MD 20742, USA E-mail address: [email protected] Norbert Wiener Center, Department of Mathematics, University of Maryland, College Park, MD 20742, USA E-mail address: [email protected]
1010.0999
3
1010
2015-09-14T07:57:21
On continuity of measurable group representations and homomorphisms
[ "math.FA", "math.GN" ]
Let G be a locally compact group, and let U be its unitary representation on a Hilbert space H. Endow the space L(H) of linear bounded operators on H with weak operator topology. We prove that if U is a measurable map from G to L(H) then it is continuous. This result was known before for separable H. To prove this, we generalize a known theorem on nonmeasuralbe unions of point finite families of null sets. We prove also that the following statement is consistent with ZFC: every measurable homomorphism from a locally compact group into any topological group is continuous. This relies, in turn, on the following theorem: it is consistent with ZFC that for every null set S in a locally compact group there is a set A such that AS is non-measurable.
math.FA
math
ON CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS AND HOMOMORPHISMS YULIA KUZNETSOVA Abstract. Let G be a locally compact group, and let U be its unitary representation on a Hilbert space H. Endow the space L(H) of linear bounded operators on H with weak operator topology. We prove that if U is a measurable map from G to L(H) then it is continuous. This result was known before for separable H. We prove also that the follow- ing statement is consistent with ZFC: every measurable homomorphism from a locally compact group into any topological group is continuous. Let G be a locally compact group. We consider its unitary representa- tions, that is, homomorphisms U from G into the group U(H) of unitary operators on a Hilbert space H. One gets a rich representation theory if the representations considered are weakly continuous, i.e. such that for every x, y ∈ H the coefficient f (t) = hU(t)x, yi is a continuous function on G. This requirement is equivalent to strong continuity, i.e. continuity of the function F (t) = kU(t)xk for every x ∈ H. Representations satisfying any of these conditions will be further called continuous. In certain cases it happens that every representation is automatically continuous, as, notably, every finite dimensional unitary representation of a connected semisimple Lie group. This theorem was proved for compact groups by Van der Waerden [27] and in general case by A. I. Shtern [26]. But in general it is easy to construct discontinuous representations, so for automatic continuity, one has to assume some sort of measurability at least. A commonly used notion is as follows. Say that a representation U of a locally compact group G on a Hilbert space H is weakly measurable if every coefficient f (t) = hU(t)x, yi is a measurable function on G. Every weakly measurable unitary representation must be continuous if it acts on a separable Hilbert space [14, Theorem V.7.3]. However, in general this does not imply continuity: if G is non-discrete, then the regular representation of G on the space ℓ2(G) of countably sum- mable sequences on G is weakly measurable but discontinuous. In this paper we prove that separability restriction can be removed if we use a slightly stronger notion of measurability. Let L(H) be the space of bounded linear operators on the Hilbert space H, endowed with the weak operator topology (it is generated by the functions fxy for all x, y ∈ H, where fxy(A) = hAx, yi, Current address: University of Franche-Comté, 25030 Besançon Cedex, France. Email: [email protected] 2010 Mathematics Subject Classification. 22D10, 43A05, 28A05, 54H11. Key words and phrases. automatic continuity, group representations, group homomor- phisms, nonmeasurable unions, zero measure sets. 1 2 YULIA KUZNETSOVA A ∈ L(H)). Say that U is weakly operator measurable if U −1(V ) is mea- surable for every open set V ⊂ L(H). Now we can formulate the main result of this paper (Theorem 5): every weak operator measurable unitary representation of a locally compact group is continuous. The proof is based on a generalization of the so called Four Poles Theorem: if A is a point-finite family of null sets with non-null union in a Polish space, then there is a subfamily in A with a nonmeasurable union (this was proved initially by L. Bukovsky [5] and then much simpler by J. Brzuchowski, J. Cichoń, E. Grzegorek and C. Ryll-Nardzewski [4]). In Lemma 4, we prove the same result for subsets of any locally compact group, with a restriction that cardinality of A is not more than continuum. The second part of the paper deals with automatic continuity of more general group homomorphisms. Most actively this question is studied for homomorphisms between Polish groups, see a recent review of C. Rosendal [25]. The notion of Haar measurability of f : G → H is here replaced by universal measurability: the inverse image of every open set is measurable with respect to every Radon measure on G. It is known that every univer- sally measurable homomorphism from a locally compact or abelian Polish group into a Polish group, or from a Polish group to a metric group is con- tinuous. There are also generalizations to other subclasses of Polish groups by S. Solecki and Rosendal. We omit results on other types of measurability (in the sense of Souslin, Christensen etc.) If G is not supposed to be Polish, the results are fewer. The most general statement is probably the theorem of A. Kleppner [22]: every measurable homomorphism between two locally compact groups is continuous. It has been generalized to some special classes of groups by J. Brzdęk [3]. If one makes no assumptions on the image group, it seems inevitable to impose additional set-theoretic axioms instead. The only result in this direction known to the author, belongs to J. P. R. Christensen [6]: under Luzin’s hy- pothesis, every Baire, in particular, every Borel measurable homomorphism from a Polish group to any topological group is continuous. Our Theorem 10 is proved under Martin’s axiom (MA): every measurable homomorphism from a locally compact group to any topological group is continuous. Theorem 10 is reduced to the following question. Let G be a locally compact group; call a set A ⊂ G extra-measurable if SA is measurable for any S ⊂ G. An obvious example of an extra-measurable set is an open set. Existence of discontinuous measurable homomorphisms implies existence of null extra-measurable sets; but under MA, as Theorem 9 shows, the latter do not exist, so every measurable homomorphism is continuous. In the commutative case, the question of automatic continuity is even equivalent to the existence of a certain sequence of null extra-measurable sets (Theorem 7). 1. Continuity of unitary representations Definitions and notations. On a locally compact group G, we fix a left Haar measure µ and the corresponding outer measure µ∗. A map CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS 3 f : G → Y , where Y is a topological space, is called measurable if f −1(Y ) is Haar measurable for every open set U ⊂ Y . For a set A, A denotes its cardinality. There are two approaches to the construction of Haar measures. One, used by E. Hewitt and K. Ross [16], yields an outer regular measure: for every measurable set E, one has µ(E) = inf{µ(U) : E ⊂ U, U open}. Another one, accepted by D. Fremlin [13], leads to an inner regular measure: µ(E) = sup{µ(F ) : F ⊂ E, F compact}. In the σ-finite case, in particular, on a σ-compact group, both construc- tions give the same resulting measure, which is both inner and outer regular. If G is not σ-compact, the approach of [16] gives rise to the following patho- logical sets. A set A ⊂ G is called locally null [16, 11.26] if µ(A ∩ K) = 0 for every compact set K ⊂ G. Equivalently, A does not contain any set of positive finite measure. Of course, if A is null then A is locally null. Every locally null set A is measurable, and either µ(A) = 0 or µ(A) = ∞. In the Fremlin’s treatment, there are no locally null non-null sets. The results of this paper, in particular Theorem 5, are valid for both definitions of the Haar measure. It is known [14, IV.2.16 and V.7.2] that every unitary representation of a locally compact group may be decomposed into a direct sum U = U1 ⊕ U2, where U1 is continuous and every coefficient of U2 is (locally) almost everywhere zero. We will say that U2 is singular. If U acts on a separable space then U2 = 0 [14, Theorem V.7.3]. Let U act on a Hilbert space H. Endow the space L(H) of bounded linear operators on H with the weak operator topology (generated by the functions fxy for all x, y ∈ H, where fxy(A) = hAx, yi, A ∈ L(H)). If U is a measurable map from G to L(H), we will say that U is weakly operator measurable. The following lemma is known [13, 443P]. Lemma 1. Let G be a σ-compact locally compact group, K its compact normal subgroup, and let π : G → G/K be the quotient map. If A ⊂ G is such that A = AK then A is measurable (respectively null) in G if and only if π(A) is measurable (null) in G/K. The two following facts will be used in further proofs several times, so we prefer to state them separately. Remark 2 (Pro-Lie and Polish groups). Recall that a topological group is called pro-Lie if it is an inverse (projective) limit of (finite-dimensional) Lie groups (see [18]). It is known that in every locally compact group there is an open pro-Lie subgroup [18, p. 165]. If G is a locally compact group and G = lim Gi, where every Gi is a Lie group, then these groups can be chosen as Gi = G/Ki, where every Ki is a compact normal subgroup of G, and the order on I is just inclusion of Ki. Every σ-compact Lie group is Polish (it is first countable, hence metrizable [17, Theorem A4.16], and further apply [2, Chapter IX, § 6], Corollary of Proposition 2). If all Gi are σ-compact and I is countable, G is Polish too ([2, § 6, Proposition 1a,b]). ←− i∈I 4 YULIA KUZNETSOVA Remark 3 (Baire sets in direct products). Baire sets ([15, §51]) are the elements of the σ-algebra generated by all compact Gδ-sets. In the case of a σ-compact locally compact group (the only case we will need), this is also the σ-algebra generated by all zero sets of continuous functions (in [16, 11.1], this latter property is taken as a definition). Consider the direct product of a family of locally compact groups G = Qj∈J Gj. This is a topological group, which is not necessarily locally compact. Let G ⊂ G be a closed σ-compact subgroup. For any I ⊂ J let πI : G → Qj∈I Gj be the natural projection. We say that a set X ⊂ G depends on coordinates I ⊂ J if X = G ∩ π−1 I (πIX). If F = ∩ Un is a compact Gδ set in G, then for every n the open set Un can be chosen as a finite union of basic neighbourhoods in G, which depend on finite number of coordinates. It follows that every such F , and as a consequence every Baire set depends on a countable set of coordinates. Lemma 4. Let A = {As : s ∈ S} be a point finite family of null subsets of a σ-compact locally compact group G. If A 6 c and ∪A is non-null, then there is B ⊂ A such that ∪B is nonmeasurable. Proof. Let H ⊂ G be an open pro-Lie subgroup, which we can assume compactly generated. Since G is a countable union of H-cosets, there is t0 ∈ G such that (∪A) ∩ t0H is non-null. Define A′ 0 As) ∩ H for s : s ∈ S} satisfies all conditions of the all s, then the family A′ = {A′ s : s ∈ T } is theorem and is contained in H. Moreover, if a union ∪{A′ nonmeasurable, then so is ∪{As : s ∈ T }. Thus, we can assume that Gi is σ-compact and pro-Lie. In this case every G = H, i.e. G = lim Gi = G/Ki is a σ-compact Lie group, hence Polish. We assume that every Gi is nontrivial, otherwise G = {1} and the family A would not exist. s = (t−1 ←− i∈I (s) k . (s) k m (s) k k , m(s) k We can assume that S ⊂ R. Let Q = {qm : m ∈ N} be an enumeration of the rationals, and let Wmn = ∪{As : s − qm < 1/n}. For every s ∈ S, choose sequences n(s) < 1/n(s) k → ∞ while n(s) k+1 > n(s) for every s. Indeed, every point x ∈ As is contained in this intersection; if x /∈ As then x ∈ Ati for at most finite set of points ti 6= s; and every ti can be separated from s by some interval qm k for all k. Then As = ∩kWn k , so that x /∈ Wn so that s − qm − s < 1/n(s) k and n(s) (s) k m (s) k (s) k If one of the sets Wmn is nonmeasurable, the lemma is proved. Suppose that every Wmn is measurable. By [16, 19.30b], there exists a Baire set Bmn ⊂ Wmn such that Nmn = Wmn \ Bmn is null. Further, for every n, m there is a null Baire set N ′ mn. Then N is a Baire set, so Wmn \ N = Bmn \ N is Baire for all m, n. Let Wmn \ N depend on the countable set of coordinates Jmn. Then every As \ N = ∩ \ N) (Wn k depends on coordinates J = ∪mnJmn, and the set J is countable. mn ⊃ Nmn. Let N = ∪m,nN ′ (s) k m (s) k Extending J, if necessary, we can assume that the family {Kj : j ∈ J} is closed under finite intersections. Denote K = ∩j∈J Kj. Then G/K = G/Kj is a Polish group. Let π : G → G/K be the quotient map, lim ←− j∈J s = π(As \ N). Then, since As \ N = (As \ N)K, the family and put A′ CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS 5 s : s ∈ S} is point finite, and by Lemma 1 we have that A′ s is null A′ = {A′ for all s, while ∪A′ = π(∪A) is non-null. By the Four Poles Theorem for the Polish case [4] we get B′ ⊂ A′ such that ∪B′ is nonmeasurable. Put s ∈ B′}. Then ∪B \ N = π−1(∪B′) is nonmeasurable, so B is as B = {As : A′ desired. (cid:3) A simple example shows that in the Hewitt&Ross approach, this theorem is not true for a non-σ-compact group. Let Rd be the real line with the discrete topology, and consider the direct product Rd × R. Then X = Rd × {0} is measurable of infinite measure (this is an example of a locally null, non-null set [16, 11.33]). Every uncountable subset of X is also measurable with infinite measure, and every countable subset is null. Thus, if we put At = {(t, 0)} and A = {At : t ∈ Rd}, then every At is null, ∪A is non- null but every subfamily of A has a measurable union. In the Fremlin’s approach, this example does not appear. Theorem 5. Let G be a locally compact group. Then every weakly operator measurable unitary representation of G is continuous. Proof. Let U : G → L(H) be a representation acting on a Hilbert space H. Clearly U is continuous if and only if its restriction to any open subgroup is continuous. In G, there is an open compactly generated pro-Lie subgroup (e.g., the intersection of an open pro-Lie subgroup and the subgroup gener- ated by a pre-compact neighbourhood of identity). So we can assume that G/Ki is itself compactly generated and pro-Lie; in particular, G G = lim is σ-compact. ←− i∈I Take any x ∈ H, kxk = 1. Put f (t) = hU(t)x, xi and S = {t ∈ G : f (t) 6= 0}. We can assume that U is singular, then S is null. Suppose towards a contradiction that U 6≡ 0, then e ∈ S. Let us show that S has a null projection onto a Polish quotient of G. By [16, 19.30b], there exists a null Baire set B ⊃ S. Every Baire set (Remark 3) depends on a countable number of coordinates. Let J ⊂ I be a countable set such that B = π−1 J πJ B. By extending J if necessary, we can assume that the family {Kj : j ∈ J} is closed under finite intersections. Denote K = ∩j∈J Kj. Then G/K = lim G/Kj is a Polish group. Let π : G → G/K be the quotient map and let S ′ = π(S), then S ′ ⊂ π(B) is null. ←− j∈J Choose an enumeration (probably with repetitions) Pα : α < c of perfect non-null sets in G/K. It is known that there is at most continuum such sets. By induction, we will choose points tα : α < n with some ordinal n 6 c so that ∪{tαS ′ : α < n} is non-null (in G/K) and tα /∈ ∪{tβS ′ : β < α} for every α > 0. Set t0 = e. For every α, let Tα = {tβ : β < α}. If TαS ′ is non-null, stop the procedure. Otherwise Pα \TαS ′ 6= ∅, and choose tα as any point of this set. Let n be the ordinal on which we stopped the induction, or n = c if it was not stopped. Set T = Tn. If n < c then as assumed µ∗(T S ′) > 0; if n = c then T S ′ intersects every non-null perfect set in G/K, so it is of full measure. In either case T S ′ is non-null. 6 YULIA KUZNETSOVA For every α < n, choose any zα ∈ π−1(tα) and set Z = {zα : α < n}. It follows that ZSK = π−1(π(ZS)) = π−1(T S ′) is non-null in G. Recall that K is a normal subgroup, so ZSK = ZKS. Define now Sn = {t ∈ G : f (t) = hU(t)x, xi > 1/n}. (1) Then S = ∪nSn and ZKS = ∪nZKSn. It follows that ZKSN is non-null for some N ∈ N. We claim that the family A = {zαKSN : α < n} is point-finite. To prove this, first show that U(zαk2)x ⊥ U(zβk1)x for any α 6= β and any k1, k2 ∈ K. Suppose that α > β. Then (zβk1)−1zαk2 /∈ S, because otherwise we would have zα ∈ zβk1Sk−1 2 ⊂ zβKSK = zβSK 2 = zβSK and hence tα = π(zα) ∈ π(zβSK) = tβS ′, what is impossible by the choice of tα. This gives us 0 = f ((zβk1)−1zαk2) = hU((zβk1)−1zαk2)x, xi = hU(zαk2)x, U(zβk1)xi, that is, U(zαk2)x ⊥ U(zβk1)x. Next, if t ∈ zαKSN then there is kα ∈ K such that (zaka)−1t ∈ SN, hU(t)x, U(zαkα)xi > 1/N. As we have shown above, U(zαkα)x are i.e. orthogonal for different α; since U is unitary, they have norm 1. By Bessel’s inequality we have for any t ∈ G: 1 = kxk2 = kU(t)xk2 > X hU(t)x, U(zαkα)xi2 α: t∈zαKSN > N −2 · {α : t ∈ zαKSN }. So A is point finite. It has cardinality A = Z = n 6 c and a non-null union ∪A = ZKSN . Every zαKSN ⊂ zαKS = zαπ−1(S ′) is a null set. Thus, we can apply Lemma 4 to get B ⊂ A such that ∪B is nonmeasurable. Now recall that SN, by formula (1), is the inverse image of an open set in L(H). The same is true for every translate of SN and for unions of such translates, in particular for every zαKSN and for ∪B. As the inverse image of an open set, ∪B must be measurable. This contradiction proves the theorem. (cid:3) 2. Continuity of group homomorphisms The contents of this section is valid for both treatments of the Haar measure. For the inner regular variant accepted by [13], it suffices to ignore the bracketed “locally” everywhere. Let G be a locally compact group. Call a set A ⊂ G extra-measurable if SA is measurable for every set S ⊂ G. Every open set is extra-measurable, while a one-point set is not, unless the group is discrete. As shown below (Theorem 6), existence of discontinuous measurable homomorphisms im- plies existence of [locally] null (definition below) extra-measurable sets. It is consistent with ZFC (Theorem 9) that a nonempty [locally] null set cannot be extra-measurable, so it is consistent that every measurable homomor- phism from a locally compact group to any topological group is continuous. It is an open question, whether this statement is true in ZFC without any additional axioms. Already in the basic case of the real line the answer CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS 7 is unknown, but for commutative groups one can make the question more precise (Proposition 7). Theorem 6. Let G be a locally compact group. If there exists a homomor- phism ϕ : G → H to a topological group H which is measurable but discon- tinuous, then there is a family A of nonempty [locally] null extra-measurable sets such that for every A ∈ A: A−1 = A; ∃B ∈ A : B2 ⊂ A; (2) ∀x ∈ G ∃C ∈ A : x−1Cx ⊂ A. Proof. Suppose that such ϕ exists. Let U be an open neighbourhood of identity in H and let A = ϕ−1(U). Then for any S ⊂ G we have SA = ϕ−1(ϕ(S)U). Indeed, the inclusion ϕ(SA) ⊂ ϕ(S)ϕ(A) = ϕ(S)U is obvious. For the opposite inclusion, take z ∈ ϕ−1(ϕ(S)U) and choose s ∈ S, a ∈ A such that ϕ(z) = ϕ(s)ϕ(a) = ϕ(sa). Let (sa)−1z = t, then t ∈ ker ϕ; since ϕ(at) = ϕ(a) ∈ U, we have at ∈ A and z = sat ∈ SA. Now SA is the inverse image of an open set ϕ(S)U, so it must be mea- surable. Thus, A is extra-measurable. Suppose that ϕ−1(U) is not [locally] null for every U. Take an open neighbourhood of identity V such that V −1V ⊂ U. Then B = ϕ−1(V ) is by assumption also non-[locally] null. It contains then a set C with 0 < µ(C) < ∞, so C −1C contains a neighbourhood of identity W ⊂ G [16, 20.17]. Then ϕ(W ) ⊂ ϕ(C −1C) ⊂ V −1V ⊂ U, so ϕ−1(U) ⊃ W . Since U was arbitrary, this implies that ϕ is continuous. Thus, in assumptions of the theorem there is U such that ϕ−1(U) is [locally] null. Let V be a base of neighbourhoods of identity in H, such that V ⊂ U and V = V −1 for every V ∈ V. Denote A = {ϕ−1(V ) : V ∈ V}, then this family has properties (2). (cid:3) The properties (2) guarantee that if we take A as a base of neighbour- hoods of identity in G, this turns G into a topological group [2, IV, §2]. However, it does not follow immediately that the identical map on G is measurable, i.e. in general we do not get a converse of this theorem. An equivalence holds in the commutative case: Proposition 7. Let G be a commutative locally compact group. The fol- lowing are equivalent: (i) There is a homomorphism ϕ : G → H to a topological group H which is measurable but discontinuous; (ii) There is a sequence of [locally] null extra-measurable sets An such that for every n: A−1 n = An; A2 n+1 ⊂ An. Proof. (i)⇒(ii): Proved in Theorem 6. (ii)⇒(i): Take the sets An as a base of neighborhoods of identity in G, then this turns G into a topological group which we can denote H. (Note that H is metrizable if ∩An = ∅.) The identical map ϕ : G → H is 8 YULIA KUZNETSOVA obviously discontinuous. At the same time, for every open set U ⊂ H we have U = ∪TnAn for some sets Tn; every TnAn and hence U = ϕ−1(U) are measurable, so ϕ is a measurable map and (i) holds. (cid:3) Existence of sets as in Proposition 7(ii) is an open question even on the real line. Known results on automatic continuity mostly concern Polish groups; here they do not give a ready answer, since the group H obtained in the proof may be not complete (i.e. not Polish). Let add(N ) be the minimal cardinality of a family J of null sets on the real line G such that ∪J is non null. This is called the additivity of the ideal N of Lebesgue null sets in G. It is known that additivity of the ideal of Haar null sets is the same for every non-discrete locally compact Polish group [13, 522Va]. It is consistent with ZFC that add(N ) < c, but it follows from Martin’s axiom (MA) that add(N ) = c (see [12]). This is, in fact, the assumption that we use in our proof. It is known that Martin’s axiom follows from the Continuum hypothesis, but is consistent also with its negation. For further discussion of Martin’s axiom, we refer to the Fremlin’s monograph [12]. A. Kharazishvili has indicated in private correspondence that the follow- ing statement holds for a commutative Polish group: Lemma 8 (MA). Let G be a locally compact Polish group, and let A ⊂ G be a nonempty set of measure zero. Then there is a set S ⊂ G such that SA is nonmeasurable. Proof. If G is countable, then by local compactness it has isolated points, and the measure of every point is positive. Then the set A in assumption cannot exist. Thus, G is uncountable without isolated points. Note that G is σ-compact [11, Theorems 3.3.1, 3.8.1, 3.8.C(b)]. We will construct S so that both SA and G \ SA intersect every perfect set of positive measure. Then SA must be nonmeasurable, since the inner measure of SA and G \ SA is zero. By translating A, and then S, if necessary, one can assume that e ∈ A. Note that A−1 has also measure zero—this follows, e.g., from [16, 20.2] or [13, 442K]. Since G is separable and uncountable, there is exactly continuum many closed sets in it. Let {Pξ : ξ < c} be an enumeration of all perfect sets of positive measure. By induction, we will choose sξ, dξ ∈ Pξ so that the condition dξ ∈ Pξ \ sηA holds for every ξ, η. Then S = {sξ : ξ < c} will be as needed, since sξ ∈ S ∩ Pξ ⊂ SA ∩ Pξ and dξ ∈ Pξ\SA, so both Pξ ∩ (SA) and Pξ ∩ (G \ SA) are nonempty. Suppose that for all η < ξ such sη, dη are chosen, or that ξ = 0 (the base of induction). Set Dξ = {dη : η < ξ} and note that Dξ < c. Since Pξ cannot be covered by a less that continuum translates of A−1 (here we use the Martin’s axiom), we can choose a point sξ ∈ Pξ \ DξA−1 6= ∅. This implies (sξA) ∩ Dξ = ∅. Next, set Sξ = {sη : η 6 ξ}. Then Sξ < c and similarly we can choose dξ ∈ Pξ \ SξA. By this choice, we have dξ /∈ sηA for all η 6 ξ, and for η > ξ we have dξ /∈ sηA by the choice of sη. This concludes the proof. (cid:3) CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS 9 Theorem 9 (MA). Let G be a locally compact group, and let A ⊂ G be a nonempty [locally] null set; then there is a set S ⊂ G such that SA is nonmeasurable. Proof. Let H be an open pro-Lie subgroup of G. Clearly, H can be chosen σ-compact (e.g., generated by any pre-compact neighborhood of identity). Translating A, if necessary, we can assume that e ∈ A. Then A1 = A ∩ H is nonempty and [locally] null with respect to the Haar measure of H, and due to σ-compactness it is just null in H. If we find a set S ⊂ H such that SA1 is nonmeasurable in H, then (SA) ∩ H = SA1 is nonmeasurable in G, and so SA is nonmeasurable too. We can assume therefore that G = H, that is: G is σ-compact and pro-Lie, and A is null. As in the proof of Theorem 5, either G is Polish (and we can apply Lemma 8), or we can find a Polish quotient G/K such that π(A) is null, where π : G → G/K is the quotient map. By Lemma 8, there is a set S1 ⊂ G/K such that S1π(A) is nonmeasurable. By Lemma 1, π−1(S1π(A)) = π−1(S1)A is also nonmeasurable. Thus, we can take S = π−1(S1), and the theorem is proved. (cid:3) This theorem together with Theorem 6 imply: Theorem 10 (MA). Every measurable homomorphism from a locally com- pact group to any topological group is continuous. In conclusion, let us review some close results. Say that a set S is small if the union of every family of translates of S of cardinality less than continuum is null. We use Martin’s axiom to guarantee that every null set is small. Without MA, this depends on the set S. Gruenhage [9] has proved that the ternary Cantor set is small, and Darji and Keleti — that every subset of R of packing dimension less than 1 is small. From the other side, Elekes and Tóth [10] and Abért [1] proved the following: it is consistent with ZFC that in every locally compact group there is a non-small compact set of measure zero. It is however unknown whether for a non-small set the statement of Theorem 9 is false. Finally, we say a few words on results in ZFC concerning nonmeasurable products of sets. One should better say “sums of sets” because there is a tradition to do everything in the commutative case. This restriction is reasonable since the principal difficulties lie already in the case of the real line. The advances most close to our topic are: for every null set S on the real line such that S +S has positive outer measure there is a set A ⊂ S such that A + A is nonmeasurable (Ciesielski, Fejzic and Freiling [8]). Cichoń, Morayne, Rałowski, Ryll-Nardzewski, and Żeberski [7] proved that there is a subset A of the Cantor set C such that A+C is nonmeasurable, and under additional axioms they prove the same statement for every closed null set P such that P + P has positive measure. There is also a series of results going back to Serpiński which find null sets A and B such that A + B is non- measurable (see, e.g., a monograph [19] and recent papers of Kharazishvili and Kirtadze [20], [21]), where the task is to make A and B “maximally negligible” (in different senses), and A + B “maximally nonmeasurable”. 10 YULIA KUZNETSOVA Author thanks S. Akbarov for drawing her attention to this problem. I thank also S. Akbarov, J. Cichoń, M. Morayne, R. Rałowski and J. Żeberski for useful discussions. References [1] M. Abért. Less than continuum many translates of a compact nullset may cover any infinite profinite group. J. Group Theory 11 (2008), 545–553. [2] N. Bourbaki. General topology, Chapters 5-10. Springer, 1989. [3] J. Brzdęk. Continuity of measurable homomorphisms. Bull. Austral. Math. Soc. 78 (2008), 171–176. [4] J. Brzuchowski, J. Cichoń, E. Grzegorek and C. Ryll-Nardzewski, On the existence of nonmeasurable unions, Bull. Polish Acad. Sci. Math. 27 no. 6 (1979), 447–448. [5] L. Bukovský, Any partition into Lebesgue measure zero sets produces a non- measurable set, Bull. Acad. Polon. Sci. 27 no. 6 (1979), 431–435. [6] J. P. R. Christensen, Borel structures in groups and semigroups, Math. Scand. 28 (1971), 124–128. [7] J. Cichoń, M. Morayne, R. Rałowski, C. Ryll-Nardzewski, S. Żeberski, On nonmea- surable unions. Topology and its Applications 154 (2007), 884–893. [8] K. Ciesielski, H. Fejzic, C. Freiling. Measure zero sets with non-measurable sum. Real Anal. Exch. 27 no. 2 (2001), 783–794. [9] U. B. Darji, T. Keleti. Covering R with Translates of a Compact Set. Proc. AMS, 131 No. 8 (2003), 2593–2596. [10] M. Elekes, Á. Tóth. Covering locally compact groups by less than 2ω translates of a compact nullset. Fund. Math. 193 no. 3 (2007), 243–257. [11] R. Engelking, General topology. Taylor & Francis, 1977. [12] D. H. Fremlin, Consequences of Martin’s axiom. Cambridge tracts in mathematics, no. 84. Cambridge University Press, 1984. [13] D. H. Fremlin. Set-theoretic measure theory: Vol. V. Torres Fremlin, 2008. [14] S. A. Gaal, Linear analysis and representation theory. Springer, 1973. [15] P. R. Halmos. Measure theory. Springer, 1974. [16] E. Hewitt, K. A. Ross, Abstract harmonic analysis I, II. Springer, 3rd printing, 1997. [17] K. H. Hofmann, S. A. Morris, The Structure of Compact Groups. De Gruyter, 2006. [18] K. H. Hofmann, S. A. Morris. The Lie theory of connected pro-Lie groups. EMS, 2007. [19] A. Kharazishvili. Nonmeasurable sets and functions. North Holland, 2004. [20] A. Kharazishvili, A. Kirtadze. On algebraic sums of measure zero sets in uncountable commutative groups. Proc. A. Razmadze Math. Inst. 135 (2004), 97–103. [21] A. Kharazishvili. The algebraic sum of two absoultely negligible sets can be an absolutely nonmeasurable set. Georgian Math. J. 12 no. 3 (2005), 455–460. [22] A. Kleppner. Measurable Homomorphisms of Locally Compact Groups, Proc. AMS 106, no. 2 (1989), 391–395. [23] B. J. Pettis, On continuity and openness of homomorphisms in topological groups, Ann. of Math. (2), 52 (1950), 293–308. [24] H. Reiter, J. D. Stegeman. Classical harmonic analysis and locally compact groups, Oxford: Clarendon Press, 2000. [25] C. Rosendal. Automatic continuity of group homomorphisms. Bull. Symb. Logic 15 no. 2 (2009), 184–214. [26] A. I. Shtern. Van der Waerden continuity theorem for semisimple Lie groups. Russ. J. Math. Phys. 13 No. 2 (2006), 210–223. [27] B. L. van der Waerden, Stetigkeitssätze für halbeinfache Liesche Gruppen, Math. Z. 36 no. 1 (1933), 780–786. CONTINUITY OF MEASURABLE GROUP REPRESENTATIONS 11 University of Luxembourg 6, rue Richard Coudenhove-Kalergi L-1359 Luxembourg Tel: (+352) 466644 6802
1902.02073
1
1902
2019-02-06T09:07:20
Hardy spaces of general Dirichlet series - a survey
[ "math.FA" ]
The main purpose of this article is to survey on some key elements of a recent $\mathcal{H}_p$-theory of general Dirichlet series $\sum a_n e^{-\lambda_{n}s}$, which was mainly inspired by the work of Bayart and Helson on ordinary Dirichlet series $\sum a_n n^{-s}$. In view of an ingenious identification of Bohr, the $\mathcal{H}_p$-theory of ordinary Dirichlet series can be seen as a sub-theory of Fourier analysis on the infinite dimensional torus $\mathbb{T}^\infty$. Extending these ideas, the $\mathcal{H}_p$-theory of $\lambda$-Dirichlet series is build as a sub-theory of Fourier analysis on what we call $\lambda$-Dirichlet groups. A number of problems is added.
math.FA
math
HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY ANDREAS DEFANT AND INGO SCHOOLMANN 9 1 0 2 b e F 6 ] Abstract. The main purpose of this article is to survey on some key elements of a recent Hp-theory of general Dirichlet seriesP ane−λns, which was mainly inspired by the work of Bayart and Helson on ordinary Dirichlet seriesP ann−s. In view of an ingenious identification of Bohr, the Hp-theory of ordinary Dirich- let series can be seen as a sub-theory of Fourier analysis on the infinite dimen- sional torus T∞. Extending these ideas, the Hp-theory of λ-Dirichlet series is build as a sub-theory of Fourier analysis on what we call λ-Dirichlet groups. A number of problems is added. . A F h t a m [ 1 v 3 7 0 2 0 . 2 0 9 1 : v i X r a 1. Introduction Within the last two decades the theory of ordinary Dirichlet seriesP ann−s saw a sort of renaissance. The study of these series in fact was one of the hot topics in mathematics at the beginning of the 20th. Among others, H. Bohr, Besicovitch, Bohnenblust, Hardy, Hille, Landau, Perron, M. Riesz, and Neder, were the leading mathematicians in this issue. The theory lived a sort of golden moment between the 1910s and the 1930s but after that it was somehow forgotten. Some 20 years ago the seminal article [33] of Hedenmalm, Lindqvist and Seip called again the attention from analysis to Dirichlet series. Since then a lot has been going on, and ordinary Dirichlet series have been studied with new techniques from functional and harmonic analysis. Bohr's main interest was to derive properties of Dirichlet series from the analyt- ical properties of the holomorphic functions defined by them. It is well known that Dirichlet series converge on half planes, and where they converge, they define a holomorphic function. Bohr considered three abscissas for a given Dirichlet series D =P ann−s: σc(D), σu(D), and σa(D) that define the maximal half planes on which the series respectively converges, converges uniformly, or converges abso- lutely. He also considered a fourth abscissa, σb(D), that gives the maximal half plane on which the series defines a bounded and holomorphic function. Then σc(D) ≤ σb(D) ≤ σu(D) ≤ σa(D) . 2010 Mathematics Subject Classification: Primary Key words and phrases: 1 2 DEFANT AND SCHOOLMANN In [7] Bohr was interested on describing the absolute convergence abscissa of an ordinary Dirichlet series in terms of analytic properties of its limit function. He proved (1) σb(D) = σu(D) , (see e.g. [26, Corollary 1.14]), and then considered the number S = sup{σa(D) − σu(D) : D =X ann−s} , that gives the maximal width of the band on which a Dirichlet series can con- verge uniformly but not absolutely. Bohr showed that S ≤ 1/2. The problem of whether or not this was the correct value remained open for some 15 years, until Bohnenblust and Hille in [5] indeed proved that (2) S = 1 2 (see also [26, Theorem 4.1]). These ideas are the seeds of a recent revival of interest in the research area opened up by these early contributions. A new field emerged intertwining the classical work in novel ways with modern functional analysis, infinite dimensional holomorphy, probability theory as well as analytic number theory. As a consequence, a number of challenging research problems crystallized and were solved over the last decades. We refer to the monographs [36], [41], and [26] were many of the key elements of this new developments for ordinary Dirichlet series are described in detail. of all ordinary Dirichlet series D =P ann−s which converge and define a bounded, A fundamental object in these investigations is given by the Banach space D∞ and then necessarily holomorphic, function on [Re > 0] (endowed with the supre- mum norm on [Re > 0]), and a celebrated result from [33] shows that D∞ in fact equals the Hardy space H∞(T∞) on the infinitely dimensional torus. Let us explain this in more detail. The infinitely dimensional torus T∞ is the infinite product of T = {w ∈ C : w = 1} which forms a natural compact abelian group on which the Haar measure is given by the normalized Lebesgue measure. The characters on T∞, so the elements in the dual group, consists of all monomials z 7→ zα, where α = (αk) ∈ Z(N) (all a finite sequence of integers), and H∞(T∞) denotes the closed subspace of all f ∈ L∞(T∞) such that the Fourier coefficient f (α) =ZT∞ f (w)w−αdw = 0 , whenever α < 0 (in the sense that some αk < 0). Then, based on Bohr's work, it is proven in [33] (see again [26, Corollary 5.3]) that there is a unique linear isometry (3) H∞(T∞) = D∞ , which preserves Fourier- and Dirichlet coefficients in the sense that f (α) = an whenever n = p (4) where α = (α1, . . . , αN , 0 . . .) ∈ N(N) and p = 2, 3, 5 . . . the sequence of primes. We refer to this result as the Bohr-Hedenmalm-Lindqvist-Seip theorem (see also [26, 0 α := pα1 1 . . . pαN N , HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 3 Corollary 5.3]). The crucial point of its proof is the fact that each natural number has a unique prime number decomposition, as well as Kronecker's theorem on Diophantine approximation (see e.g. [26, Proposition 3.4]): The continuous group homomorphism has dense range, which in particular implies that for each n and α with n = pα the following diagram commutes: β : R → T∞ , t 7→ (p−it k )∞ k=1 zα T e−it log n T∞ β R (5) The original proof of (3) goes a detour through infinite dimensional holomorphy using a result of Cole and Gamelin from [28]. Denote by H∞(Bc0) all holomorphic (Fr´echet differentiable) functions f on the open unit ball Bc0 of c0, which endowed with the sup norm forms a Banach space. Then Cole and Gamelin show that there is a unique isometric linear bijection (6) H∞(T∞) = H∞(Bc0) , f 7→ g which preserves Fourier and monomial coefficients in the sense that for every multiindex α (7) f (α) = ∂αg(0) α! (see [26, Theorem 5.1]). The equality of Banach spaces from (3) shows that a Dirichlet series D belongs to D∞ if and only if there is a function f ∈ H∞(T∞) such that the Dirichlet coefficients (an(D)) and the Fourier coefficients ( f (α)) coincide in the sense of (4), if and only if there is a function f ∈ H∞(Bc0) such that the Dirichlet coefficients (an(D)) and the monomial coefficients ( ∂αg(0) ) coincide in the sense of (7). This links the theory of ordinary Dirichlet which generate bounded, holomorphic functions on the positive half plane intimately with Fourier analysis on the group T∞ as well as infinite dimensional holomorphy on the open unit ball of c0. α! More generally, Bayart in [3] developed an Hp-theory of Dirichlet series. Re- call that the Hardy space Hp(T∞) , 1 ≤ p ≤ ∞, is the closed subspace of all f ∈ Lp(T∞) such that f (α) = 0 only if α < 0. Then the Banach spaces Hp of ordinary Dirichlet series by definition is the isometric image of Hp(T∞) under the identification from (4): and the Bohr-Hedenmalm-Lindqvist-Seip theorem from (6) reads Hp = Hp(T∞) , (8) Again the Banach spaces Hp can be reformulated in terms of holomorphic functions in infinitely many variables. Instead of looking at holomorphic functions on Bc0, D∞ = H∞ = H∞(T∞) . 4 DEFANT AND SCHOOLMANN we look at holomorphic functions on ℓ2 ∩ Bc0, understood as an open subset of ℓ2. Then Hp(ℓ2 ∩ Bc0) is the Banach space of all holomorphic functions g : ℓ2 ∩ Bc0 → C for which (9) 0<r<1(cid:18)ZTn g(rw1, . . . , rwn, 0, 0, . . .)pd(w1, . . . , wn)(cid:19) 1 kgkHp(ℓ2∩Bc0 ) = sup Then for 1 ≤ p < ∞ there is a unique isometric identity (10) sup n∈N p Hp(ℓ2 ∩ Bc0) = Hp(T∞) . < ∞ . identifying Fourier coefficients and monomial coefficients in the sense of (7) (a detailed proof is given in [26, Chapter 13]). Let us now face general Dirichlet series -- our main object of interest. Given a frequency λ := (λn) (i.e. a non-negative sequence of real numbers tending to +∞), a λ-Dirichlet series is a formal sum with complex Dirichlet coefficients (an) and a complex variable s. Denote by D(λ) the space of all formal λ-Dirichlet series. We start recalling some basic facts. The natural domains of convergence of Dirichlet series are half spaces (see [32, §II.2, Theorem 1]), and as in the ordinary case the 'abscissas' X ane−λns σc(D) = inf {σ ∈ R D converges on [Re > σ]} , σa(D) = inf {σ ∈ R D converges absolutely on [Re > σ]} , σu(D) = inf {σ ∈ R D converges uniformly on [Re > σ]} , rule the convergence theory of general Dirichlet series. Again general Dirichlet series D define holomorphic functions on [Re > σc(D)], which relies on the fact that they converge uniformly on all compact subsets of [Re > σc(D)] (see [32, §II.2, Theorem 2]). There are useful Bohr-Cahen formulas for the abscissas σc and σa, that are, given D =P ane−λns, σc(D) ≤ lim sup N log(cid:16)(cid:12)(cid:12)PN n=1 an(cid:12)(cid:12)(cid:17) λN and σa(D) ≤ lim sup N log(cid:16)PN n=1 an(cid:17) λN , where in each case equality holds if the left hand side is non negative. See [32, §II.6 and §II.7] for a proof. The formula for σu (and its proof) extends from the ordinary case in [26, §1.1, Proposition 1.6] canonically to arbitrary λ's: (11) σu(D) ≤ lim sup N where again equality holds if the left hand side is non negative. log(cid:16) supt∈R(cid:12)(cid:12)PN n=1 ane−λnit(cid:12)(cid:12)(cid:17) λN , HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 5 We point out that making the jump from the ordinary case λ = (log n) to arbi- trary frequencies reveals serious difficulties and that many of the above ideas fail for general Dirichlet series. Much of the ordinary theory relies on Bohr's theorem from (1), the fact that for each ordinary Dirichlet series the abscissa of uniform convergence and boundedness coincide. This phenomenon fails for general Dirich- let series. Further due to the prime number theorem each natural number n has its prime number decomposition n = pα and so the frequency (log n) can be written as a linear combination of (log pj) with natural coefficients. This intimately links the theory of ordinary Dirichlet series with the theory of holomorphic functions many complex variables. One of several consequences is that m-homogeneous on polydiscs, and in particular with the theory of polynomialsP cαzα in finitely Dirichlet seriesP ann−s, i.e. an 6= 0 only if n has m prim factors, are linked with m-homogeneous polynomials. This way powerful tools enter the game, as e.g. polynomial inequalities (like the Bohnenblust-Hille inequalities, hypercontractiv- ity of convolution with the Poisson kernel, etc.), m-linear forms, or polarization. So modern Fourier analysis and infinite dimensional holomorphy enrich the theory of ordinary Dirichlet series considerably. All this is carefully explained in [26]. Unfortunately facing general Dirichlet series many of these powerful bridges seem to collapse and new questions arise which make the theory of general Dirichlet quite challenging. Nevertheless inspired by the ordinary theory we are able to introduce a Fourier analysis setting for the study of general Dirichlet series by restricting ourserlves to λ-Dirichlet series with Dirichlet coefficients that actually are Fourier coefficients defined by functions on compact abelian groups G of a certain type; compact abelian groups allowing a continuous homomorphism β : R → G with dense range. 2. Bohr's theorem To get started we define the space D∞(λ) analogously to the space D∞ = D∞((log n)). Definition 2.1. Let λ be a frequency. Then D∞(λ) denotes the space of all λ- Dirichlet series D which converge on [Re > 0], and define (a then necessarily holomorphic) bounded function there. For instance, if λ = (n)∞ n=0, then looking at the transformation z = e−s we easily conclude that D∞((n)) is simply H∞(D), the space of all bounded and holomorphic functions on the open unit ball D. And if λ = (log n), then we are in the ordinary case. Recall that there is a unique coefficient preserving isometry identifying H∞(D) and H∞(T). Hence, in both cases, λ = (n) and λ = (log n), the space D∞(λ) can be described in terms of Fourier analysis, that is, it can be considered as a Hardy space, namely D∞((n)) = H∞(T) and D∞((log n)) = H∞(T∞). In view of these two examples the following question arises naturally: Given an arbitrary frequency λ, is it possible to describe D∞(λ) in terms of a sort of Hardy space on a compact abelian group? 6 DEFANT AND SCHOOLMANN The first step towards the solution of this problem, is to follow ideas of Bohr and Landau, who asked under which assumptions the abscissa σu(D) can be described in terms of analytic properties of the limit function of D. Define σb(D) = inf σ , where the infimum is taken over all σ ∈ R such that D converges and defines a bounded function on [Re > σ] (which then is automatically holomorphic). Addi- tionally define, provided σc(D) < ∞, σext b (D) = inf σ , the infimum now taken over all σ ∈ R such that the limit function of D allows a holomorphic and bounded extension to [Re > σ]. By definition σc(D) ≤ σb(D) ≤ σu(D) ≤ σa(D) and σext (D) ≤ σb(D). But in general all these abscissas differ. For instance an example of Bohr from [11] shows that σc(D) = σext (D) = σb(D) = −∞ and σu(D) = +∞ is possible. b b A very prominent research project at the beginning of the 20th century was to find conditions for frequencies λ under which (12) σext b (D) = σu(D) holds for all somewhere convergent λ-Dirichlet series D. In this context the space Dext ∞ (λ) of all somewhere convergent λ-Dirichlet series D =P ane−λns, which have a holo- morphic and bounded extension to [Re > 0], appears naturally. Endow Dext ∞ (λ) with the semi norm kDk∞ := sup[Re>0] f (s), where f is the (unique) extension of D. We will later in fact see that this always defines a norm -- but that, in general, neither D∞(λ) nor Dext ∞ (λ) form Banach spaces. Much of the abstract theory of ordinary Dirichlet series is based on a fundamen- tal theorem of Bohr from [8] which shows that every D ∈ Dext ∞ ((log n)) converges uniformly on all half spaces [Re > ε] for all ε > 0, and this then easily implies that Dext ∞ ((log n)) = D∞((log n)) (see also [26, Theorem 1.13]). For certain classes of frequencies λ, Bohr [7] and Landau [38] extended this result to λ-Dirichlet series. In [7] Bohr shows that (12) holds if λ satisfies the following condition (we call it Bohr's condition (BC)): (13) ∃ l = l(λ) > 0 ∀ δ > 0 ∃ C > 0 ∀n ∈ N : λn+1 − λn ≥ Ce−(l+δ)λn; roughly speaking this condition prevents the λn's from getting too close too fast. Motivated by Bohr's work we introduce the following definition. Definition 2.2. We say that a frequency λ satisfies Bohr's theorem, whenever every D ∈ Dext ∞ (λ) converges uniformly on all half spaces [Re > ε] , ε > 0, i.e. σext (D) = σu(D). b Observe that λ = (log n) satisfies (BC) with l = 1 and so Bohr's theorem holds. In [38] Landau gave a weaker sufficient condition than (BC) (we call it HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 7 Landau's condition (LC)), which extends the class of frequencies which satisfy Bohr's theorem: ∀ δ > 0 ∃ C > 0 ∀ n ∈ N : λn+1 − λn ≥ Ce−eδλn . (14) We like to mention that in [39, §1] Neder went a step further and considered λ's satisfying ∃ x > 0 ∃ C > 0 ∀ n ∈ N : λn+1 − λn ≥ Ce−exλn . Then Neder proved that this condition is not sufficient for Bohr's theorem by constructing, given some x > 0, a Dirichlet series D (belonging to some frequency λ) for which σc(D) = σa(D) = x and σext (D) ≤ 0 holds, which in particular shows that D∞(λ) ( Dext Analysing Bohr's original proof from [7], Bohr's theorem in the ordinary case ∞ (λ). b was improved in [4] by a quantitative version (see again [26, Theorem 1.13]). Theorem 2.3. There is a constant C > 0 such that for every D ∈ D∞((log n)) and all N ≥ 2 (15) sup [Re>0](cid:12)(cid:12)(cid:12)(cid:12) NXn=1 ann−s(cid:12)(cid:12)(cid:12)(cid:12) ≤ C log(N)kDk∞. Note that in the ordinary case λ = (log n), this inequality together with (11) implies (12). Like in the work of Bohr, Landau only proves the qualitative version of the fact that each frequency λ under his condition (LC) satisfies Bohr's theorem. To establish quantitative versions in the sense of (15) means to control the norm of the partial sum operator SN : Dext ∞ (λ) → D∞(λ), D 7→ an(D)e−λns . NXn=1 Using the summation method of typical means of order k > 0 invented by M. Riesz (see Proposition 2.6), the following estimate of kSNk, which holds under no further conditions on λ, is the main result of [43, Theorem 3.2]. Theorem 2.4. Let D = P ane−λns ∈ Dext N ∈ N: sup [Re>0](cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXn=1 ane−λns(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C ∞ (λ). Then for all 0 < k ≤ 1 and Γ(k + 1) k (cid:18) λN +1 λN +1 − λN(cid:19)k kDk∞, where C > 0 is a universal constant and Γ denotes the gamma function. As a consequence, assuming Bohr's condition (BC) for λ, the choice kN := 1 λN , N ≥ 2 (note that λ1 = 0 is possible) leads to (16) kSN : Dext ∞ (λ) → D∞(λ)k ≤ C1(λ)λN , which reproves (15) for λ = (log n). Assuming Landau's condition (LC), the estimate from Theorem 2.4 with kN := e−δλN , δ > 0, gives (17) kSN : Dext ∞ (λ) → D∞(λ)k ≤ C2(λ, δ)eδλN ; 8 DEFANT AND SCHOOLMANN the quantitative version of Bohr's theorem under (LC). For particular frequencies λ the bounds from (16) and (17) may be bad, which isn't too surprising since these results in fact hold for all frequencies satisfying the conditions of Bohr or Landau. For instance, consider the case λ = (n). Then the projection SN : H∞(T) → H∞(T) , SN (f ) = f (n)zn NXn=0 is nothing else than the convolution operator which assigns to every f its con- volution with the Dirichlet kernel DN , and this immediately gives that kSNk = kDNk1 ∼ log(N). To our knowledge the optimal upper and lower bounds for the norm of SN in the ordinary case λ = (log(n)) are still unknown. Problem 2.5. Determine optimal bounds for kSN : D∞((log n)) → D∞((log n))k. The proof of Theorem 2.4 relies on the following independently interesting result from [43, Proposition 3.4] which was inspired by the work of Hardy and M. Riesz from [32]. k > 0 the Dirichlet polynomials Proposition 2.6. Let D =P ane−λns ∈ Dext an(cid:18)1 − Rk x(D) = Xλn<x λn x(cid:19)k e−λns ∞ (λ) with extension f . Then for all converge uniformly to f on [Re > ε] for all ε > 0. Moreover, (18) x≥0 kRk sup x(D)k∞ ≤ e 2π Γ(k + 1) k kDk∞. In the language of [32] Proposition 2.6 states that, given any order k > 0, then ∞ (λ) is on every halfplane [Re > ε] the limit function of a Dirichlet series D ∈ Dext the uniform limit of its first typical means of order k. Moreover, Proposition 2.6 gives a direct link to the theory of almost periodic functions on R, and proves that Dext ∞ (λ) in fact is a normed space (see Corollary 2.8). Note that a priori, k · k∞ is only a semi norm, or equivalently, it is not clear whether Dext ∞ (λ) can be considered as a subspace of H∞[Re > 0], the Banach space of all holomorphic and bounded functions on [Re > 0]. Here it is important to distinguish Dirichlet series from their limit functions, and to prove that k · k∞ in fact is a norm on Dext ∞ (λ) requires to check that all Dirichlet coefficients of D vanish provided kfk∞ = 0. Recall that by definition a continuous function f : R → C is called (uniformly) almost periodic, whenever for every ε > 0 there is a number l > 0 such that for all intervals I ⊂ R with I = l there is a translation number τ ∈ I such that supx∈R f (x − τ ) − f (x) ≤ ε (see [6] for more information). Then by a result of Bohr a bounded and continuous function f is almost periodic if and only if it is the uniform limit of trigonometric polynomials on R, which are of the form HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 9 p(t) := PN n=1 ane−itxn for some x1, . . . , xN ∈ R (see e.g. 1.5.5]). In particular, the Dirichlet polynomials Rk considered as functions on vertical lines [Re = σ] are almost periodic. Corollary 2.7. If D ∈ Dext f (σ + it) : R → C is almost periodic and [41, §1.5.2.2, Theorem x(D) stated in Proposition 2.6 ∞ (λ) with extension f , then the function fσ(t) := an(D) = lim T →∞ f (σ + it)e(σ+it)λn 1 2T Z T −T for all σ > 0. In particular, supn∈N an ≤ kfk∞. This result is taken from [43, Corollary 3.8], and the next corollary is then an immediate consequence. Corollary 2.8. Dext spaces for any frequency λ. ∞ (λ), and consequently also its subspace D∞(λ), are normed Another particular consequence of Corollary 2.7 is as follows. Clearly, we may ∞ (λ) → D∞(λ)k ≤ N for all N, so equality deduce from this corollary that kSN : Dext (12) follows, i.e. λ satisfies Bohr's theorem, whenever L(λ) := lim sup N→∞ log(N) λN = 0. We like to mention that the number L(λ) has the following geometric meaning in terms of abscissas. In [9, §3, Hilfssatz 3, Hilfssatz 2] Bohr proved that (19) L(λ) = σc(cid:16)X e−λns(cid:17) = σa(cid:16)X e−λns(cid:17) = sup For instance we have that L((n)) = 0, and see as a consequence that for power se- ries we up to ε can't distinguished between uniform convergence and boundedness of the limit function. σa(D) − σc(D). D∈D(λ) Besides λ's with (BC) or (LC), there is another class of frequencies λ for which Bohr's theorem holds. In [12] Bohr proved that Q-linearly independent frequencies λ satisfy σext b (D) = σa(D) for all somewhere convergent λ-Dirichlet series D. The use of Kronecker's theorem, quency λ. Then (an) ∈ ℓ1 and k(an)k1 = kDk∞. Moreover, sequence (λn) is Q-linearly independent, combined with Proposition 2.6 in [43, Theorem 4.7] lead to an alternative proof of this fact. which states that the set (cid:8)(e−λnit)n t ∈ R(cid:9) is dense in T∞ whenever the real Theorem 2.9. Let D =P ane−λns ∈ Dext ane−λns(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞ (λ) = D∞(λ) = ℓ1 , X ane−λns 7→ (an) , Dext [Re>0](cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXn=1 ∞ (λ) with a Q-linearly independent fre- and λ satisfies Bohr's theorem. In particular, = kDk∞. sup sup N ∈N (20) (cid:26)λn + j sn (λn+1 − λn) n ∈ N, j = 0, . . . sn − 1(cid:27) 10 DEFANT AND SCHOOLMANN The following theorem summarizes some of the preceding results on frequencies λ satisfying Bohr's theorem. Theorem 2.10. A frequency λ satisfies Bohr's theorem, if one of following con- ditions holds: • L(λ) = 0 • λ is Q-linearly independent • λ satisfies (LC) (or the stronger condition (BC)) Moreover, in each of these cases we have Dext conditions is necessary for Bohr's theorem. ∞ (λ) = D∞(λ), but non of these Unfortunately D∞(λ) may fail to be a Banach space. The following result from [43, Theorem 5.2] is inspired by a construction of Neder in [39]. Proposition 2.11. Let λ be a frequency. Then there is a strictly increasing se- quence (sn) of natural numbers such that D∞(η), where η is the frequency obtained by ordering the set increasingly, is not complete and D∞(η) ( Dext fails for η. ∞ (η). In particular, Bohr's theorem The good news is that the completeness of D∞(λ) is guaranteed by several sufficient conditions on λ which in concrete cases have a good chance to be checked. The following collection of results was proved in [43, Theorem 5.1] (for the case of (BC) see also [19]). Theorem 2.12. Dext a Banach space and coincides with Dext following conditions: In particular, D∞(λ) is ∞ (λ) provided that λ satisfies one of the ∞ (λ) is complete, if L(λ) < ∞. • L(λ) = 0 • λ is Q-linearly independent • λ satisfies (LC) and L(λ) < ∞ (this includes (BC)). Moreover, none of these conditions is necessary for the completeness of D∞(λ). In view of the different nature of the stated conditions in Theorem 2.10 and The- orem 2.12, it seems that we are far away from a characterization of completeness of D∞(λ) (or Dext Problem 2.13. Characterize completeness of D∞(λ) and Dext In this context more interesting questions appear naturally. ∞ (λ)). ∞ (λ). Problem 2.14. Is there any relation between the λ's satisfying Bohr's theorem and the λ's for which D∞(λ) is complete? For instance the frequency λ := (√log n) fulfils (LC) (and so satisfies Bohr's theorem), but we don't know wether D∞((√log n)) is complete. Problem 2.15. Is D∞((√log n)) complete? HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 11 Recall Theorem 2.10 where we conditions on frequencies under which Bohr's theorem hold, that is σext b (D) = σb(D) for all λ-Dirichlet series D. Problem 2.16. Find a reasonable condition on the frequency λ which is weaker than (LC), but is sufficient for the equality σext (D) = σb(D) for all λ-Dirichlet series D. b By Theorem 2.12, D∞(λ) is complete, if σext (D) = σb(D) for all λ-Dirichlet series D and L(λ) < ∞. So any contribution to Problem 2.16 may give partial solutions to Problem 2.14. b We finish summarizing a few relations of the three conditions (BC), (LC) and L(λ) < ∞ (see [43, Remark 4.1]). Remark 2.17. • (BC) implies L(λ) < ∞ and (LC). • (LC) plus L(λ) < ∞ does not necessarily imply (BC). • L(λ) < ∞ does not necessarily imply (LC), and so neither (BC). • (LC) does not necessarily imply L(λ) < ∞, and so neither (BC). 3. Hardy spaces of general Dirichlet series Now we start our Hp-theory on general Dirichlet series. As already mentioned in the introduction we restrict ourselves to general Dirichlet series with Dirichlet coefficients which actually are Fourier coefficients of functions on certain compact abelian groups. This has several advantages. One is that the class of all general Dirichlet series simply is too large to obtain a good understanding. Assuming that the Dirchlet coefficients are Fourier coefficients gives more structure and allows to use tools from harmonic analysis like the Hausdorff-Young inequality or Plancherel's theorem (among others). A further advantage of our setting is that Bayart's Hp-theory of ordinary Dirichlet series embeds in a natural way. Whereas the Hp-theory of ordinary Dirichlet series is basically Fourier analysis on the infinite dimensional torus T∞, this group fails to be the right model for general Dirichlet series. In fact, the Bohr compactification R of R and products ofcQd (the dual group of the rationals endowed with the discrete topology) turn out to be suitable substitutes. Finally, fixing some λ, regarding the different realisations of λ-Dirichlet series of this type, another feature of our approach is that the Hp- theory of general Dirichlet series we intend to present will be independent of the chosen suitable group for λ. We are interested in the subclass of all compact abelian groups G which allow a continuous homomorphism β : R → G with dense range. We call such pairs (G, β) Dirichlet groups. Hence, given such a pair, the characters x = e−ix· ∈ bβ(bG) ⊂ bR are precisely those for which there is a unique character hx ∈ bG with e−ix· = hx◦β; to understand this recall that R = R, x → e−ix· is a group isomorphism. particular, we have In (21) bG = {hx x ∈ bβ(bG)}. 12 DEFANT AND SCHOOLMANN Then the following notion (first given in [24]) turns out to be fundamental for our purposes. Definition 3.1. Let λ be a frequency and (G, β) a Dirichlet group. Then G is called a λ-Dirichlet group whenever λ ⊂ bβ(bG), which means that the following diagram commutes for every n ∈ N: hλn T e−iλn· β G R Let us give examples. Denoting by p := (pn) the sequence of prime numbers, the compact group T∞ together with the mapping t 7→ p βT∞ : R → T∞, −it forms a (log n)-Dirichlet group (see again Kronecker's theorem). This example keeps us in track to recover results on ordinary Dirichlet series. The 'mother' of all possible examples is as follows: Given a subgroup U of R, the topological group \(U, d) together with the mapping β\(U,d) : R → \(U, d), t 7→(cid:2)u 7→ e−itu(cid:3) forms a Dirichlet group. So in particular, for U = Z and identifying T = bZ, we obtain the (n)-Dirichlet group (T, βT), where βT : R → T, t 7→ e−it. The compact abelian group R := \(R, d) is the so-called Bohr compactification of R which forms a λ-Dirichlet group for all frequencies λ with the embedding βR : (R, · ) ֒→ R, x 7→(cid:2)t 7→ e−ixt(cid:3) . Besides R, countable products ofcQd := \(Q, d), where d denotes the discrete topol- ogy, also form 'universal' Dirichlet groups. To explain this , let B := (b1, b2, . . .) be a Q-linearly independent sequence of real numbers of length N ∈ N ∪ {∞}. Then (22) is an injective homomorphism, and hence its dual map TB : Qd, NMn=1 Q ֒→ R, α 7→X αjbj \∞Mn=1 t 7→(cid:2)(qj)j 7→ e−it P qjbj(cid:3) n=1 Qd, the pair(cid:16)QN n=1cQd = \L∞ n=1cQd,cTB(cid:17) is a λ-Dirichet group. cTB : R → has dense range. SinceQN suitable B such that(cid:16)QN group, which is 'universal' in the following sense: For every frequency λ there is a n=1cQd,cTB(cid:17) is a Dirichlet Together with the norm kDkp := kfkp the space Hp(λ) clearly forms a Banach space, and then by definition the Bohr map P ane−λns for which there is some f ∈ H λ p (G) ֒→ D(λ), f ∼Xγ∈ bG Hp(λ) = H λ gives an isometric onto isomorphism B : H λ p (G) . p (G) such that an = bf (hλn) for all n. f (γ)γ 7→ Xn∈N bf (hλn)e−λns HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 13 To see this, rational sequence (rn let us recall that to every λ there is a Q-linearly independent sequence B := (bn) of real numbers, called basis for λ (which can always be k bn for some (unique) finite k )n,k is said to be a Bohr matrix of λ chosen as a subsequence of λ), such that λn = P rn with respect to the basis B and we write λ = (R, B). Hence(cid:16)QN λ-Dirichlet group for every λ with decomposition λ = (R, B). k ). In this case, R := (rn n=1cQd,cTB(cid:17) is a Given a λ-Dirichlet group (G, β) and 1 ≤ p ≤ ∞, we define the Banach space p (G) :=nf ∈ Lp(G) f : bG → C is supported by all hλn, n ∈ No , and use it to define the following natural scale of Hardy spaces of general Dirichlet series. Definition 3.2. The Hardy space Hp(λ) of λ-Dirichlet series is the space of all H λ The following fact, proved in [24, Theorem 3.19], is fundamental. Theorem 3.3. Hp(λ) is independent of the chosen λ-Dirichlet group G. So the above definition of Hp(λ) actually coincides with Bayart's definition of Hp from [3] in the ordinary case (see also [26, §11]). Moreover, recall that in this case the groups T∞ and R are suitable (log n)-Dirichlet groups, which immediately leads to the following consequence extending (3). Corollary 3.4. For all 1 ≤ p ≤ ∞ we have H log(n) p (R) = Hp(log(n)) = Hp(T∞), kfkp = kDkp = kgkp, polynomials we define Let us mention that there also is an internal description of Hp(λ) through λ-Dirichlet polynomials without considering λ-Dirichlet groups. We denote by n=1 ane−λns. For such where bf (hlog n) = an(D) =bg(α), if n = pα. P ol(λ) the space of all λ-Dirichlet polynomials D(s) = PN ane−λnit(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −T(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt! 1 2T Z T NXn=1 kDkp := lim Then this limit exists and gives a norm on P ol(λ) (see e.g. [41, Theorem 1.5.6] or [26, Theorem 11.9]). The following description was noted in [24, Theorem 3.25]. T →∞ 1 . p p 14 DEFANT AND SCHOOLMANN Theorem 3.5. Let 1 ≤ p < ∞ and λ be a frequency. Then the space Hp(λ) is the completion of (P ol(λ),k · kp). ordinary case λ = (log n). Then log n =P αj log pj whenever n = p 3.1. Frequencies of integer type. Recall that given a frequency λ there is a basis B for λ and a Bohr matrix R such that λ = (R, B). Let us look at the α, and hence a basis is given by B = (log pj) and every multi index α > 0 appears as a row in the corresponding Bohr matrix R. Moreover recall that T∞ is a (log n)-Dirichlet group with β : R → T∞, β(t) := p−it. More generally, we call of frequency λ of integer (natural) type, if there is a basis B = (bn) such that the Bohr matrix R associated to B and λ only has integer In this case T∞ is a λ-Dirichlet group with β(t) := e−itB := (natural) entries. (e−itb1, e−itb2, . . .), t ∈ R, and the rows of the corresponding Bohr matrix are multi indices α (with integer entries). Actually one can show that a frequency λ is of integer type if and only if there is a homomorphism β : R → T∞ such that (T∞, β) is a λ-Dirichlet group (see [24, Remark 3.32]). To see an example, consider the set M := {n +√2m n, m ∈ Z} which is dense in R. Then any frequency λ ⊂ M is of integer type with basis B = (1,√2). See Let B be a basis of length N ∈ N ∪ {∞} for some frequency λ of integer type with Bohr matrix R, and let us write α ∈ R if the multi index α appears as a row in R. Then we define H R p (TN ) to be the space of all g ∈ Lp(TN ) for which [24, Example 3.36] for a frequency which is not of integer type. bg(α) 6= 0 implies α ∈ R. By Theorem 3.3, the Hp(λ)'s can be identified with H R p (TN ) in the following sense (see Theorem [24, Theorem 3.30]). Theorem 3.6. Let 1 ≤ p ≤ ∞ and λ = (R, B) a frequency of integer type with basis of length N ∈ N ∪ {∞}. Then there is a unique onto isometry such thatbg(α) = an(D) for all multi indices α ∈ R and λn =P αjbj. An immediate consequence is that the ordinary Hp's in the following sense are the largest spaces for λ's of natural type. Corollary 3.7. Let 1 ≤ p ≤ ∞ and λ = (R, B) a frequency of natural type. Then there is a unique into isometry ψ : Hp(λ) → H R p (TN ), D 7→ g ψ : Hp(λ) ֒→ Hp, X ane−λns →X bnn−s such that an = bpα for all α ∈ R, where λn =P αjbj. 4. Some structure theory We summarize further properties of Hp(λ) which extend important key stones from Bayart's theory of ordinary Dirichlet series to our new theory of general Dirichlet series. HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 15 4.1. Coincidence of H∞H∞H∞'s. Note first that there are now two 'H∞-spaces of λ- Dirichlet series' around, namely D∞(λ) and H∞(λ). Recall that by Proposition 2.11 there are frequencies λ such that D∞(λ) is not complete, hence in these cases D∞(λ) 6= H∞(λ). The following result is given in [24, Theorem 4.10, 4.12], and an far reaching extension of the Bohr-Hedenmalm-Lindqvist-Seip theorem from (8). Theorem 4.1. For every frequency λ there is a coefficient preserving injective contraction and if Dext embedding is even an isometric equality ∞ (λ) = D∞(λ) and L(λ) < ∞ (see again Theorem 2.10), then the Dext ∞ (λ) ⊂ H∞(λ) , D∞(λ) = H∞(λ) . We mention an interesting by-product of Theorem 4.1 which is an immediate consequence of the definition of H2(λ) and Parseval's equality. Corollary 4.2. For each D ∈ Dext ∞ (λ) we have (an(D)) ∈ ℓ2 with k(an(D))k2 ≤ kDk∞, i.e. the embedding Dext ∞ (λ) ⊂ H2(λ) , is a well-defined contraction. 4.2. Schauder bases. Given a frequency λ and 1 < p < ∞, the following ques- tion is fundamental: Do the e−λns form a Schauder basis of Hp(λ)? In the ordinary case λ = (log n) the answer is affirmative as discovered in [1], and we will see that the same result holds true whenever λ is an arbitrary frequency. In (21) we note that, given a Dirichlet group (G, β), we have This in particular shows that the dual group bG inherits the order of R, and hence we deduce from [42, Theorem 8.7.2.], that the 'Riesz projection' bG = {hx x ∈ bβ(bG)} . Φ(cid:16)X akhxk(cid:17) := Xxk≥0 akhxk is bounded on the subspace P ol(G) of all polynomials in Hp(G), 1 < p < ∞. Then standard arguments shows the following important theorem from [24, Theorem 4.16]. Theorem 4.3. Let 1 < p < ∞ and λ be a frequency. Then the monomials e−λns form a Schauder basis for Hp(λ). An equivalent formulation of Theorem 4.3 is that for 1 < p < ∞ all projections Sp N : Hp(λ) → Hp(λ), X ane−λns 7→ NXn=1 ane−λns are uniformly bounded. But for the border cases p = 1 and p = ∞ this in general is false (e.g., for the frequencies λ = (log n) or λ = (n)). Upper bounds for the 16 DEFANT AND SCHOOLMANN growth of the partial sum operators in H∞(λ) were given in (16) and (17). The following result handles the case p = 1, and its proof in a sense reduces to the case p = ∞ (see [24, Proposition 4.17] and [26, 12.5] in the ordinary case). Proposition 4.4. Let λ be a frequency. Assuming (BC) for λ there is a constant C = C(λ) such that for all N ≥ 2 kS1 N : H1(λ) → H1(λ)k ≤ CλN , and, assuming (LC) and L(λ) < ∞, for every δ there is a constant D = D(δ, λ) such that for all N kS1 N : H1(λ) → H1(λ)k ≤ DeδλN . 4.3. Brothers Riesz theorem. Recall the classical brothers Riesz theorem with states that the Hardy space H1((n)) = H1(T) coincides with the space M(T) of all bounded, regular and analytic Borel measures on T. The corresponding result in ordinary case reads H1((log n)) = H1(T∞) = M(T∞) (due to Helson and Lowdenslager from [37]). See also [26, Theorem 13.5] for a proof within the setting of ordinary Dirichlet series. The brothers Riesz theorem extends to the case of general Dirichlet series. Theorem 4.5. Let λ be any frequency and let G a λ-Dirichlet group. Then the map is an onto isometry. In particular H1(λ) = Mλ(G). H λ 1 (G) → Mλ(G), f 7→ f dm Actually it was discovered in [24, Theorem 4.25] that this is a fairly simple consequence of a more general result due to [29, Theorem 4]; see [24, §4.7] for a discussion on this. 4.4. Montel theorem. In [3, Lemma 18] Bayart proved that for every bounded sequence (DN ) ⊂ D∞((log n)) there is a subsequence (DNk) and D ∈ D∞((log n)) such that (DNk) converges to D on [Re > ε] for all ε > 0. This fact, sometimes called 'Montel's theorem', extends to the following classes of λ's. See also [41], and [26, Theorem 3.11], where Bayart's Montel theorem is deduced from a Montel type theorem for H∞(Bc0) combined with (6) and (8). Theorem 4.6. Let λ satisfy L(λ) = 0, or L(λ) < ∞ and (LC), or let λ be Q-linearly independent. Then for every 1 ≤ p ≤ ∞ and every bounded sequence (DN ) ⊂ Hp(λ) there is a subsequence (DNk) and D ∈ Hp(λ) such that for all ε > 0 the translations about ε DNk ε =X an(DNk)e−ελne−λns converge to D in Hp(λ) as k tends to ∞. HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 17 The proof in a sense reduces the Hp(λ)-case to the H∞(λ)-case. This reduction modifies an idea from [27] showing that, under the assumption of the completeness of D∞(λ), the map Φ : Hp(λ) ֒→ D∞(λ,Hp(λ)), X ane−λns 7→X(cid:0)ane−λns(cid:1) e−λnz is an into isometry; here D∞(λ,Hp(λ)) stands for the Banach space of all λ- bounded on [Re > 0]. Dirichlet seriesP Ane−λns with coefficients An ∈ Hp(λ), which converge and are 4.5. Hilbert's criterion. As already mentioned the 'ordinary' D∞ isometrically equals H∞(Bc0), identifying Dirichlet and monomial coefficients. A crucial ar- gument in the proof of this result (see [26, §2.3]) is that a continuous function f : Bc0 → C is in H∞(Bc0) if and only if all restriction maps fN : DN → C are in H∞(DN ) and supN kfNk∞ < ∞. Formulated for ordinary Dirichlet series this shows that D =P ann−s ∈ D∞ if and only if all of its so-called N-th abschnitte DN =P ann−s , where the sum is only taken over those n which have only the first N primes as divisors, belong to D∞ with uniformly bounded norms. In more vague terms, D ∈ D∞ if and only if all its finite dimensional blocks are in D∞ with uniformly bounded norms (see also [26, Theorem 3.11]). This phenomenon is also true in the general case. Given some decomposition λ = (R, B) the N-th abschnitt DN of a λ-Dirichlet series D is the sumP an(D)e−λns, where an(D) 6= 0 implies that λn depends only on the first N basis elements b1, . . . , bN . As proven in [24, Theorem 4.22], general Dirichlet series D ∈ Hp(λ) under appropriate assumptions on the frequency are again 'determined by their finite dimensional parts'. Theorem 4.7. Let 1 ≤ p ≤ ∞ and let λ = (R, B) be a frequency satisfying one of the conditions of Theorem 4.6. Let D be a formal λ-Dirichlet series. Then D ∈ Hp(λ) if and only if the N-th abschnitt DN ∈ Hp(λ) for all N ∈ N and supN kDNkp < ∞. Moreover, in this case kDkp = supN ∈N kDNkp. By Theorem 4.1 we obtain the following particular case: If λ satisfies σext b = σb and L(λ) < ∞, then D ∈ D∞(λ) if and only if DN ∈ D∞(λ) for all N ∈ N and supN kDNk∞ < ∞. 4.6. Helson's theorem. A celebrated result of Helson [35] on general Dirichlet converges on the open right half plane [Re > 0], or equivalently, for all u > 0 seriesP ane−λns states that if λ satisfies (BC) and (an) is 2-summable, then for almost all homomorphism ω : (R, +) → T the Dirichlet series P anω(λn)e−λns the seriesP ane−λnuhλn converges almost everywhere on R (note that if we for all rational u's collect zero sets in R, then we get the aforementioned result). See also [33, Theorem 4.4] for the ordinary case. 18 DEFANT AND SCHOOLMANN Using our teminology, let us rephrase this result for square summable functions on Dirichlet groups: Given a frequency λ with (BC), a λ-Dirichlet group (G, β), and some f ∈ H λ 2 (G), then for every u > 0 the series f (hλn)e−λnuhλn Xn converges almost everywhere in G. This result extends to functions in H λ p (G). To see this recall that for u > 0 the Poisson kernel Pu(t) := 1 π u u2 + t2 : R → R is integrable with kPuk1 = 1. Hence, by the Riesz representation theorem, for every λ-Dirichlet group (G, β) the functional Tu : C(G) → C, g 7→ZR (g ◦ β)(t)Pu(t)dt defines a measure on G which we call the 'Poisson measure' on G denoted by pu. The following theorem from [25] is a far reaching extension of Helson's theo- rem. It in fact shows that, given an appropriate frequency λ, Helson's theorem extends to functions f from our scale H λ p (G), 1 ≤ p < ∞, modelled on λ-Dirichlet groups G. Moreover, it describes the G-a.e. pointwise limit of the Fourier series P∞ n=1 bf (hλn)e−λnuhλn in terms of convolution of f with the Poisson measure pu on G, and -- most important -- it adds the relevant maximal inequality. Theorem 4.8. Let (G, β) be a λ-Dirichlet group for a frequency with (BC), and f ∈ H λ p (G) with 1 ≤ p < ∞. Then for every u > 0 almost everywhere on G. Moreover, there is a constant C = C(u, λ) such that for all f ∈ H λ p (G) (23) f ∗ pu = ∞Xn=1 bf (hλn)e−λnuhλn NXn=1 bf (hλn)e−λnuhλn(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Ckfkp. (cid:13)(cid:13)(cid:13)(cid:13) sup N (cid:12)(cid:12)(cid:12) 4.7. Vertical Limits. Let us reformulate Theorem 4.8 in terms of general Dirich- let series. Given a λ-Dirichlet series D(s) = P ane−λns and a λ-Dirichlet group G, we for ω ∈ G call the Dirichlet series Dω(s) =X anhλn(ω)e−λns a vertical limit of D. The following characterization of vertical limits justifies their name ([24, Proposition 4.6]). Remark 4.9. Let (G, β) be a λ-Dirichlet group, and D =P ane−λns a λ-Dirichlet series with σa(D) ≤ 0. Then HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 19 (1) For every ω ∈ G there is a sequence (τk) ⊂ R such for all ε > 0 the 'vertical translations' converge to Dω uniformly on [Re > ε]. Diτk =X ane−iλnτk e−λns (2) Assume conversely that the (Diτk) for some (τk) ⊂ R converge uniformly on [Re > ε] to a holomorphic function f for every ε > 0. Then there is ω ∈ G such that f (s) = Dω(s) for all s ∈ [Re > 0] . Note that the notion of vertical limits allows to rephrase Helson's theorem from the preceding section as follows: Given the λ-Dirichlet group G = R and D ∈ H2(λ), then R-almost all vertical limits Dω converge on [Re > 0]. The following result is then a straight forward reformulation of Theorem 4.8. Theorem 4.10. Let (G, β) be a λ-Dirichlet group for a frequency with (BC), and D ∈ Hp(λ) with 1 ≤ p < ∞. Then almost all vertical limits converge on [Re > 0]. Moreover, for all u > 0 there is a constant C = C(u, λ) such that for all D ∈ Hp(λ) Dω(s) =X an(D)hλn(ω)e−λns, ω ∈ G, N (cid:12)(cid:12)(cid:12) anhλn(ω)e−uλn(cid:12)(cid:12)(cid:12) dm(ω)(cid:19) 1 NXn=1 p p (cid:18)ZG sup ≤ CkDkp. Remark 4.11. Note that in the maximal inequality (24) the constant C = C(u, λ) is independent of p. Hence, if p → ∞, then we recover that frequencies with (BC) satisfy Bohr's theorem, i.e. for every somewhere convergent λ-Dirichlet series D we have σext (D) = σu(D). b Again convolution allows to relate the values of the vertical limits Dω of a Dirichlet series D ∈ Hp(λ) and its associated function f ∈ H λ p (G). The vertical limits Dω in some sense recover the associated function f on vertical lines. Note that, given f ∈ Lp(G), almost all ω ∈ G define a measurable function fω(t) := f (ωβ(t)), which itself is defined for almost all t ∈ R. Hence, the convolution (24) fω ∗ Pu(t) :=ZR fω(t − y)Pu(y)dy is defined almost everywhere on R. Corollary 4.12. Let 1 ≤ p < ∞, λ satisfy (BC), and (G, β) be a λ-Dirichlet group. If f ∈ H λ p (G) and D ∈ Hp(λ) are associated to each other, then for all u > 0 and almost all t ∈ R Dω(u + it) = (fω ∗ Pu)(t). Let us consider the particular case of ordinary Dirichlet series. We denote by Ξ the set of all completely multiplicative characters χ : N → T (that is χ(nm) = 20 DEFANT AND SCHOOLMANN χ(n)χ(m) for all m,n) which with the pointwise multiplication forms an abelian group. Looking at the group isomorphism (25) ι : Ξ → T∞, χ 7→ (χ(pn))n, where p = (pn) again denotes the sequence of prime numbers, we see that Ξ also forms a compact abelian group, and its Haar measure dχ is the push forward measure of dz through ι−1. Then applying Theorem 4.8 to the (log n)-Dirichlet group Ξ we obtain the following consequence. Corollary 4.13. Let 1 ≤ p < ∞. Then for all u > 0 there is a constant C = C(u) such that for all Dirichlet seriesP ann−s ∈ Hp(λ) anχ(n)n−u(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dχ! 1 N (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXn=1 ZΞ sup p [Re > 0]. In particular, almost all vertical limits Dχ(s) =P anχ(n)n−s, χ ∈ Ξ, converge on mann's conjecture holds for almost all randomized ζ-functionsPn χ(n)n−s, in the sense that almost all of them converge pointwise on [Re > 1/2] without having any zero in this half plane (see also [33]). Remark 4.14. Helson in [36] deduces from this result the curious fact that Rie- p ≤ CkDkp . Bayart in [3] deduces the 'almost everywhere part' of Corollary 4.13 for the case p = 2 from the Menchoff-Rademacher theorem on almost everywhere convergence of orthonormal series, and then he extends it to the range 1 ≤ p < ∞ using the 'hypercontractivity of the Poisson kernel'. Let us explain what is meant by this. Recall the definition of the Poisson kernel on T pr(w) := rnwn, 0 < r < 1, ∞Xn=−∞ and consider for all 1 ≤ p < q < ∞ the following convolution operator Pr : Hp(T) → Hq(T), f 7→ f ∗ pr(ω) = ∞Xn=0 bf (n)(rw)n. An easy computation shows that kPrk ≤ 1 1−r (independent of p, q), but this bound improves considerably depending on the relation of p, q, r. Weissler in [45] shows that kPrk = 1 if and only if r2 ≤ p q ; a result which is refered to as the 'hy- percontractivity of the Poisson kernel'. Then using this result 'in each variable separately' shows that for all u > 0 and 1 ≤ p, q < ∞ the translation operator Pu : Hp 7→ Hq, D 7→ Du =X an(D)n−un−s is well-defined and bounded (see [26, Theorem 12.9] for the details on all that). This explains, why in the ordinary case 'the almost everywhere part' of Corollary 4.13 follows in the range 1 ≤ p < ∞, once this result is proven for p = 2. HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 21 It seems that in the case of general Dirichlet series no 'hypercontractivity results' are available, and so our proof of Theorem 4.10 in [25] follows an entirely different approach. Nevertheless the following problem seems interesting in itself. Problem 4.15. Given D = P aneλns ∈ Hp(λ), 1 ≤ p < ∞. Under which as- sumptions on λ is it true that Du =P aneλnueλns ∈ Hq(λ) for all 1 ≤ q < ∞ and u > 0? From [26, §12] we know that for 1 ≤ p < ∞ (26) sup D∈Hp σc(D) = sup D∈Hp σa(D) = 1 2 The second equality was implicitly proved in [3], and in view of the coefficient preserving identity Hp = Hp(T∞), it implies that the infimum over all u > 0 such that for each f ∈ Hp(T∞) the series Xα pu(cid:17)α f (α)(cid:16) z converges (here we of course talk about absolutele convergence) equals 1/2 (com- pare also with (10)). The next result shows that there is no such 'wall' if we order the sum 'blockwise according to the order of N'. Corollary 4.16. Let f ∈ Hp(T∞), 1 ≤ p < ∞. Then for all u > 0 lim N Xpα≤N N (cid:12)(cid:12)(cid:12) Xpα≤N f (α)(cid:16) z pu(cid:17)α pu(cid:17)α(cid:12)(cid:12)(cid:12) cα(cid:16) w p a.e. on T∞ , dω(cid:19)1/p ≤ C(u)kfkp , and moreover (cid:18)ZT∞ sup where C = C(u) only depends on u. What about the case ω = 1 (the neutral element of G) in Theorem 4.10, i.e. (pointwise) convergence of Dirichlet series from Hp(λ)? By (19) we always have σc(D) ≤ σa(D) ≤ L(λ) for all D ∈ Hp(λ) (since in this case the sequence (an(D)) of Dirichlet coefficients is always bounded). In the particular case λ = (log n), we know from (26) that sup D∈Hp σc(D) = sup D∈Hp σa(D) = L((log n)) 2 . Problem 4.17. Given a frequency λ, give reasonable upper and lower estimates for supD∈Hp(λ) σc(D) and supD∈Hp(λ) σc(D) , for example in terms of L(λ). almost everywhere on G, and f = f (hλn)hλn ∞Xn=1 f (hλn)hλn(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Ckfkp , (cid:13)(cid:13)(cid:13)(cid:13) sup N (cid:12)(cid:12)(cid:12) NXn=1 22 DEFANT AND SCHOOLMANN 4.8. Carleson's theorem. What happens, if we consider the case u = 0 in the preceding two sections, in particular in Theorem 4.8? It turns out that this question is intimately related with one of the most famous theorems in Fourier analysis, namely Carleson's theorem: The Fourier series of every f ∈ L2(T) converges almost everywhere. More generally, the Carleson-Hunt all z ∈ T and theorem states that if f ∈ Lp(T), 1 < p < ∞, then f (z) =Pn bf (n)zn for almost (cid:13)(cid:13)(cid:13) sup N (cid:12)(cid:12) NXn=−N f (n)zn(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)p ≤ Ckfkp , Recall that a frequency is of integer type if it has a basis B for which the Bohr matrix R only contains integers. For λ's of this type the following theorem extends the Carleson-Hunt theorem to our setting of λ-Dirichlet groups, and for such λ's it is a strong improvement of Theorem 4.3. Theorem 4.18. Let λ be a frequency of integer type, (G, β) a λ-Dirichlet group, and 1 < p < ∞. Then for every f ∈ H λ p (G) where C = C(p) is a constant which only depends on p. The case p = 2 is due to [34, Theorem 1.4], whereas the proof of the more general case given here follows closely the ideas of this article combining them with the Carleson-Hunt theorem and a magic trick of Fefferman from [30]. This is carried out in [25]. It is somewhat surprising that Theorem 4.18 doesn't require (BC) for λ. Problem 4.19. Does Theorem 4.8 hold for a wider class of frequencies than the class of λ's satisfying (BC)? Is it even true without any condition on λ? Theorem 4.18 easily translates into a sort of Carleson-Hunt theorem for func- tions on the polytorus T∞. Obviously, given some f ∈ Hp(T∞) with 1 < p < ∞, there in general is no z ∈ T∞ such that f (z) =Pα Corollary 4.20. Let f ∈ Hp(T∞) and 1 < p < ∞. Then for almost all z ∈ T∞ f (α)zα. and moreover, (cid:13)(cid:13)(cid:13)(cid:13) sup N (cid:12)(cid:12)(cid:12) Xpα≤N f (z) = lim f (α)zα , N Xpα≤N p(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(p)kfkp . f (α)zα(cid:12)(cid:12)(cid:12) HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 23 5. Vector-valued aspects of general Dirichlet series We now briefly discuss some results on vector-valued general Dirichlet series P ane−λns, where (an) ⊂ X for some Banach space X. All results are related to the above topics on general Dirichlet series with scalar coefficients, and most of them are taken from [16]. In the ordinary case λ = (log n) the study of functional analytic aspects of vector-valued Dirichlet series accumulated quite some amount of research, see e.g. [14], [15], [17], [18], [20], [21], [22], or [23]. As an example we recall the following strong extension of (2) from [20]: The maximal width of the strip of uniform, non absolute convergence of X-valued ordinary Dirichlet series is given by the formula S(X) = 1 − 1 cot(X) , where cot(X) denotes the infimum over all 2 ≤ q ≤ ∞ for which X has cotype q (see also [26, Theorem 26.4]). Given a Banach space X and a frequency λ, we denote by D(λ, X) the space of all λ-Dirichlet series with coefficients in X, and by D∞(λ, X) the space of all D ∈ D(λ, X) which on [Re > 0] define bounded (and then necessarily holomorphic) functions. Moreover, the by now obvious definition of the a priori larger space Dext ∞ (λ, X) is then clear. The Hahn-Banach theorem transports many results from scalar-valued Dirichlet series to vector-valued ones. For instance this way we may easily deduce from Corollary 2.7 that (D∞(λ, X),k · k∞) is a normed space with sup n∈N kankX ≤ kDk∞ for all D ∈ D∞(λ, X). In [16] it is shown that D∞(λ) is complete if and only if D∞(λ, X) is. Another straight forward application of the Hahn-Banach theorem yields that the vector-valued version of Theorem 2.4 holds, and consequently also the qualita- tive versions (16) and (17) of Bohr's theorem (where in all inequalities the absolute value of course has to be replaced by the norm in X). Given 1 ≤ p ≤ ∞, X a Banach space and λ a frequency, the definition of Hardy spaces of X-valued Dirichlet series extends naturally to the vector-valued situation. As in the scalar case we for any λ-Dirichlet group (G, β) define the Banach space p (G, X) :=nf ∈ Lp(G, X) f : G → X is supported on all hlog n, n ∈ No , and consider the vector-valued Bohr map H λ B : H λ p (G, X) ֒→ D(λ, X), f 7→ D =X f (hλn)e−λns . Then it again turns out that the Hardy space Hp(λ, X) := B(H λ p (G, X)) 24 DEFANT AND SCHOOLMANN together with the norm kDkp := kfkp of all X-valued λ-Dirichlet series, forms a Banach space which is independent of the chosen λ-Dirichlet group. In this final section of our survey we focus on two different topics from [16]. In the first one we ask up to which extend Theorem 4.1 transfers to the vector-valued setting? More precisely, for which Banach spaces X and frequencies λ does the equality D∞(λ, X) = H∞(λ, X) hold (with an coefficient preserving isometry). Secondly, we revisit the maximal inequalities from the Sections 4.6 and 4.8 in the vector-valued case. 5.1. Coincidence of vector-valued H∞H∞H∞'s. We recall that a Banach space X has the analytic Radon Nikodym property (short ARNP) if every f ∈ H∞(D, X) has radial limits in almost all w ∈ T. This happens if and only if H∞(T, X) = H∞(D, X) (via an isometry identifying Fourier and monomial coefficients). Then it is straight forward to see that D∞((n), X) = H∞((n), X) if and only if X has ARNP. The main result of [27] (see also [26, Theorem 24.17]) shows that this result holds true for ordinary Dirichlet series, so for the frequency λ = (log n), and in the following theorem we state an extension of all this for general Dirichlet series which is proven in [16]. Theorem 5.1. Let λ be any frequency and X a non trivial Banach space with ARNP. Then D∞(λ) = H∞(λ) if and only if (27) D∞(λ, X) = H∞(λ, X) . In particular, (27) holds whenever X has ARNP and λ satisfies (BC). But in fact there are Banach spaces X without ARNP as well as frequencies λ for which D∞(λ) = H∞(λ), such that (27) is no equality. To see this we extend Theorem 2.9 to the vector-valued case. Denote by ℓw 1 (X) the Banach space of all weakly summable X-valued sequences (an) with norm w((an)) = sup x′∈BX ′ x′(an) , ∞Xn=1 1 (X) the closed subspace of ℓw and by ℓw,0 1 (X) consisting of all sequences (an) which are unconditionally summable. Recall that ℓw,0 1 (X) if and only if X does not contain an isomorphic copy of c0. The following theorem is the announced extension of Theorem 2.9 from [16]. (X) = ℓw 1 Theorem 5.2. If λ is Z-linearly independent, then for every Banach space X the following isometric inclusions hold: ℓw,0 1 (X) ⊂ H∞(λ, X) ⊂ D∞(λ, X) ⊂ ℓw 1 (X) In particular, D∞(λ, X) = H∞(λ, X) holds isometrically whenever c0 is no iso- morphic subspace of X. HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 25 Banach lattices have ARNP if and only if they don't contain an isomorphic copy of c0. But in contrast to this, there are Banach spaces which fail to have ARNP as well as an isomorphic copy of c0 (see [13, §4]). Consequently, Theorem 5.2 shows that if λ is Z-linearly independent, then the equality H∞(λ, X) = D∞(λ, X) does not necessarily imply that X has ARNP. 5.2. Maximal inequalities - vector-valued. Now we discuss the inequalities from the Sections 4.2, 4.6 and 4.8. Recall that our proof of Theorem 4.3 relies explicitely on the boundedness of the Riesz projection on Lp(G), where G is some Dirichlet group. This fact remains valid for X-valued functions whenever X is a so-called UMD-space which, as observed in [16], is an immediate consequence of results from [2]. Theorem 5.3. Let G be a Dirichlet group, X a Banach space and 1 < p < ∞. Then X has UMD if and only if the Riesz projection R : Lp(G, X) → Lp(G, X), f 7→Xx≥0 bf (hx)hx is bounded. As done in [24] for the scalar case, a standard argument leads to the following consequence generalizing Theorem 4.3 (see once again [16]). Corollary 5.4. For every frequency λ and every UMD-space X we have sup N (cid:13)(cid:13)(cid:13)Sp,X N : Hp(λ, X) → Hp(λ, X), X ane−λns 7→ NXn=1 ane−λns(cid:13)(cid:13)(cid:13) < ∞. Let us finally again turn to the Carleson-Hunt theorem. The proof of Theorem 4.18 given in [25], extends in the following sense from scalar to vector-valued functions. Theorem 5.5. Assume that X is a Banach space for which the Carleson-Hunt maximal inequality holds, i.e. for every 1 < p < ∞ there is some C > 0 such that for every f ∈ Lp(T, X) (28) Then for every frequency λ of integer type, every λ-Dirichlet group (G, β), and every 1 < p < ∞ we have that for every f ∈ H λ f (hλn)hλn f = almost everywhere on G, and (29) (cid:13)(cid:13)(cid:13)(cid:13) sup N (cid:13)(cid:13)(cid:13) NXn=1 where C = C(p) is a constant which only depends on p. (cid:13)(cid:13)(cid:13)(cid:13) sup N (cid:13)(cid:13)(cid:13) NXn=−N p (G, X) f (n)zn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T) ≤ CkfkLp(T,X) . ∞Xn=1 f (hλn)hλn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Ckfkp , 26 DEFANT AND SCHOOLMANN In particular, if X is a Banach lattice with UMD, then (28) holds true (see e.g. [31]), hence in this case its X-valued variant (29) is valid. Actually, it is proven in [16] that for Z-linearly independent frequencies λ the assumption UMD on X in fact is superfluous. Proposition 5.6. If λ is Z-linearly independent, then (29) holds for all Banach spaces X. Since the proof of Theorem 4.8 needs for example tools like the Hausdorff-Young inequality, there so far seems no satisfying answer to the following final problem. Problem 5.7. Under which assumptions on the Banach space X and the frequency λ for all f ∈ H λ p (G, X) and u > 0 ZG sup N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) NXn=1 bf (hλn)e−uλnhλn(ω)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p dω! 1 p X ≤ Ckfkp ? If there is such an estimate, one may ask on which of the parameters λ, X, p and u the constant C actually depends. Recall that in Theorem 4.8, which deals with the scalar case X = C and λ's with (BC), the constant C in (23) does not depend on p (but on u and λ). This is a useful fact since by Remark 4.11, Theorem 4.8 then implies Bohr's theorem for λ's with (BC). Since Bohr's theorem in the vector-valued case holds for every Banach space X provided that λ has (LC), Problem 5.7 may hold for every Banach space X and λ's satisfying (LC) with some constant C independent of p. If we ask for a constant C depending on p, then in view of Theorem 5.5 it may happen that Problem 5.7 has an affirmative answer for every frequency λ but with restrictions on the Banach space X. References [1] A. Aleman, J.F. Olsen, and E. Saksman, Fourier multipliers for Hardy spaces of Dirichlet series, International Mathematics Research Notices (16) (2014) 4368-4378 [2] N. H.Asmar, B. P. Kelly, and S. Montgomery-Smith. A note on UMD spaces and trans- ference in vector-valued function spaces, in Proceedings of the Edinburgh Mathematical Society 39(3) (1996) 485-490 [3] F. Bayart, Hardy spaces of Dirichlet series and their compostion operators, Monatsh. Math. 136(3) (2002) 203-236 [4] R. Balasubramanian, B. Calado, and H. Queffelec, The Bohr inequality for ordinary Dirich- let series, Studia Math. 175(39) (2006) 285-304 [5] H. F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. 32(3) (1931) 600-622 [6] A. S. Besicovitch, Almost periodic functions, Dover publications (1954) [7] H. Bohr, Uber die gleichmassige Konvergenz Dirichletscher Reihen, J. Reine Angew. Math. 143 (1913) 203-211 [8] H. Bohr, Uber die Bedeutung der Potenzreihen unendlich vieler Variabeln in der Theorie ns , Nachr. Ges. Wiss. Gott. Math. Phys. Kl. 4 (1913) 441-488 [9] H. Bohr, Einige Bemerkungen uber das Konvergenzproblem Dirichletscher Reihen, Rendi- der Dirichletschen ReihenP an conti del circolo Matematica di Palermo 37 (1913) 1-16 [10] H. Bohr, Zur Theorie der fastperiodischen Funktionen II, Acta Math. 46 (1925) 101-214 [11] H. Bohr, En bemaerkning om Dirichletske raekkers ligelige konvergens, Mat. Tidsskr. B (1951) 1-8 HARDY SPACES OF GENERAL DIRICHLET SERIES -- A SURVEY 27 [12] H. Bohr, Losung des absoluten Konvergenzproblems einer allgemeinen Klasse Dirichletscher Reihen, Acta Math. 36(1) (1913) 197-240. [13] A. V. Bukhvalov and A. A. Danilevich, Boundary properties of analytic and harmonic functions with values in Banach spaces, Mat. Zametki 31,2 (1982) 203-214 [14] D. Carando, A. Defant, and P. Sevilla-Peris, Bohr's absolute convergence problem for Hp- Dirichlet series in Banach spaces, Anal PDF 7(2) (2014) 513-527 [15] D. Carando, A. Defant, and P. Sevilla-Peris, Almost sure-sign convergence of Hardy-type Dirichlet series, J. Anal. Math. 135(19) (2018) 225-247 [16] D. Carando, A. Defant, F. Marceca, and I. Schoolmann, Vector-valued Hardy spaces of general Dirichlet series, in preparation [17] D. Carando, F. Macerca, M. Scotti, and P. Tradacette, Random unconditional convergence of vector-valued Dirichlet series, preprint 2019 [18] J. Castillo-Medina, Domingo Garc´ıa, and M. Maestre, Isometries between spaces of multiple Dirichlet series, J. Math. Anal. and Appl., to appear 2019 [19] Y. S. Choi, U. Y. Kim, and M. Maestre, Banach spaces of general Dirichlet series, J. Math. Anal. and Appl. 465 (2018) 839-856 [20] A. Defant, D. Garc´ıa, M. Maestre, and D. P´erez-Garc´ıa, Bohr's strip for vector-valued Dirichlet series, Math. Ann. 342(3) (2008) 53-555 [21] A. Defant, D. Popa, and U. Schwarting, Coordinatewise multiple summing operators in Banach spaces, J. Funct. Anal. 259(1) (2010) 220-242 [22] A. Defant, U. Schwarting, and P. Sevilla-Peris, Estimates for vector-valued Dirichlet series, Monatsh. Math. 175(1) (2014) 89-116 [23] A. Defant and P. Sevilla-Peris, Convergence of Dirichlet polynomials in Banach spaces, Trans. Amer. Math. Soc. 363(23) (2011) 681-697 [24] A. Defant I. Schoolmann, Hp-theory of general Dirichlet series, preprint 2018 [25] A. Defant and I. Schoolmann, On a theorem for Helson for general Dirichlet series, in preparation [26] A. Defant, D. Garc´ıa, M. Maestre, and P. Sevilla Peris, Dirichlet series and holomor- phic functions in high dimensions, to appear in: New Mathematical Monographs Series, Cambrigde University Press 2019 [27] A. Defant and A. P´erez, Hardy spaces of vector-valued Dirichlet series, Studia Math., 243(1) (2018) 53-78 [28] B. J. Cole and T. W. Gamelin, Representing measures and Hardy spaces for the infinite polydisk algebra, Proc. London Math. Soc. 53(1) (1986) 112-142 [29] R. Doss, One the Fourier-Stieltjes transforms of singular or absolutely continuous measures, Math. Zeitschr. 97 (1967) 77-84 [30] C. Fefferman, On the convergence of multiple Fourier series, Bull. Amer. Math. Soc. 77 (1971) 744-745 [31] J. L. Rubio de Francia, Martingale and integral transforms of Banach space valued functions, In Probability and Banach spaces (Zaragoza, 1985), Springer, Lecture Notes in Math. 1221 (1986) 195-222 [32] G. H. Hardy, and M. Riesz, The general theory of Dirichlet series, Cambridge Tracts in Mathematics and Mathematical Physics 18 (1915) [33] H. Hedenmalm, P. Lindqvist, and K. Seip, A Hilbert space of Dirichlet series and systems of dilated function in L2(0, 1), Duke Math. J. 86 (1997) 1-37 [34] H. Hedenmalm and E. Saksman, Carleson's convergence theorem for Dirichlet series, Pacific J. of Math. 208 (2003) 85-109 [35] H. Helson, Compact groups and Dirichlet series, Ark. Mat. 8 (1969) 139-143 [36] H. Helson, Dirichlet series, Regent Press, 2005 [37] H. Helson and D. Lowdenslager. Prediction theory and Fourier series in several variables, Acta Math. 99 (1958) 165-202 28 DEFANT AND SCHOOLMANN [38] E. Landau, Uber die gleichmassige Konvergenz Dirichletscher Reihen, J. Reine Angew. Math. 143 (1921) 203-211 [39] L. Neder, Zum Konvergenzproblem der Dirichletschen Reihen beschrankter Funktionen, Math. Zeitschrift 14 (1922) 149-158 [40] O. Perron, Zur Theorie der Dirichletschen Reihen, J. Reine Angew. Math. 134 (1908) 95-143 [41] H. Queff´elec and M. Queff´elec, Diophantine approximation and Dirichlet series, Hindustan Book Agency, Lecture Note 2 (2013) [42] W. Rudin, Fourier analysis on groups, Interscience Publishers (1962) [43] I. Schoolmann, On Bohr's theorem for general Dirichlet series, preprint 2019 [44] G. Valiron, Th´eorie g´en´erale des s´eries de Dirichlet, M´emorial des Sience Math´ematiques, 17, 1926 [45] F. B. Weissler, Logarithmic Sobolev inequalities and hypercontractive estimates on the circle, J. Funct. Anal. 37(2) (1980) 218-234 Andreas Defant Institut fur Mathematik, Carl von Ossietzky Universitat, 26111 Oldenburg, Germany. E-mail address: [email protected] Ingo Schoolmann Institut fur Mathematik, Carl von Ossietzky Universitat, 26111 Oldenburg, Germany. E-mail address: [email protected]
1103.1030
1
1103
2011-03-05T09:29:07
Quasi-nearly subharmonic functions and quasiconformal mappings
[ "math.FA" ]
We prove that the composition of a quasi-nearly subharmonic function and a quasiregular mappings of bounded multiplicity is quasi-nearly subharmonic. Also, we prove that if $u\circ f$ is quasi-nearly subharmonic for all quasi-nearly subharmonic $u$ and $f$ satisfies some additional conditions, then $f$ is quasiconformal. Similar results are further established for the class of regularly oscillating functions.
math.FA
math
QUASI-NEARLY SUBHARMONIC FUNCTIONS AND QUASICONFORMAL MAPPINGS PEKKA KOSKELA AND VESNA MANOJLOVI ´C Dedicated to Professor Miroslav Pavlovic Abstract. We prove that the composition of a quasi-nearly subharmonic function and a quasiregular mappings of bounded multiplicity is quasi-nearly subharmonic. Also, we prove that if u ◦ f is quasi-nearly subharmonic for all quasi-nearly subharmonic u and f satisfies some additional conditions, then f is quasiconformal. Similar results are further established for the class of regularly oscillating functions. 1. Introduction and results Let Ω be a domain in the Euclidean space Rn. If h is a function harmonic in Ω, then the function hp, which need not be subharmonic in Ω for 0 < p < 1, behaves like a subharmonic function: the inequality (1.1) h(a)p ≤ C rn ZB(a,r) hp dm, B(a, r) ⊂ Ω, 0 < p < ∞, holds, whenever 0 < p < ∞, B(a, r) = {x : x−a < r} ⊂ Ω, and dm is the Lebesgue measure normalized so that B(0, 1) := m(B(0, 1)) = 1. The constant C in (1.1) depends only on n and p when p < 1, and C = 1 when p ≥ 1. This fact is essentially due to Hardy and Littlewood (see [5, Theorem 5]), although they never formulated it. The proof was first given by Fefferman and Stein [4], and independently by Kuran [11]. It follows from Fefferman and Stein's proof that (1.1) remains true if h is replaced by a nonnegative subharmonic function. Hence: Theorem 1.1. If u ≥ 0 is a function subharmonic in a domain Ω ⊂ Rn, then (1.2) u(a)p ≤ C rn ZB(a,r) up dm, B(a, r) ⊂ Ω, 0 < p < ∞, where C depends only on p and n, when p < 1, and C = 1 when p ≥ 1. Let u ≥ 0 be a locally bounded, measurable function on Ω. We say (see [12], [13]) that u is C-quasi-nearly subharmonic (abbreviated C-qns) if the following condition is satisfied: (1.3) u(a) ≤ C rn ZB(a,r) u dm, whenever B(a, r) ⊂ Ω. One can view (1.3) as a weak mean value property. Besides of nonnegative subharmonic functions it also holds for nonnegative subsolutions to a large family 1991 Mathematics Subject Classification. 31C05,30C65. The first author was supported by grants from the Academy of Finland and the second author by MN Project 174024, Serbia. 1 2 PEKKA KOSKELA AND VESNA MANOJLOVI ´C of second order elliptic equations, see [6]. In fact, (1.3) is typically proven as a step towards Harnack inequalities for second order elliptic equations, using the Moser iteration scheme. Notice that u is quasi-nearly subharmonic if and only if u is everywhere dominated by its centered minimal function [2]. Our first result is an invariance property. Theorem 1.2. If u ≥ 0 is a C-qns function defined on a domain Ω′ ⊂ Rn, n ≥ 2, and f is a K-quasiregular mapping, with bounded multiplicity N , from a domain Ω onto Ω′, then the function u ◦ f is C1-qns in Ω, where C1 only depends on K, C, N , and n. Remark 1.1. The hypothesis of bounded multiplicity of f is necessary as the fol- lowing example shows. Let f (z) = ez, Ω = C, Ω′ = C \ {0}, E = [j≥2 [exp(2j), exp(2j + 1)], and u(w) = χE(w). Then it is easy to check that u is quasi-nearly subharmonic in Ω′ but u ◦ f is not quasi-nearly subharmonic in Ω. Above quasiregularity requires that f is continuous, the component functions of f belong locally to the Sobolev class W 1,n and that there is a constant K ≥ 1 so that Df (x)n ≤ KJ(x, f ) holds almost everywhere in Ω. Injective quasiregular mappings are called quasicon- formal. It was previously only known that the invariance holds under conformal mappings in the planar case [10] and under bi-Lipschitz mappings [3] in all dimen- sions. Let us consider the above morphism property in more detail. Definition 1.1. Let Ω and Ω′ be subdomains of Rn. A mapping f : Ω 7→ Ω′ is a qns-morphism if there is a constant C < ∞ such that for every qns u defined in Ω′ we have ku ◦ f kqns ≤ Ckukqns, where kukqns = inf nC ≥ 0 : u(a) ≤ C rn ZB(a,r) u dm for all a ∈ Ω′, 0 < r ≤ d(x, ∂Ω′)o. If the above holds with a constant C, we call f a C-qns-morphism. Finally, f is a strong qns-morphism if there is a constant C so that f restricted to any domain G ⊂ Ω, f : G 7→ G′, is a C-qns-morphism. Theorem 1.3. Let Ω, Ω′ ⊂ Rn, n ≥ 2, be domains. Then a homeomorphism f : Ω 7→ Ω′ is a strong qns-morphism if and only if f is quasiconformal. If we assume sufficient a priori regularity for f, a version of Theorem 1.3 holds also for qns-morphisms. The reader may wish to compare this with related quasi- conformal invariance properties for other function classes [1], [14], [16], [17], [18]. Theorem 1.4. Let n ≥ 2 and let f : Ω 7→ Ω′ be a qns-morphism that belongs to W 1,n loc . If, additionally, J(x, f ) ≥ 0 almost everywhere, then f is quasiregular. Each quasiregular mapping f is either constant or both open and discrete; in the latter case the multiplicity of f is locally finite. The condition J(x, f ) ≥ 0 in Theorem 1.4 cannot be dropped, as the mapping f (x, y) = (x, y) is a planar QUASI-NEARLY SUBHARMONIC FUNCTIONS AND QUASICONFORMAL MAPPINGS 3 (strong) qns-morphism. The Sobolev regularity assumption can be slightly relaxed: if f above is a C-qns-morphism, then local p-integrability of the distributional derivatives suffices for p = p(n, C) < n; this can be inferred from the proof of Theorem 1.4 using [9]. In the planar, injective setting, even W 1,1 loc suffices. Let us close this introduction by commenting on the invariance of a related function class, introduced in [12]. Definition 1.2. A function u : Ω′ 7→ Rk is said to be regularly oscillating if x ∈ Ω′, B(x, r) ⊂ Ω′, Lip u(x) ≤ Cr−1 (1.4) sup u(y) − u(x), y∈B(x,r)⊂Ω′ where C ≥ 0 is a constant independent of x and r. Here u(y) − u(x) Lip u(x) = lim sup . y→x y − x Note that Lip u(x) = grad u(x) if u is differentiable at x. The smallest C satisfying (1.4) will be denoted by kukro. We have the following invariance. Theorem 1.5. Let f : Ω 7→ Ω′ be quasiregular, regularly oscillating and of bounded multiplicity in Ω. If u is regularly oscillating in Ω′, then u ◦ f is regularly oscillating in Ω with ku ◦ f kro ≤ C′kukro, where C′ depends only on the multiplicity of f . The assumption that f be regularly oscillating is necessary, as seen by notic- ing that the coordinate projections are regularly oscillating; not all quasiregular mappings are regularly oscillating. In the case of an analytic function, this can naturally be dropped. Similarly to Theorem 1.4, quasiregularity is necessary if we assume that J(x, f ) ≥ 0 almost everywhere, but no a priori Sobolev regularity is needed because regularly oscillating functions and mappings are locally Lipschitz continuous. The invariance property of Theorem 1.5 was established in [10] when f is conformal (and n = 2). Remark 1.2. The assumption of bounded multiplicity of f in Theorem 1.5 is nec- essary as in the case of Theorem 1.2. To see this simply let f (z) = ez, Ω = C, Ω′ = C \ {0}, E = [j≥2 [exp(2j), exp(2j + 1)], and v(w) = R w 0 χE(t) dt. Then v is regularly oscillating but v ◦ f is not. 2. Proof of the theorem 1.2 For the proof we need some lemmas. The first says that if up is qns for some p, then so is u. Lemma 2.1. [12] If u is C-qns, and p > 0, then up is C1(C, n)-qns. We also need the following lemma that can be distilled from the arguments in [7]. For the sake of completeness, we give a short proof below. Lemma 2.2. Let f : Ω 7→ Ω′ be K-quasiregular and of bounded multiplicity N. Let x ∈ Ω and 0 < r ≤ 1 2 d(x, ∂Ωr)). Then d(cid:0)f (x), ∂f (B(x, r))(cid:1) ≥ δ sup f (y) − f (x), y∈B(x,r) 4 PEKKA KOSKELA AND VESNA MANOJLOVI ´C where δ = δ(n, K, N ). Proof. Let x ∈ Ω and let 0 < r ≤ 1 f (B(x, r)) because f is open. Moreover 2 d(x, ∂Ω). Now f (x) is an interior point of sup f (y) − f (x) = f (z) − f (x) y∈B(x,r) for some z ∈ ∂B(x, r), and 0 < d(f (x), ∂f (B(x, r)) = f (ω) − f (x) for some ω ∈ ∂B(x, r). Let E = [f (x), f (ω)] be the segment between f (w) and f (ω), and F be a segment that joins f (z) to ∂Ω′ (or to infinity) outside the ball We may assume that B(f (x), f (z) − f (ω)). f (z) − f (x) ≥ 2f (ω) − f (x). 1, 0, log f (z)−f (x) y−f (x) . log f (z)−f (x) f (ω)−f (x) y ∈ B(f (x), f (ω) − f (x)), y ∈ Bc(f (x), f (z) − f (x)), elsewhere. Then, by a change of variables, see page 21 in [15], Let u(y) =   ZΩ ∇(u ◦ f )n dm ≤ KZΩ′ ≤ KN ZΩ′ ∇un dm (∇u)(f (y))nJf (y) dm(y) ≤ KN Cn logn−1 f (z) − f (x) f (ω) − f (x) . On the other hand, since f is open, the set f −1([f (x), f (ω)]) contains a contin- 2 r). uum joining x to ∂B(x, r), and f −1(F ) a continuum joining ∂B(x, r) to ∂B(x, 3 By usual capacity estimates, see e.g. [6] ZΩ ∇(u ◦ f )n dm ≥ δ0(n, K) > 0. The claim follows. (cid:3) As an immediate consequence of Lemma 2.2 we have: Lemma 2.3. Let B = B(0, 1) and let f : 2B 7→ Ω′ be a K-qr mapping with bounded multiplicity N and such that f (0) = 0. Then there exist ρ ∈ (0, 1) and R > 0 such that B(0, R) ⊃ f (B) ⊃ B(0, ρ), where R/ρ ≤ 1/δ, and δ depends only on K, n, N. Finally we need the following fundamental fact: Lemma 2.4. [8, p. 258] Under the hypotheses of Lemma 2.3, there exists p > 1 such that (cid:16)ZB J(y, f )p dm(cid:17)1/p ≤ CZB J(y, f ) dm, where p depends only on K, N and n. QUASI-NEARLY SUBHARMONIC FUNCTIONS AND QUASICONFORMAL MAPPINGS 5 Proof of Theorem 1.2. As is easily seen, the proof reduces to the case Ω = B(0, 2), f (0) = 0. Let B = B(0, 1) and write v = u ◦ f. By using translations and rotations, we see that it is enough to prove that v(0) ≤ CZB v(y) dm(y), where C = C(K, n, kukqns). By Lemma 2.1, it suffices to find q = q(K, N, n) ≥ 1 so that (2.1) v(0) ≤ C(cid:16)ZB v(y)q dm(y)(cid:17)1/q . To prove this, we start from Holder's inequality: ZB v(y)J(y, f ) dm(y) ≤ (cid:16)ZB v(y)q dm(y)(cid:17)1/q(cid:16)ZB J(y, f )p dm(y)(cid:17)1/p , where p = q/(q − 1). By a change of variables, see page 21 in [15], we have u(y) N (y, f, B) dm(y) ZB v(y)J(y, f ) dm(y) = Zf (B) ≥ Zf (B) ≥ ZB(0,ρ) u(y) dm(y) u(y) dm(y) Here we have used Lemma 2.3 and the hypothesis that u is qns. On the other hand, by Lemmas 2.4 and 2.3, we have ≥ cρnu(0) = cρnv(0). (cid:16)ZB J(y, f )p dm(y)(cid:17)1/p dy ≤ CZB = CZf (B) J(y, f ) dm(y) N (y, f, B) dm(y) Combining these inequalities, we obtain ≤ CN f (B) ≤ CN B(0, R) = CN knRn. Hence cρnv(0) ≤ CN knRn(cid:16)ZB (cid:16)ZB CN knRn v(0) ≤ cρn . v(y)q dm(y)(cid:17)1/q v(y)q dm(y)(cid:17)1/q . Now the desired result follows from the inequality R/ρ ≤ 1/δ, where δ depends only on K, n, N. 3. Proof of Theorem 1.4 Even though our definition of quasiregular mappings requires them to be contin- uous, this condition is superfluous and it suffices to show that there exists K ≥ 1 so that Df (x)n ≤ KJ(x, f ) 6 PEKKA KOSKELA AND VESNA MANOJLOVI ´C holds almost everywhere, see e.g. page 177 in [15]. Next, every mapping f with Sobolev regularity W 1,1 loc is approximatively differentiable almost everywhere. That is, for almost every x0 and every ǫ > 0, the set Aǫ := {x : f (x) − f (x0) − Df (x0)(x − x0) x − x0 < ǫ} has density one at x0, see e.g. page 140 in [19]. Because of our a priori Sobolev regularity, it thus suffices to show the above distortion inequality at every such point x0. For simplicity, we only give the proof in the planar case, assuming differentiability instead of approximate differenentiability. The higher dimensional setting and the switch to approximate differentiability only require technical modifications that should be obvious to the reader after examining the argument below. Thus suppose that f is differentiable at x0. Case (a). Suppose f is differentiable at x0 with Jf (x0) 6= 0. In some coordinate systems we have Df (x0) = (cid:20)a 0 b(cid:21) 0 Assume 0 < a < b. We want to show that b/a is bounded. Consider the function u(ω) = χ{ω=x′+iy ′:0≤y ′≤x′}(ω − f (x0)). Then kukqns = 8. Now T −1(cid:0){ω = x′ + iy′ : 0 ≤ y′ ≤ x′}(cid:1) = {z = x + iy : 0 ≤ by ≤ ax} = {z = x + iy : 0 ≤ y ≤ (a/b)x} for the linear transformation T associated to Df (x0). Now u ◦ f (x0) = 1. If r > 0 is such that B(x0, r) ⊂ Ω, we conclude from the morphism property of f that 1 ≤ C r2 ZB(0,r) u ◦ f dm = → C r2 C 2 1 2 a b r2 arctan a b + o(r) , when r → 0, where C > 0 comes from the morphism property. Hence b/a ≤ C/2. Case (b). Now suppose that f is differentiable at x0 with Jf (x0) = 0. We want to prove that Df (x0) = 0. We argue by contradiction: suppose Df (x0) 6= 0. We may assume Df (x0) = (cid:20)0 0 0 1(cid:21) . Define u(ω) = χ{ω=x′+iy ′:0≤y ′≤x′}(ω − f (x0)). Then kukqns = 4, and f −1(t + it + f (x0)) = s1(t) + is2(t) + x0. QUASI-NEARLY SUBHARMONIC FUNCTIONS AND QUASICONFORMAL MAPPINGS 7 where limt→0 s2(t) s1(t) = 0. But then there is no C > 0 such that (u ◦ f )(x0) ≤ C r2 ZB(x0,r) u ◦ f dm for all small r > 0, which contradicts the morphism property of f. 4. Proof of Theorem 1.3 It is an immediate consequence of Theorem 1.2 that quasiconformality of the homeomorphism f is a sufficient condition for f to be a strong qns-morphism. In the other direction, it suffices to prove that f −1 : Ω′ 7→ Ω is quasiconformal. Thus it suffices to verify the existence of H < ∞ such that (4.1) lim sup r→0 diam (f −1(B(y, r)))n f −1(B(y, r)) ≤ H for all y ∈ Ω′, see page 64 in [8]. To simplify our notation, we write x′ = f (x) for x ∈ Ω, in what follows. Fix y′ ∈ Ω′ and let r > 0. Towards proving (4.1), we may assume that r is so small that B(y, 2 diam (f −1(B(y′, 2r)))) ⊂ Ω. Fix y′ 0 ∈ ∂B(y′, 2r) and pick y′ 1 ∈ ∂B(y′, r) so that y′ 0 − y′ 1 = max ω′∈∂B(y ′,r) ω − y′ 0. Set G′ = Ω′ \ {y′ 0} and G = Ω \ {y0}. Now B(y1, y1 − y0/2) ⊂ G and (4.2) Define u(ω′) = χB(y ′,r)(ω′) for ω′ ∈ G′. Then u is qns in G′ with kukqns ≤ 3n. Since f is is C-qns-morphism in G, we conclude that diam (f −1(B(y′, r)) ≤ 2y1 − y0. u ◦ f (y1) ≤ C3n 1 B(y1, y1 − y0/2) ZB(y1,y1−y0/2) u ◦ f dm = C3n f −1(B(y′, r)) ∩ B(y1, y1 − y0/2) B(y1, y1 − y0/2) Recalling (4.2) and that u ◦ f (y1) = u(y′ 1) = 1, we arrive at diam (f −1(B(y′, r))n ≤ CCn(f −1(B(y′, r)) as desired. . 5. Proof of Theorem 1.5 5.1. Proof of Theorem 1.5. Let x ∈ Ω and 0 < r < 1 f is regularly oscillating and quasiregular we have, by Lemma 2.2, 2 d(x, ∂Ω). Since the mapping Lip f (x) ≤ Cr−1 sup f (y) − f (x) y∈B(x,r) ≤ C δ r−1d(f (x), ∂f (B(x, r))). 8 PEKKA KOSKELA AND VESNA MANOJLOVI ´C Recall that non-constant quasiregular mappings are open. Since u is regularly oscillating and d(f (x), ∂f (B(x, r))) > 0, we have that Lip u(f (x)) ≤ Cd(f (x), ∂f (B(x, r)))−1 sup u(f (x)) − z z∈B(f (x),d(f (x),∂f (B(x,r))) ≤ Cd(f (x), ∂f (B(x, r)))−1 sup u ◦ f (y) − u ◦ f (x). y∈B(x,r) Now we have Lip (u ◦ f )(x) ≤ Lip (u(f (x)) Lip f (x) ≤ Cd(f (x), ∂f (B(x, r)))−1 sup u ◦ f (y) − u ◦ f (x) y∈B(x,r) × C δ r−1d(f (x), ∂f (B(x, r))) = C′r−1 sup u ◦ f (y) − u ◦ f (x). y∈B(x,r) This completes the proof of the theorem. References 1. K. Astala, A remark on quasiconformal mappings and BMO-functions, Michigan Math. J. 30 (1983), 209-212. 2. D. Cruz-Uribe, The minimal operator and the geometric maximal operator in Rn, Studia Math. 144 (2001), 1-37. 3. O. Dovgoshey and J. Riihentaus, Bi-Lipschitz mappings and quasinearly subharmonic func- tions, Int. J. Math. Math. Sci. 2010, Art. ID 382179, 8 pp. 4. C. Fefferman and E.M. Stein, H p-spaces of several variables, Acta Math. 129 (1972), 137 -- 193. 5. G.H. Hardy and J.E. Littlewood, Some properties of conjugate functions, J. Reine Angew. Math. 167 (1931), 405 -- 423. 6. J. Heinonen, T. Kilpelainen and O. Martio, Nonlinear potential theory of degenerate elliptic equations, Oxford Mathematical Monographs. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1993. vi+363 pp. 7. J. Heinonen and P. Koskela, Weighted Sobolev and Poincar´e inequalities and quasiregular mappings of polynomial type, Math. Scand. 77 (1995), 251 -- 271. 8. J. Heinonen and P. Koskela, Definitions of quasiconformality, Invent. Math. 120 (1995), 61-79. 9. T. Iwaniec, p-harmonic tensors and quasiregular mappings, Ann. of Math. (2) 136 (1992), 589-624. 10. V. Koji´c, Quasi-nearly subharmonic functions and conformal mappings, Filomat (Nis), 21:2(2007), 243 -- 249. 11. U. Kuran, Subharmonic behaviour of hp (p > 0, h harmonic), J. London Math. Soc. 8(2) (1974), 529 -- 538. 12. M. Pavlovi´c, On subharmonic behaviour and oscillation of functions on balls in Rn, Publ. Inst. Math. (Belgrade) 55(1994), 18 -- 22. 13. M. Pavlovi´c and J. Riihentaus, Classes of quasi-nearly subharmonic functions, Potential Anal- ysis 29:1(2008), 89 -- 104. 14. H. M. Reimann, Functions of bounded mean oscillation and quasiconformal mappings, Com- ment. Math. Helv. 49 (1974), 260-276. 15. S. Rickman, Quasiregular mappings. Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], 26. Springer-Verlag, Berlin, 1993. x+213 pp. 16. S. G. Staples, Maximal functions, A∞ measures and quasiconformal maps, Proc. Amer. Math. Soc. 113 (1991), no. 3, 689-700 17. S. G. Staples, Doubling measures and quasiconformal maps, Comment. Math. Helv. 67 (1992), 119-128. 18. A. Uchiyama, Weight functions of the class (A∞) and quasi-conformal mappings, Proc. Japan Acad. 51 (1975), suppl., 811-814. QUASI-NEARLY SUBHARMONIC FUNCTIONS AND QUASICONFORMAL MAPPINGS 9 19. W. P. Ziemer, Weakly differentiable functions. Sobolev spaces and functions of bounded vari- ation, Graduate Texts in Mathematics, 120. Springer-Verlag, New York, 1989. xvi+308 pp. Department of Mathematics and Statistics, P.O. Box 35, FIN-40014, University of Jyvaskyla, Finland E-mail address: [email protected] University of Belgrade, Faculty of Organizational Sciences, Jove Ilica 154, Bel- grade, Serbia E-mail address: [email protected]
1902.05692
1
1902
2019-02-15T05:44:09
Non-linear functionals, deficient topological measures, and representation theorems on locally compact spaces
[ "math.FA" ]
We study non-linear functionals, including quasi-linear functionals, p-conic quasi-linear functionals, d-functionals, r-functionals, and their relationships to deficient topological measures and topological measures on locally compact spaces. We prove representation theorems and show, in particular, that there is an order-preserving, conic-linear bijection between the class of finite deficient topological measures and the class of bounded p-conic quasi-linear functionals. Our results imply known representation theorems for finite topological measures and deficient topological measures. When the space is compact we obtain four equivalent definitions of a quasi-linear functional and four equivalent definitions of functionals corresponding to deficient topological measures.
math.FA
math
Non-linear functionals, deficient topological measures, and representation theorems on locally compact spaces Svetlana V. Butler∗ February 18, 2019 Abstract We study non-linear functionals, including quasi-linear functionals, p-conic quasi-linear functionals, d-functionals, r-functionals, and their relationships to deficient topological measures and topological measures on locally compact spaces. We prove representation theorems and show, in particular, that there is an order-preserving, conic-linear bijection between the class of finite deficient topological measures and the class of bounded p-conic quasi-linear functionals. Our results imply known representation theorems for finite topo- logical measures and deficient topological measures. When the space is compact we obtain four equivalent definitions of a quasi-linear functional and four equivalent definitions of functionals corresponding to deficient topological measures. AMS Subject Classification (2010): Primary 28A25, 28C05, 46G99; Secondary 46E27, 28C15 Keywords: quasi-linear functional, p-conic quasi-linear functional, r-functional, s-functional, deficient topological measure, topological measure, right and left measure, representation theorem, locally compact space 1 Introduction This paper is devoted to the study of various non-linear functionals and their relationships to deficient topological measures and topological measures on locally compact spaces. J. F. Aarnes first discovered quasi-linear functionals and the corresponding set functions, topo- logical measures, (initially called quasi-measures) in [1]. Since then many works devoted to these objects have appeared. For their use in symplectic topology one may consult numerous papers beginning with [11] and a monograph ([12]). We prove representation theorems for deficient topological measures on locally compact spaces. Our results imply known results, including the representation theorem for quasi-linear functionals on compact spaces ([1]), the representation theorem for deficient topological mea- sures on compact spaces ([15]), and the representation theorems for quasi-linear functionals on locally compact spaces ([13],[5]). We also obtain new consequences where X is compact, including four equivalent definitions of a quasi-linear functional and four equivalent definitions of functionals corresponding to deficient topological measures. ∗Department of Mathematics, University of California, Santa Barbara, 552 University Rd, Isla Vista, CA 93117, USA. Email: [email protected] 1 Deficient topological measures were first defined and used by A. Rustad and ∅. Johansen In these in [10]. They were later independently rediscovered by M. Svistula ([14], [15]). works, deficient topological measures were defined as real-valued functions on a compact space. In this paper we use deficient topological measures on locally compact spaces as functions into extended real numbers. Quasi-linear functionals have been generalized to other non-linear functionals in order to represent deficient topological measures on compact spaces. See [10, section 6]; see also [15], where such functionals are called r- and l-functionals. To represent deficient topological measures on locally compact spaces we will need the above mentioned non-linear functionals and some others. This paper is influenced by many works, including [1], [10], [15], [13], [8], and [5]. The paper is organized as follows. Section 2 contains necessary definitions and background facts. In Section 3 we consider cones generated by a function, as well as p-conic and n- conic quasi-linear functionals. In Section 4 we consider various related non-linear functionals, including d-, r-, l-, and s-functionals. In Section 5 we show how to obtain deficient topological measures from d-functionals. In Section 6, given a finite deficient topological measure and a bounded continuous function, we define left and right measures. We give a criterion for left and right measures to coincide. Left and right measures coincide if one starts from a finite topological measure; we also give examples which show that left and right measures may or may not coincide even for {0, 1}-valued deficient topological measures. In section 7 using right and left measures we obtain by integration p-conic and l-conic quasi-linear functionals, which are also r- and l-functionals. Section 8 is devoted to representation theorems for finite deficient topological measures in terms of bounded p-conic and l-conic quasi-linear functionals, and also r- and l-functionals. In particular, we show that there is an order-preserving, conic-linear bijection between the class of finite deficient topological measures and the class of bounded p-conic quasi-linear functionals. When X is compact, we obtain four equivalent definitions of quasi-linear functionals and four equivalent definitions of functionals corresponding to deficient topological measures. 2 Preliminaries In this paper X is a locally compact, connected space. By C(X) we denote the set of all real-valued continuous functions on X with the uniform norm, by Cb(X) the set of bounded continuous functions on X, by C0(X) the set of continuous functions on X vanishing at infinity, and by Cc(X) the set of continuous functions with compact support. By C + 0 (X) we denote the collection of all nonnegative functions vanishing at infinity; similarly, C − 0 (X) is the collection of all nonpositive functions from C0(X). When we consider maps into [−∞, ∞] we assume that any such map attains at most one of ∞, −∞, and is not identically ∞ or −∞. By D(ρ) we denote the domain of a functional ρ. For example, we may take D(ρ) = C + 0 (X) or Cc(X). If D(ρ) = C(X), where X is compact, then D(ρ) contains constants. We denote by E the closure of a set E, and by F a union of disjoint sets. We denote by 1 the constant function 1(x) = 1, by id the identity function id(x) = x, and by 1K the characteristic function of a set K. By supp f we mean {x : f (x) 6= 0}. Several collections of sets are used often. They include: O(X), the collection of open subsets of X; C (X) the collection of closed subsets of X; K (X) the collection of compact subsets of X. 2 Definition 2.1. Let X be a topological space and µ be a nonnegative set function on E, a family of subsets of X that contains O(X) ∪ C (X). We say that • µ is inner regular (or inner compact regular) if µ(U ) = sup{µ(C) : C ⊆ U, C ∈ K (X)} for each open set U . • µ is outer regular if µ(F ) = inf{µ(U ) : F ⊆ U, U ∈ O(X)} for each closed set F . • k µ k= µ(X) and µ is finite if µ(X) < ∞. • µ is compact-finite if µ(K) < ∞ for any K ∈ K (X). • µ is monotone if A ⊆ B implies µ(A) ≤ µ(B). • µ is τ -smooth on compact sets if for every decreasing net Kα ց K, Kα, K ∈ K (X) we have µ(Kα) → µ(K). • µ is τ -smooth on open sets if for every increasing net Uα ր U, Uα, U ∈ O(X) we have µ(Uα) → µ(U ). • µ is simple if it only assumes values 0 and 1. Definition 2.2. A deficient topological measure on a locally compact space X is a set function ν : C (X) ∪ O(X) −→ [0, ∞] which is finitely additive on compact sets, inner compact regular, and outer regular, i.e. : (DTM1) if C ∩ K = Ø, C, K ∈ K (X) then ν(C ⊔ K) = ν(C) + ν(K); (DTM2) ν(U ) = sup{ν(C) : C ⊆ U, C ∈ K (X)} for U ∈ O(X); (DTM3) ν(F ) = inf{ν(U ) : F ⊆ U, U ∈ O(X)} for F ∈ C (X). For a closed set F , ν(F ) = ∞ iff ν(U ) = ∞ for every open set U containing F . Remark 2.3. Note that a deficient topological measure ν is monotone on O(X) ∪ C (X) and ν(Ø) = 0. If ν and µ are deficient topological measures that agree on K (X) then ν = µ, and if ν ≤ µ on K (X) (or on O(X)) then ν ≤ µ. Definition 2.4. A topological measure on X is a set function µ : C (X) ∪ O(X) → [0, ∞] satisfying the following conditions: (TM1) if A, B, A ⊔ B ∈ K (X) ∪ O(X) then µ(A ⊔ B) = µ(A) + µ(B); (TM2) µ(U ) = sup{µ(K) : K ∈ K (X), K ⊆ U } for U ∈ O(X); (TM3) µ(F ) = inf{µ(U ) : U ∈ O(X), F ⊆ U } for F ∈ C (X). We denote by T M (X) the collection of all topological measures on X, by DT M (X) the collection of all deficient topological measures on X, and by M (X) the collection of all Borel measures on X that are inner regular on open sets and outer regular (restricted to O(X) ∪ C (X)). 3 Remark 2.5. Let X be locally compact. In general, M (X) $ T M (X) $ DT M (X). When X is compact, there are examples of topological measures that are not measures and of deficient topological measures that are not topological measures in numerous papers, be- ginning with [1], [10], and [14]. When X is locally compact, see [2], [4, Sections 5 and 6], and [3, Section 9] for more information on proper inclusion, criteria for a deficient topological measure to belong to M (X) or T M (X) (in particular, to be a Radon measure or a regular Borel measure), as well as various examples. The proof of the next result is in [4, Section 4]. Theorem 2.6. (I) Let X be compact, and ν a deficient topological measure. The following are equivalent: (a) ν is a topological measure. (b) ν(X) = ν(C) + ν(X \ C), C ∈ C (X). (c) ν(X) ≤ ν(C) + ν(X \ C), C ∈ C (X). (II) Let X be locally compact, and ν a deficient topological measure. The following are equivalent: (a) ν is a topological measure. (b) ν(U ) = ν(C) + ν(U \ C), C ∈ K (X), U ∈ O(X). (c) ν(U ) ≤ ν(C) + ν(U \ C), C ∈ K (X), U ∈ O(X). Recall the following fact (see, for example, [6, Chapter XI, 6.2]): Lemma 2.7. Let K ⊆ U, K ∈ K (X), U ∈ O(X) in a locally compact space X. Then there exists a set V ∈ O(X) with compact closure such that K ⊆ V ⊆ V ⊆ U. The following is proved in [4, Section 3]. Lemma 2.8. Let X be a locally compact space. (a) A deficient topological measure is τ -smooth on compact sets and τ -smooth on open sets. In particular, a topological measures is additive on open sets. (b) A deficient topological measure µ is superadditive, i.e. if Ft∈T At ⊆ A, where At, A ∈ O(X) ∪ C (X), and at most one of the closed sets (if there are any) is not compact, then µ(A) ≥Pt∈T µ(At). We recall the following theorem: Theorem 2.9 (Integration by parts for Lebesque-Stieltjes integrals). Let F and G be real- valued nondecreasing functions on R with corresponding Lebesque-Stieltjes measures mF , mG. Then for a < b we have: Z b a G(x+)dmF (x) +Z b a F (x−)dmG(x) = F (b+)G(b+) − F (a−)G(a−). 4 Definition 2.10. Let ρ be a functional on C(X). We say • ρ is homogeneous if ρ(af ) = aρ(f ) for any a ∈ R. • ρ is positive-homogeneous if ρ(af ) = aρ(f ) for any a ≥ 0. • ρ is conic-linear on a cone S if it is linear on conic combinations of elements of S, i.e. ρ(af + bg) = aρ(f ) + bρ(g) for any a, b ≥ 0 and any f, g ∈ S. • ρ is positive if f ≥ 0 =⇒ ρ(f ) ≥ 0. • ρ is monotone if g ≤ f =⇒ ρ(g) ≤ ρ(f ). • ρ is orthogonally additive if f · g = 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). • ρ is real-valued if ρ(f ) ∈ R for any f . • k ρ k= sup{ρ(f ) : k f k≤ 1} and ρ is bounded if k ρ k< ∞. The remaining definitions and facts related to quasi-linear functionals can be found in [5, Section 2]: Definition 2.11. Let X be locally compact. (a) Let f ∈ Cb(X). Define A(f ) to be the smallest closed subalgebra of Cb(X) containing f and 1. Hence, when X is compact, we take f ∈ C(X) and define A(f ) to be the smallest closed subalgebra of C(X) containing f and 1. We call A(f ) the singly generated subalgebra of C(X) generated by f . (b) Let B be a sublagebra of Cb(X). Define B(f ) to be the smallest closed subalgebra of B containing f . We call B(f ) the singly generated subalgebra of B generated by f . We may take, for example, Cc(X), C0(X) as B. Remark 2.12. When X is compact, A(f ) for f ∈ C(X) contains all polynomials of f . It is not hard to show that A(f ) has the form: A(f ) = {φ ◦ f : φ ∈ C(f (X))}. When X is locally compact, B = C0(X) and f ∈ C0(X) (or B = Cc(X) and f ∈ Cc(X)) the singly generated subalgebra has the form: B(f ) = {φ ◦ f : φ(0) = 0, φ ∈ C(f (X))}. Definition 2.13. Let X be locally compact, and let B be a subalgebra of C(X) containing Cc(X). A real-valued map ρ on B is a signed quasi-linear functional on B if (QI1) ρ(af ) = aρ(f ) for a ∈ R; (QI2) for each h ∈ B we have: ρ(f + g) = ρ(f )+ ρ(g) for f, g in the singly generated subalgebra B(h) generated by h. We say that ρ is a quasi-linear functional (or a positive quasi-linear functional) if, in addition, 5 (QI3) f ≥ 0 =⇒ ρ(f ) ≥ 0. When X is compact, we call ρ a quasi-state if ρ(1) = 1. Definition 2.14. We say that a signed quasi-linear functional ρ is compact-finite if ρ(f ) < ∞ for f ∈ Cc(X). 3 Cones generated by functions When X is a compact or locally compact, non-compact space, there is a correspondence between topological measures and quasi-linear functionals. (See [1], [13], [5]). When X is compact, there is also a correspondence between deficient topological measures and r- or l-functionals, see [15] and [10]. When X is locally compact, we shall consider relations between deficient topological measures and various non-linear functionals. We shall start with functionals that are conic-linear on cones generated by functions. Definition 3.1. Let X be compact, f ∈ C(X). Define cones: A+(f ) = {φ ◦ f : φ is non-decreasing }, A−(f ) = {φ ◦ f : φ is non-increasing }. If X is a locally compact, non-compact space, f ∈ C0(X) let A+(f ) = {φ ◦ f : φ is non-decreasing, φ(0) = 0}, A−(f ) = {φ ◦ f : φ is non-increasing, φ(0) = 0}. Remark 3.2. Since f = id ◦ f , for f 6= 0 (f 6= const if X is compact) we have f ∈ A+(f ), f /∈ A−(f ), and −f ∈ A−(f ), −f /∈ A+(f ). Note that 0 ∈ A+(f ), A−(f ). Also, f + ∈ A+(f ), f − ∈ A−(f ), since f + = (id ∨ 0) ◦ f, f − = ((−id) ∨ 0) ◦ f . Obviously, A+(f ), A−(f ) ⊆ B(f ) (respectively, A+(f ), A−(f ) ⊆ A(f ) if X is compact). Definition 3.3. We call a functional ρ on C0(X) with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ a p-conic quasi-linear functional if it is orthogonally additive and monotone on nonnegative functions and conic-linear on A+(f ) for each f ∈ C0(X), i.e. (p1) If f g = 0, f, g ≥ 0 then ρ(f + g) = ρ(f ) + ρ(g). (p2) If 0 ≤ g ≤ f then ρ(g) ≤ ρ(f ). (p3) For each f , if g, h ∈ A+(f ), a, b ≥ 0 then ρ(ag + bh) = aρ(g) + bρ(h). Similarly, a functional ρ on C0(X) with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called an n-conic quasi-linear functional if it is orthogonally additive and monotone on non-positive functions and conic-linear on A−(f ) for each f ∈ C0(X). In other words, (n1) If f g = 0, f, g ≤ 0 then ρ(f + g) = ρ(f ) + ρ(g). (n2) If f ≤ g ≤ 0 then ρ(f ) ≤ ρ(g). 6 (n3) For each f , if g, h ∈ A−(f ), a, b ≥ 0 then ρ(ag + bh) = aρ(g) + bρ(h). Remark 3.4. From [5, Lemma 20(iii), Section 3 and Lemma 44, Section 4] it follows that a positive quasi-linear functional is a p-conic quasi-linear functional and also an n-conic quasi- linear functional. Remark 3.5. Given a functional ρ, consider also the functional π defined by π(f ) = −ρ(−f ) for every f ∈ D(ρ). Then π is an n-conic quasi-linear functional iff ρ is a p-conic quasi-linear functional. This allows us to transfer results for p-conic quasi-linear functionals to n-conic quasi-linear functionals and vice versa. Lemma 3.6. Let ρ be a functional on a locally compact space with ρ(0) < ∞. (I) Suppose a functional ρ is conic-linear on A+(f ) or on A−(f ) for some f . Then ρ(0) = 0. (II) Suppose a functional ρ is conic-linear on A+(h) for each function h ∈ C0(X). Then f g = 0, f ≥ 0, g ≤ 0 implies ρ(f + g) = ρ(f ) + ρ(g). (III) Suppose 0 ≤ g(x) ≤ f (x) ≤ c, f = c on {x : g(x) > 0}. Then f, g ∈ A+(f + g). In particular, if ρ is a p-conic quasi-linear functional then ρ(af + bg) = aρ(f ) + bρ(g) for any a, b ≥ 0. If ρ is a quasi-linear functional, then ρ(af + bg) = aρ(f ) + bρ(g) for any a, b ∈ R. (IV) Suppose g ≥ 0, 0 ≤ h ≤ 1, and h = 1 on {x : g(x) > 0}. If ρ is a p-conic quasi-linear If ρ is a quasi-linear functional then ρ(ag + bh) = aρ(g) + bρ(h) for any a, b ≥ 0. functional, then ρ(ag + bh) = aρ(g) + bρ(h) for any a, b ∈ R. (V) Suppose ρ is a p-conic quasi-linear functional, f, g ∈ Cc(X), f, g ≥ 0. Then ρ(f ) − ρ(g) ≤k ρ kk f − g k. If X is compact we also have: (i) If f ∈ C(X), and a functional ρ is conic-linear on A+(f ) (or on A−(f )), then ρ(f +c) = ρ(f ) + ρ(c). (ii) If ρ is conic-linear on A+(h) for each function h ∈ C(X) (respectively, conic-linear on A−(h) for each function h ∈ C(X)) and ρ(1) ∈ R, then ρ is monotone. Proof. (I) Since 0 ∈ A+(f ), ρ is conic-linear on A+(f ), and ρ(0) < ∞ we have ρ(0) = 0. A similar argument works for A−(f ). (II) Suppose f g = 0, f ≥ 0, g ≤ 0. Taking h = f + g, observe that f = (id ∨ 0) ◦ h, g = (id ∧ 0) ◦ h, so f, g ∈ A+(h). Then ρ(f + g) = ρ(f ) + ρ(g). (III) Note that c ≥ 0 and f = (id ∧ c) ◦ (f + g), g = (0 ∨ (id − c)) ◦ (f + g). (IV) We may assume that g 6= 0. Since 0 ≤ g ≤k g k h ≤k g k, by part (III) ρ(ag + bh) = ρ(ag + b k g k k g k h) = aρ(g) + b k g k ρ(k g k h) = aρ(g) + bρ(h). 7 (V) Suppose f, g ∈ Cc(X), f, g ≥ 0. Choose h ∈ Cc(X) such that h ≥ 0, h = 1 on supp f ∪ supp g. Since f − g ≤k f − g k h, i.e. f ≤ g+ k f − g k h, using Definition 3.3 and part (IV) we have: ρ(f ) ≤ ρ(g+ k f − g k h) = ρ(g)+ k f − g k ρ(h). Similarly, ρ(g) ≤ ρ(f )+ k f − g k ρ(h), so ρ(f ) − ρ(g) ≤k f − g k ρ(h) ≤k f − g k k ρ k . (3.1) (i) Since every constant c ∈ A+(f ) (and also c ∈ A−(f )) we immediately see that ρ(f +c) = ρ(f ) + ρ(c). (ii) Note that ρ(c) ∈ R for any constant c ≥ 0. Let g ≤ f . Choose a constant c > 0 such that 0 ≤ g + c ≤ f + c. We have: ρ(g) + ρ(c) = ρ(g + c) ≤ ρ(f + c) = ρ(f ) + ρ(c), so ρ(g) ≤ ρ(f ). Proposition 3.7. Suppose X is locally compact and ρ is a functional on C0(X) that is positive-homogeneous and monotone on nonnegative functions. If ρ is real-valued, then kρk < ∞. Proof. The statement can be obtained by adapting the argument from Lemma 2.3 in [13] (see Proposition 50 in [5]). 4 d-functionals In this section we shall define several functionals. The domains of these functionals vary, but the most common are C(X), C0(X), Cc(X), C + 0 (X). We shall not specify the domains in the definitions and results that hold for different domains, but shall indicate them later when we use these functionals on specific collections of functions. Definition 4.1. A functional ρ with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called a d-functional if on nonnegative functions it is positive-homogeneous, monotone, and orthogonally additive, i.e. for f, g ∈ D(ρ) (d1) f ≥ 0, a > 0 =⇒ ρ(af ) = aρ(f ), (d2) 0 ≤ g ≤ f =⇒ ρ(g) ≤ ρ(f ), (d3) f · g = 0, f, g ≥ 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). Remark 4.2. It is easy to see that ρ(0) = 0, and so ρ is positive. Definition 4.3. We say that a functional ρ satisfies the c-level condition if f ≤ c, f = c on supp g =⇒ ρ(f + g) = ρ(f ) + ρ(g), where c is a constant, and f, g ∈ D(ρ). If the domain of ρ includes constants, we say that ρ satisfies the constant condition if for any function g ∈ D(ρ) and any constant c ρ(g + c) = ρ(g) + ρ(c). (4.1) 8 Lemma 4.4. Suppose the domain of ρ includes constants. (i) The c-level condition implies the constant condition. (ii) Suppose ρ is a d-functional that satisfies the constant condition, ρ(1) ∈ R, and c ≥ 0. If f ≤ c, f = c on supp g, g ≥ 0, then ρ(g − f ) = ρ(g) + ρ(−f ). (iii) Suppose a functional ρ is conic-linear on A+(h) for each function h (for example, ρ is a p-conic quasi-linear). If f ρ satisfies the constant condition then ρ satisfies the c-level condition for all g ≥ 0. Proof. (i) Take f ≡ c. (ii) Since c − f ≥ 0 and (c − f )g = 0, we have: ρ(c) + ρ(g − f ) = ρ(c + g − f ) = ρ(c − f ) + ρ(g) = ρ(c) + ρ(−f ) + ρ(g), so ρ(g − f ) = ρ(g) + ρ(−f ). (iii) Suppose ρ satisfies the constant condition (4.1). Let f ≤ c, f = c on supp g, g ≥ 0. Let h = f − c. Then h ≤ 0 and g h = 0. By part (II) of Lemma 3.6, ρ(h + g) = ρ(h) + ρ(g). Then ρ(f + g) = ρ(h + c + g) = ρ(h + g) + ρ(c) = ρ(h) + ρ(g) + ρ(c) = ρ(h + c) + ρ(g) = ρ(f ) + ρ(g). Remark 4.5. If ρ is a monotone, positive-homogeneous functional that satisfies the constant condition then for any functions f, g ∈ D(ρ) ρ(f ) − ρ(g) ≤ ρ(k f − g k) =k f − g k ρ(1). One can show this by noticing that f ≤ g+ k f − g k and using an argument similar to the one for part (V) of Lemma 3.6. We modify condition (d3) in Definition 4.1 and obtain the definition of a c-functional: Definition 4.6. A functional ρ with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called a c-functional if for f, g ∈ D(ρ) (c1) f ≥ 0, a > 0 =⇒ ρ(af ) = aρ(f ); (c2) 0 ≤ g ≤ f =⇒ ρ(g) ≤ ρ(f ); (c3) f · g = 0, f, g ≥ 0 or f ≥ 0, g ≤ 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). Note that for a c-functional ρ, ρ(f ) = ρ(f +) + ρ(−f −). Definition 4.7. A real-valued functional ρ is called an s-functional if (s1) f ≥ 0, a ∈ R =⇒ ρ(af ) = aρ(f ); (s2) 0 ≤ g ≤ f =⇒ ρ(g) ≤ ρ(f ); 9 (s3) f · g = 0, f, g ≥ 0 or f ≥ 0, g ≤ 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). Lemma 4.8. Suppose ρ is an s-functional. Then (i) ρ(f ) = ρ(f +) − ρ(f −). (ii) ρ(af ) = aρ(f ) for any f and any a ∈ R. (iii) If g ≤ f then ρ(g) ≤ ρ(f ). (iv) If f · g = 0 then ρ(f + g) = ρ(f ) + ρ(g). Proof. (i) ρ(f ) = ρ(f + − f −) = ρ(f +) + ρ(−f −) = ρ(f +) − ρ(f −). (ii) Easy to see from part (i). (iii) If g ≤ f then g+ ≤ f +, f − ≤ g− and ρ(g) = ρ(g+) − ρ(g−) ≤ ρ(f +) − ρ(f −) = ρ(f ). (iv) If f · g = 0 then (f + g)+ = f + + g+, (f + g)− = f − + g−, and f + · g+ = 0, f − · g− = 0. Then ρ(f + g) = ρ((f + g)+) − ρ((f + g)−) = ρ(f + + g+) − ρ(f − + g−) = ρ(f +) + ρ(g+) − ρ(f −) − ρ(g−) = ρ(f ) + ρ(g). Lemma 4.8 shows that Definition 4.7 is equivalent to the following: Definition 4.9. A functional ρ is called an s-functional if it is homogeneous, monotone, and orthogonally additive, i.e. (sa1) ρ(af ) = aρ(f ) for any a ∈ R; (sa2) g ≤ f =⇒ ρ(g) ≤ ρ(f ); (sa3) f · g = 0 =⇒ ρ(f + g) = ρ(f ) + ρ(g). Remark 4.10. Let X be locally compact, and let ρ be a quasi-linear functional. By [5, Lemma 20(q2), Section 3 and Lemma 44, Section 5], ρ is an s-functional. Definition 4.11. Let QI and L denote, respectively, the families of all quasi-linear and linear functionals. By Φd, Φc, Φs, Φ+, Φ− we denote, respectively, the families of all d-functionals, c-functionals, s-functionals, p-conic quasi-linear functionals, and n-conic quasi-linear func- tionals. Remark 4.12. We have: L ⊆ QI ⊆ Φs ⊆ Φc ⊆ Φd. Proposition 4.13. (i) ρ is a s-functional iff it is a real-valued c-functional with a property that ρ(−g) = −ρ(g) for every g ≤ 0 in the domain of ρ. (ii) If ρ is a d-functional, D(ρ) contains constants, ρ(1) ∈ R, and the constants condition (4.1) is satisfied, then ρ is monotone and positive. 10 Proof. (i) Suppose ρ is a c-functional with the property that ρ(−g) = −ρ(g) for every g ≤ 0 in the domain of ρ. We have: ρ(f ) = ρ(f +) + ρ(−f −) = ρ(f +) − ρ(f −) for any function f in the domain of ρ. Then ρ(−f ) = −ρ(f ) for any function f in the domain of ρ, and condition (sa1) of Definition 4.7 follows. Thus, ρ is an s-functional. (ii) Let g ≤ f . It is enough to assume that 0 ≤ g ≤ f , for choosing a positive constant k such that g +k ≥ 0 we see that ρ(g) ≤ ρ(f ) iff ρ(g)+ρ(k) ≤ ρ(f )+ρ(k) iff ρ(g +k) ≤ ρ(f +k). But ρ is a d-functional, so ρ(g) ≤ ρ(f ) for 0 ≤ g ≤ f . Since ρ(0) = 0, monotonicity of ρ implies positivity. Now we will introduce the closely related r- and l- functionals. Definition 4.14. A functional ρ with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called an r-functional if for f, g ∈ D(ρ) (r1) a > 0 =⇒ ρ(af ) = aρ(f ); (r2) 0 ≤ g ≤ f =⇒ ρ(g) ≤ ρ(f ); (r3) If f · g = 0 where f, g ≥ 0 or f ≥ 0, g ≤ 0 then ρ(f + g) = ρ(f ) + ρ(g). If D(ρ) contains constants then we also require for f ∈ D(ρ) and a constant c ρ(f + c) = ρ(f ) + ρ(c). Definition 4.15. A functional ρ with values in [−∞, ∞] (assuming at most one of ∞, −∞) and ρ(0) < ∞ is called an l-functional if (l1) a > 0 =⇒ ρ(af ) = aρ(f ); (l2) g ≤ f ≤ 0 =⇒ ρ(g) ≤ ρ(f ); (l3) If f · g = 0 where f, g ≤ 0 or f ≥ 0, g ≤ 0 then ρ(f + g) = ρ(f ) + ρ(g). If D(ρ) contains constants then we also require for f ∈ D(ρ) and a constant c ρ(f + c) = ρ(f ) + ρ(c). Definition 4.16. By Φr, Φl we denote, respectively, the families of all r-functionals, and l-functionals on X. Remark 4.17. Here are a few easy observations. (i) Suppose ρ, R are r-functionals with the same domain that contain constants and ρ(1) ∈ R. Then ρ = R iff ρ = R on nonnegative functions. Indeed, for an arbitrary function f choose a constant c ≥ 0 such that f + c ≥ 0 and see that ρ(f ) + ρ(c) = ρ(f + c) = R(f + c) = R(f ) + R(c) = R(f ) + ρ(c), i.e. ρ(f ) = R(f ). (ii) We have: Φs ⊆ Φr ⊆ Φc ⊆ Φd and from Definition 4.9 Φs ⊆ Φl. (iii) Given a functional ρ, consider also the functional π defined by π(f ) = −ρ(−f ) for every f ∈ D(ρ). Then π is an l-functional iff ρ is an r-functional. 11 Lemma 4.18. Each p-conic quasi-linear functional is an r-functional. Each n-conic quasi- linear functional is an l-functional. So Φ+ ⊆ Φr and Φ− ⊆ Φl. Proof. Using part (II) and part (i) of Lemma 3.6 we see that a p-conic quasi-linear functional is an r-functional. The second statement follows from Remark 3.5 and part (iii) of Remark 4.17. Lemma 4.19. Let X be compact and ρ be a functional on C(X). (I) If ρ is an r-functional with ρ(1) ∈ R or an l-functional with ρ(−1) ∈ R then ρ is monotone. (II) If ρ is an r-functional with ρ(1) ∈ R then ρ(f ) − ρ(g) ≤k ρ k k f − g k= ρ(1) k f − g k . (III) If ρ is an r-functional with ρ(−1) ∈ R then ρ satisfies the c-level condition for any g ≥ 0. (IV) If ρ is an r-functional, f · g = 0, f ≥ 0 then ρ(f + g) = ρ(f ) + ρ(g). Similarly, if ρ is an l-functional, f · g = 0, f ≤ 0 then ρ(f + g) = ρ(f ) + ρ(g). Proof. (I) Suppose g ≤ f . Choose a constant c > 0 such that 0 ≤ g + c ≤ f + c. Then ρ(g)+ρ(c) = ρ(g+c) ≤ ρ(f +c) = ρ(f )+ρ(c), which gives ρ(g) ≤ ρ(f ). The monotonicity of an l-functional can be proved similarly. (II) Use part (I) and Remark 4.5. (III) Assume that f ≤ c, f = c on supp g, g ≥ 0. Then (f − c)g = 0, f − c ≤ 0, and ρ(−c) + ρ(f + g) = ρ(f + g − c) = ρ(f − c) + ρ(g) = ρ(f ) + ρ(−c) + ρ(g). Thus, ρ(f + g) = ρ(f ) + ρ(g). (IV) Let ρ be an r-functional, f · g = 0, f ≥ 0. Then (f + g+)(−g−) = 0, f · g+ = 0, and by Definition 4.14 ρ(f + g) = ρ(f + g+ − g−) = ρ(f + g+) + ρ(−g−) = ρ(f ) + ρ(g+) + ρ(−g−) = ρ(f ) + ρ(g). The proof for an l-functional is similar. Remark 4.20. Part (IV) of Lemma 4.19 and part (iii) of Remark 4.17 were first observed for a compact space in [15, Propositions 17 and 18]. Definition 4.21. Let ρ be a d-functional. Let g ≥ 0. Define ρg(f ) = ρ(f g) for f ∈ D(ρ). Lemma 4.22. If ρ is a d-,c-, r-, l-,s- functional or a linear functional, then so is ρg. Proof. Easy to check. 12 5 Deficient topological measures from d-functionals Definition 5.1. Let X be locally compact, and let ρ be a d-functional with C + c (X) ⊆ D(ρ) ⊆ Cb(X). Define a set function µρ : O(X) ∪ C (X) → [0, ∞] as follows: for an open set U ⊆ X let µρ(U ) = sup{ρ(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U }, and for a closed set F ⊆ X let µρ(F ) = inf{µρ(U ) : F ⊆ U, U ∈ O(X)}. Note that Definition 5.1 is consistent for clopen sets. Lemma 5.2. For the set function µρ from Definition 5.1 the following holds: y1. µρ is nonnegative. y2. µρ is monotone. y3. Given an open set U , for any compact K ⊆ U µρ(U ) = sup{ρ(g) : 1K ≤ g ≤ 1, g ∈ Cc(X), supp g ⊆ U }. y4. For any K ∈ K (X) y5. For any K ∈ K (X) µρ(K) = inf{ρ(g) : g ∈ Cc(X), g ≥ 1K }. µρ(K) = inf{ρ(g) : g ∈ Cc(X), 1K ≤ g ≤ 1}. y6. Given K ∈ K (X), for any open U such that K ⊆ U µρ(K) = inf{µρ(V ) : V ∈ O(X), K ⊆ V ⊆ V ⊆ U }. y7. For any U ∈ O(X) µρ(U ) = sup{µρ(K) : K ∈ K (X), K ⊆ U }. y8. For any disjoint compact sets K and C µρ(K ⊔ C) = µρ(K) + µρ(C). y9. Suppose X is compact, ρ(1) ∈ R, and ρ is an s-functional satisfying the constant con- dition (4.1). If K ⊆ U, K ∈ K (X), U ∈ O(X) then µρ(U ) = µρ(K) + µρ(U \ K). 13 Proof. For part y1, µρ is nonnegative since ρ is a positive functional by Remark 4.2. Part y2 is easy to see. Proofs for parts y3 - y8 follow proofs of the corresponding parts of [5, Lemma 35, Sect. 4]. We shall show part y9. Let K ⊆ U, K ∈ K (X), U ∈ O(X). First we shall show that µρ(U \ K) + µρ(K) ≥ µρ(U ). (5.1) If µρ(K) = ∞, the inequality (5.1) trivially holds, so we assume that µρ(K) < ∞. By Lemma 2.7 let V ∈ O(X) with compact closure be such that K ⊆ V ⊆ V ⊆ U. For ǫ > 0 choose W1 ∈ O(X) such that K ⊆ W1 ⊆ V and µρ(W1) < µρ(K) + ǫ. Also, there exists W ∈ O(X) with compact closure such that K ⊆ W ⊆ W ⊆ W1 ⊆ V ⊆ V ⊆ U. Choose an Urysohn function g ∈ Cc(X) such that 1W ≤ g ≤ 1, supp g ⊆ W1. Then ρ(g) ≤ µρ(W1) < µρ(K) + ǫ. First assume that µρ(U ) < ∞. By part y3 choose f ∈ Cc(X) such that 1V ≤ f ≤ 1, supp f ⊆ U , and ρ(f ) > µρ(U ) − ǫ. Note that 0 ≤ f − g ≤ 1, and, since f − g = 0 on W , we have supp (f − g) ⊆ U \ K. Since f = 1 on supp g and ρ is an s-functional, by Remark 4.12 and part (ii) of Lemma 4.4 ρ(g − f ) = ρ(g) + ρ(−f ) = ρ(g) − ρ(f ), so Then we have: ρ(f − g) = ρ(f ) − ρ(g). (5.2) µρ(U \ K) ≥ ρ(f − g) = ρ(f ) − ρ(g) ≥ µρ(U ) − ǫ − µρ(K) − ǫ, which gives us inequality (5.1). If µρ(U ) = ∞, use instead of f functions fn with 1V ≤ fn ≤ 1, supp fn ⊆ U, ρ(fn) ≥ n in the above argument to show that µρ(U \ K) = ∞. Then inequality (5.1) holds. Now we would like to show that µρ(U ) ≥ µρ(U \ K) + µρ(K). (5.3) By monotonicity of µρ it is enough assume that µρ(U \ K), µρ(K) < ∞. Given ǫ > 0, choose g ∈ Cc(X), 0 ≤ g ≤ 1 such that C = supp g ⊆ U \ K and ρ(g) > µρ(U \ K) − ǫ. 14 Note that K ⊆ U \ C. If µρ(U \ C) = ∞, then µρ(U ) = ∞, so (5.3) holds. So assume that µρ(U \ C) < ∞. By part y3 choose f ∈ Cc(X) such that 1K ≤ f ≤ 1, supp f ⊆ U \ C, and ρ(f ) > µρ(U \ C) − ǫ. Then ρ(f ) > µρ(U \ C) − ǫ ≥ µρ(K) − ǫ. Since f g = 0, f, g ≥ 0, we have ρ(f +g) = ρ(f )+ρ(g). Since f +g ∈ Cc(X) with supp (f +g) ⊆ U , we obtain: µρ(U ) ≥ ρ(f + g) = ρ(f ) + ρ(g) ≥ µρ(K) + µρ(U \ K) − 2ǫ. Therefore, µρ(U ) ≥ µρ(U \ K) + µρ(K). Remark 5.3. In the proof of part y9 the only place where we need the fact that X is compact and D(ρ) contains constants is when we use part (ii) of Lemma 4.4 to obtain formula (5.2) in order to get inequality (5.1). If ρ is a quasi-linear functional on a locally compact space then formula (5.2) holds by part (III) of Lemma 3.6, and we again obtain inequality (5.1). Our means of obtaining inequality (5.1) resembles one from [13, Theorem 3.9]. Theorem 5.4. Suppose X is locally compact, ρ is a d-functional with Cc(X) ⊆ D(ρ) ⊆ Cb(X), and µρ defined in Definition 5.1. Then (i) µρ is a deficient topological measure. (ii) If ρ is real-valued on Cc(X), then µρ is compact-finite. (iii) If ρ is bounded, then µρ is finite. (iv) If X is compact and D(ρ) = C(X) then µρ(X) = ρ(1). (v) If the domain of ρ includes constants, ρ(1) ∈ R, and ρ is an s-functional satisfying constant condition (4.1) then µρ is a topological measure. (vi) If ρ is a quasi-linear functional then µρ is a topological measure. Proof. (i) Note that since ρ is not identically ∞, then neither is µρ. By part y1 of Lemma 5.2 µρ is nonnegative. Part y8 of Lemma 5.2 gives (DTM1) of Definition 2.2. Defini- tion 5.1 and part y7 of Lemma 5.2 give regularity conditions (DTM2) and (DTM3) of Definition 2.2. Thus, µρ is a deficient topological measure. (ii) Follows from part y4 of Lemma 5.2. (iii) Evident from Definition 5.1. (iv) See Definition 5.1. (v) Follows from part y9 (or just inequality (5.1)) of Lemma 5.2 and Theorem 2.6. (vi) Follows form Remark 5.3 and Theorem 2.6. 15 6 Left and right measures Given a deficient topological measure and a bounded continuous function we may consider four distribution functions. Definition 6.1. Let µ be a finite deficient topological measure on a locally compact space X. Let f ∈ Cb(X). Define the following nonnegative functions on R: L1(t) = L1,µ,f (t) = µ(f −1((−∞, t))), L2(t) = L2,µ,f (t) = µ(f −1((−∞, t])), R1(t) = R1,µ,f (t) = µ(f −1((t, ∞))), R2(t) = R2,µ,f (t) = µ(f −1([t, ∞))). Remark 6.2. For particular µ and f to simplify notations we use L1, L2, R1, R2. When we need to emphasize the dependence on µ and f , we use notations L1,µ,f , R1,µ,f and so on. When we need to use, say, L1 as a function of f we denote it by L1,f . Lemma 6.3. Let µ be a finite deficient topological measure on a locally compact space X, f ∈ Cb(X). Let nonnegative real-valued functions L1, L2, R1, R2 be as in Definition 6.1. Then I. Functions L1, L2 are non-decreasing; R1, R2 are non-increasing. If f (X) ⊆ [a, b] then L1(a) = L2(a−) = 0; L1(b+) = L2(b+) = L2(b) = µ(X), R1(a−) = R2(a−) = R2(a) = µ(X); R1(b) = R2(b+) = 0. II. L1 is left-continuous, R1 is right-continuous. III. L1(t−) = L2(t−) = L1(t) for any t. If L2 is left-continuous at t (in particular, continu- ous at t) then L1(t) = L2(t). Similarly, R2(t+) = R1(t+) = R1(t) for any t, and if R2 is right-continuous at t then R1(t) = R2(t). In particular, the set of t where L1(t) 6= L2(t) and the set of t where R1(t) 6= R2(t) are, at most, countable sets. IV. L1(t) + R1(t) ≤ µ(X) for every t. V. If X is compact, the function L2 is right-continuous, and R2 is left-continuous. If X is locally compact, f ∈ C0(X), then R2(t) is left-continuous at a and any t > 0, and L2(t) is right-continuous at b and any t < 0. In particular, R2 is left-continuous at all t except, possibly, t ∈ E for some countable set E ⊆ (−∞, 0] \ {a} and L2(t) is right-continuous at all t except, possibly, t ∈ E1 for some countable set E1 ⊆ [0, ∞) \ {b}. Proof. I. Easy to see. II. The sets Us = f −1((−∞, s)) are open, Us ր Ut as s → t−, so by Lemma 2.8 L1 is left-continuous. The argument for R1 is similar. III. Let s < t. Then f −1((−∞, s)) ⊆ f −1((−∞, s]) ⊆ f −1((−∞, t)), and so L1(t−) ≤ L2(t−) ≤ L1(t) ≤ L2(t). By left-continuity of L1 we have L1(t−) = L2(t−) = L1(t); if L2 is left-continuous at t then L1(t) = L2(t). Similarly for R1 and R2. IV. The sets f −1((−∞, t)) and f −1((t, ∞)) are disjoint open sets, so from superadditivity of µ we see that L1(t) + R1(t) ≤ µ(X) for every t. 16 V. If X is compact, the sets Ca = f −1([a, ∞) are compact. From Lemma 2.8 it follows that R2 is left-continuous. If X is locally compact and f ∈ C0(X) then the sets Ka = f −1([a, ∞)), a > 0 are compact. From Lemma 2.8 it follows that R2 is left-continuous at any t > 0. The assertions about L2 are proved similarly. Remark 6.4. Let µ be a finite deficient topological measure on a locally compact space X. Let f ∈ Cb(X) with f (X) ⊆ [a, b]. By Theorem 2.9 and part I of Lemma 6.3 the Riemann- Stieltjes integral Z b a id dL1 = −Z b a L1(t)dt + L1(b+)b = −Z b a L1(t)dt + bµ(X). Let l be the Lebesque-Stieltjes measure associated with L1, so l is a regular Borel measure on R. By part III of Lemma 6.3 we see that Z b a id dl =Z b a id dL1 = −Z b a L1(t)dt + bµ(X) = −Z b a L2(t)dt + bµ(X). Let r be the Lebesque-Stieltjes measure associated with −R1, a regular Borel measure on R. We have: Z b a id dr =Z b a id d(−R1) =Z b a R1(t)dt + aµ(X) =Z b a R2(t)dt + aµ(X). Definition 6.5. We call l the left measure and r the right measure. When the right and left measures are equal, we set m = r = l. Remark 6.6. The measures r and l arise from functions R1 = R1,µ,f and L1 = L1,µ,f . We use notations R1,f , rf , lf when we need to emphasize the dependence of R1 and measures r, l on the function f . If we want to use measures r and l as functions of f and µ, we write rf,µ, lf,µ. When µ is a topological measure, measure m is equal to µf in [13] and mf in [5]. See [5, Remark 28, Sect. 3]. Theorem 6.7. Let µ be a finite deficient topological measure on a locally compact space X, and let f ∈ C0(X). (I) There are regular Borel measures r and l on R such that supp r ⊆ f (X), supp l ⊆ f (X), l(R) = µ(X), r(R) = µ(X), r((t, ∞)) = µ(f −1((t, ∞))) for all t, r([t, ∞)) = µ(f −1([t, ∞))) for all t /∈ E, where E ⊆ (−∞, 0] is a countable set from part V of Lemma 6.3 l((−∞, t)) = µ(f −1((−∞, t))) for all t, l((−∞, t]) = µ(f −1((−∞, t])) for all t /∈ E1, where E1 ⊆ [0, ∞) is a countable set from part V of Lemma 6.3. 17 (II) For any open or closed set A ⊆ R µ(f −1(A)) ≤ l(A), µ(f −1(A)) ≤ r(A). Proof. Let f (X) = [a, b]. (I) By Lemma 6.3 L1 is left-continuous, so for every t l((−∞, t)) = L1(t) = µ(f −1((−∞, t))). Next, using Lemma 2.8 l(R) = lim n→∞ r((−∞, n)) = lim n→∞ µ(f −1((−∞, n))) = µ(f −1(R)) = µ(X). If t /∈ E1, then L2 is right-continuous at t. Since L2 = L1 outside of a countable set, l((−∞, t]) = lims→t+ l((−∞, s)) = lims→t+ L1(s) = lims→t+ L2(s) = L2(t) = µ(f −1((−∞, t])). Since L1 is constant on (−∞, a) and on (b, ∞), we see that l((−∞, a)) = l((b, ∞)) = 0. It follows that supp l ⊆ [a, b] = f (X). The statements for r can be proved similarly. (II) Let (a, b) ⊆ R, b /∈ E. By the superadditivity of µ and part (I) µ(f −1((a, b)) ≤ µ(f −1((a, ∞))) − µ(f −1([b, ∞))) = r((a, ∞)) − r([b, ∞)) = r((a, b)). For (a, b) with b ∈ E, choose bn /∈ E such that (a, bn) ր (a, b). Since by Lemma 2.8 µ ◦ f −1 and r are both τ -smooth on open sets, we have µ(f −1((a, b))) ≤ r((a, b)) for any (a, b). We see that µ(f −1(J)) ≤ r(J) for any finite or infinite open interval J. Then the same inequality holds for any open set W ⊆ R. Now let C ⊆ R be closed. Then µ(f −1(C)) ≤ inf{µ(f −1(W ) : C ⊆ W, W ∈ O(X)} ≤ inf{r(W ) : C ⊆ W, W ∈ O(X)} = r(C). The statements for the left measure l can be proved in a similar way. Theorem 6.8. Let µ be a finite deficient topological measure on a locally compact space. Then r = l iff L1(t) + R1(t) = µ(X) for a.e. t with respect to the Lebesque measure. Proof. Using Remark 6.4 and part IV of Lemma 6.3 we may note that r = l =⇒Z b ⇐⇒ Z b a a id d(−R1) id dL1 =Z b R1dx + aµ(X) = −Z b a a 18 L1dx + bµ(X) ⇐⇒ Z b a (L1 + R1)dx = (b − a)µ(X) ⇐⇒ L1 + R1 = µ(X) a.e., where a.e. is with respect to the Lebesque measure λ. Conversely, let L1(t) + R1(t) = µ(X) for t /∈ D, where λ(D) = 0. We may assume that D contains sets E, E1 from part V of Lemma 6.3 and all points where L1 6= L2, R1 6= R2. If [a, b] ⊆ R and a, b /∈ E then by part (I) of Theorem 6.7 we have: l([a, b]) = l((−∞, b]) − l((−∞, a)) = L2(b) − L1(a) = µ(X) − R1(b) − µ(X) + R2(a) = r([a, ∞)) − r((b, ∞)) = r([a, b]). An arbitrary interval (a, b) can be written asS∞ by inclusion, and an, bn /∈ E. It follows that measures l = r on R. n=1[an, bn], where intervals [an, bn] are ordered Theorem 6.9. Let µ be a finite topological measure on a locally compact space X. Let f ∈ C0(X). Then for the right and left measures r, l we have r = l. Proof. Let f (X) ⊆ [a, b]. By Theorem 6.8 it is enough to show that there is a countable set D such that L1(t) + R1(t) = µ(X) for all t ∈ R \ D. Let D be the countable (by part III of Lemma 6.3) set consisting of 0 and all points where L1 6= L2, R1 6= R2. If t > 0 then f −1([t, ∞)) is compact, and it follows from (TM1) of Definition 2.4 that L1(t)+R2(t) = µ(X). R1(t) = R2(t) for t ∈ (0, ∞) \ D, so L1(t) + R1(t) = L1(t) + R2(t) = µ(X). Similarly, for all t ∈ (−∞, 0) \ D we have L1(t) + R1(t) = L2(t) + R1(t) = µ(X). Theorem 6.10. Let µ be a finite topological measure on a locally compact space X, and let regular Borel measure m = r = l. (I) If f ∈ C0(X), then m(A) = µ(f −1(A)) for any open set A ⊆ R and any closed set A ⊆ R \ {0}. (II) If X is compact, f ∈ C(X) then m(A) = µ(f −1(A)) for any open or closed set A ⊆ R. Proof. (I) First let (a, b) ∈ R, b 6= 0. Note that in f −1((a, ∞)) = f −1((a, b)) ⊔ f −1({b}) ⊔ f −1((b, ∞)), all the sets are open except for the middle set on the right hand side, which is compact since f ∈ C0(X). Applying µ we obtain R1(a) = µ(f −1((a, b))) + µ(f −1({b})) + µ(f −1((b, ∞))) (6.1) Since f −1((b, ∞))⊔f −1({b}) ⊆ f −1((t, ∞)) for any t < b, by superadditivity (see Lemma 2.8) we have: µ(f −1({ b})) + µ(f −1((b, ∞))) ≤ µ(f −1((t, ∞))) = R1(t). Thus, from (6.1) we see that R1(a) ≤ µ(f −1((a, b))) + R1(t). As t → b− we have: m((a, b)) = R1(a) − R1(b−) ≤ µ(f −1((a, b))). 19 Together with part (II) of Theorem 6.7 we obtain m((a, b)) = µ(f −1((a, b))) for any interval (a, b), b 6= 0. An interval (a, 0) = ∪∞ n ). Since both µ and m are τ − smooth and additive on open sets (see Lemma 2.8), the result holds for any finite open interval in R, and then for any open set in R. Below in (III) we shall prove that µ(f −1(C)) = m(C) for closed sets in R \ {0}. n=1(a, − 1 (II) The set f −1([d, ∞)) is compact for every d, and the argument as in part (I) shows that m(W ) = µ(f −1(W )) for every open set W ⊆ R. Now let C ⊆ R be closed. Choose W ∈ O(X) such that C ⊆ W . Since f −1(C) is compact and µ is a topological measure, m(W ) − µ(f −1(C)) = µ(f −1(W )) − µ(f −1(C)) = µ(f −1(W \ C)) = m(W \ C). Thus µ(f −1(C)) = m(C). (III) Now we shall finish the proof of part (I). Let C ∈ R \ {0}. Set C1 = C ∩ (0, ∞) and C2 = C ∩ (−∞, 0). We have C = C1 ⊔ C2, and b = inf C1 > 0. Since f −1(C1) ⊆ f −1([b, ∞)), the set f −1(C1) is compact. An argument similar to the one in part (II) shows that µ(f −1(C1)) = m(C1). Similarly, µ(f −1(C2)) = m(C2), and so by finite additivity of µ and m on compact sets µ(f −1(C)) = m(C). Lemma 6.11. Let µ be a finite deficient topological measure on a locally compact space X. I. If µ is a simple deficient topological measure, then measures r and l are point masses, l = δa, r = δb, b ≤ a, where a = inf{t : L1(t) = 1} = sup{s : L1(s) = 0}, b = inf{t : R1(t) = 0} = sup{s : R1(s) = 1}. II. If X is compact and f = c is a constant function, then the measure m = µ(X)δc, where δc is a point mass at c. Proof. I. Since µ is simple, the non-decreasing function L1 assumes only two values, and has single discontinuity at a = inf{t : L1(t) = 1} = sup{s : L1(s) = 0}. Since l({a}) = L1(a+) − L1(a−) = 1, we see that l = δa. Similarly, r = δb, where b = inf{t : R1(t) = 0} = sup{s : R1(s) = 1}. If a < b then L1(t) + R1(t) = 2 on interval (a, b), which contradicts part IV of Lemma 6.3. Thus, b ≤ a. II. We have R1(t) = µ(X) for every t < c, and R1(t) = 0 for every t ≥ c. Then m({c}) = r({c}) = R1(c−) − R1(c+) = µ(X). Example 1. Let X = R, D = [0, 1] and µ be a simple deficient topological measure as in [4, Example 48, Sect. 6], i.e. µ(A) = 1 if D ⊆ A and µ(A) = 0 otherwise, where A ∈ O(X) ∪ K (X). Consider the following f ∈ C0(X) : f (0) = 1, f (t) = 0 for t ∈ (−∞, −1] ∪ [1, ∞), and 20 f is linear on [−1, 0] and [0, 1]. Note that D ⊆ f −1((−∞, t)) iff t > 1, and D ⊆ f −1((t, ∞)) iff t < 0. Thus, and L1(t) = ( 1 0 if t > 1 if t ≤ 1 R1(t) = ( 1 0 if t < 0 if t ≥ 0 So l = δ1 and r = δ0. Example 2. Let X = R, the family E = {D = [1, 2]}, λ0 = δ3/2. Let µ = λ+ as in [4, Example 49, Sect. 6]. Consider the following f ∈ C0(X) : f (0) = 2, f (t) = 1 for t ∈ [1, 2], f (t) = 0 for t ∈ (−∞, −1] ∪ [3, ∞), and f is linear on [−1, 0], [0, 1], and [2, 3]. Note that D ⊆ f −1((−∞, t)) iff t > 1, and D ⊆ f −1((t, ∞)) iff t < 1. It follows that and L1(t) = λ+(f −1((−∞, t))) = ( 1 0 if t > 1 if t ≤ 1 R1(t) = λ+(f −1((t, ∞))) = ( 1 0 if t < 1 if t ≥ 1 Thus, for measures l, r we have l = r = δ1. Remark 6.12. In Theorem 6.7 it is stated that supp l, supp r ⊆ f (X). In Example 1 and Example 2 supp l and supp r are properly contained in f (X). On the other hand, from part II of Lemma 6.11 we see that it is also possible to have supp l = supp r = f (X). Remark 6.13. From part IV of Lemma 6.3 we know that L1(t) + R1(t) ≤ µ(X). Although L1(t) + R1(t) = µ(X) a.e. when µ is a finite topological measure (see Theorems 6.8 and 6.9), for deficient topological measures we may have both situations: in Example 2 L1(t) + R1(t) = µ(X) a.e., but in Example 1 we have L1(t) + R1(t) = 0 < µ(X) for t ∈ (0, 1). In Lemma 6.11 we have b ≤ a. Example 1 and Example 2 show that both situations when b < a and b = a are possible. These examples also show that when µ is a deficient topological measure, we can have both situations for measures l and r induced by µ and a given function f : when l = r and when l 6= r. 7 Functionals from deficient topological measures When µ is a finite deficient topological measure (not a topological measure) the measures r, l are not equal in general, and we consider two different integrals: and a Z b Z b a id dl id dr. 21 Definition 7.1. Let µ be a finite deficient topological measure on a locally compact space X, and let measures r = rf,µ, l = lf,µ, m = mf,µ be as in Definition 6.5, Remark 6.4, and Remark 6.6. Define the following functionals on Cb(X): and R(f ) = Rµ(f ) =ZR L(f ) = Lµ(f ) =ZR ρ(f ) = ρµ(f ) =ZR id dr, id dl, id dm. Remark 7.2. By Theorem 6.7 supp r, supp l ⊆ f (X), so for any [a, b] containing f (X) R(f ) =ZR id dr =Z[a,b] id dr, L(f ) =Z[a,b] id dl, ρ(f ) =Z[a,b] id dm. With functions L1, L2, R1, R2 as in Definition 6.1 by Remark 6.4 we have: L2(t)dt + bµ(X). (7.1) L(f ) =ZR R(f ) =ZR id dl =Z b id dr =Z b id dl = −Z b id dr =Z b a a L1(t)dt + bµ(X) = −Z b R1(t)dt + aµ(X) =Z b a a a a R2(t)dt + aµ(X). (7.2) (7.3) (7.4) (7.5) If f = c is a constant function then If f is nonnegative with f (X) ⊆ [0, b] we have: R(f ) = L(f ) = cµ(X). R(f ) =Z b 0 id dr =Z b 0 R1(t)dt =Z b 0 R2(t)dt. When r = l (in particular, when µ is a topological measure) and f (X) ⊆ [0, b] we have ρ(f ) =Z b 0 id dm =Z b 0 R1(t)dt =Z b 0 R2(t)dt. Similarly, if f is nonpositive with f (X) ⊆ [a, 0] we have: L(f ) =Z 0 a id dl = −Z 0 a L1(t)dt = −Z 0 a L2(t)dt. When r = l (in particular, when µ is a topological measure) and f (X) ⊆ [a, 0] we have ρ(f ) =Z 0 a id dm = −Z 0 a L1(t)dt = −Z 0 a L2(t)dt. 22 Remark 7.3. Note that when µ is a topological measure, we obtain familiar formulas. See, for example, [13, Proposition 3.7] and [5, Remark 43, Section 5]. These results were, in turn, generalizations of results first obtained by J. F. Aarnes for compact spaces in [1]. For example, when X is compact and µ(X) = 1, formula (7.2) gives [1, formula (3.3)]. Remark 7.4. We have the connection between R and L (which is the same as noted in [15, p. 739]). We use notations as in Remark 6.6. Observe that R1,f (−t) = L1,−f (t). Thus, l−f = rf ◦ T −1, where T (t) = −t for t ∈ R. Then RR id dl−f = −RR id drf , i.e. L(−f ) = −R(f ). (7.6) We may prove results for R and obtain similar results for L by analogy (as we did, for example, in Theorem 6.7) or using relation (7.6). Definition 7.5. Let R be the functional as in Definition 7.1. We call the functional R a quasi-integral (with respect to a deficient topological measure µ) and write: ZX f dµ = R(f ) = Rµ(f ) =ZR id dr. Remark 7.6. If µ is a topological measure on X, by Definitions 7.1 and 6.5 we obtain exactly the quasi-integral in [5, Definition 27, Section 3]. Lemma 7.7. Let L, R be functionals as in Definition 7.1. (i) R is orthogonally additive on nonnegative functions, and L is orthogonally additive on nonpositive functions. (ii) R(0) = 0 and L(0) = 0. (iii) L, R are positive-homogeneous functionals. (iv) L, R are monotone. In particular, L, R are positive. Proof. We use notations as indicated in Remark 6.6. (i) Let f, g ≥ 0 and f g = 0. Say, f (X), g(X) ⊆ [0, b]. For any t > 0 observe that (f + g)−1((t, ∞)) = f −1((t, ∞)) ⊔ g−1((t, ∞)), so by additivity of a deficient topological measure on disjoint open sets we immediately obtain R1,f +g(t) = R1,f (t) + R1,g(t). Since (f + g)(X) ⊆ [0, b], from (7.4) we have R(f + g) = R(f ) + R(g). Thus, R is orthogonally additive on nonnegative functions. Then orthogonal additivity of L on nonpositive functions follows from (7.6). (ii) Follows from part (i) or from (7.4) and (7.5). (iii) If c = 0 then from (ii) we see that R(cf ) = L(cf ) = 0. Let c > 0. Since L1,cf (t) = L1,f (t/c), from (7.1) we see that L(cf ) = cL(f ). One can show that R is also positive- homogeneous in a similar way using (7.2) or using positive-homogenuity of L together with formula (7.6). (iv) Suppose that f ≤ g. Choose an interval [a, b] which contains both f (X) and g(X). Since L1,f ≥ L1,g and R1,f ≤ R1,g, from (7.1) and (7.2) we see that L(f ) ≤ L(g) and R(f ) ≤ R(g). 23 Lemma 7.8. If h = φ ◦ f ∈ A+(f ) and [a, b] is any interval containing f (X) then Similarly, if h = φ ◦ f ∈ A−(f ) then R(h) =Z[a,b] φ drf . L(h) =Z[a,b] φ dlf . Proof. Let rf , rh be the right measures for functions f and h as in Theorem 6.7, rf is supported on f (X) ⊆ [a, b]. Since φ is nondecreasing, for any interval (t, ∞) we have φ−1((t, ∞)) = (s, ∞), where s = min{y ∈ [a, b] : φ(y) = t}. (This is similar to [15, Proposition 13(1)].) Then rh((t, ∞)) = µ(h−1((t, ∞))) = µ(f −1(φ−1((t, ∞)))) = µ(f −1((s, ∞))) = rf ((s, ∞)) = (rf ◦ φ−1)((t, ∞)). Thus, rh = rf ◦ φ−1 are equal as measures. Using formula (7.2) we have: R(h) =ZR id drh =Z[a,b] φ drf . The formula L(h) =R[a,b] φ dlf can be proved in a similar way. Lemma 7.9. The functional R is conic-linear on each cone A+(f ), and the functional L is conic-linear on each cone A−(f ). Proof. Suppose h, g ∈ A+(f ), h = φ ◦ f, g = ψ ◦ f, f (X) ⊆ [a, b]. Applying Lemma 7.8 we have: R(h + g) = R(φ ◦ f + ψ ◦ f ) = R((φ + ψ) ◦ f ) =Z[a,b] (φ + ψ) drf =Z[a,b] φ drf +Z[a,b] ψ drf = R(h) + R(g). Since by Lemma 7.7 R is also positive-homogeneous, we see that R is conic-linear on A+(f ) for each f . The statements for L can be proved similarly. Theorem 7.10. (i) The functional R is a p-conic quasi-linear functional, and the func- tional L is an n-conic quasi-linear functional. (ii) The functional R is an r-functional , and L is an l-functional. Proof. (i) Follows from Lemma 7.7 and Lemma 7.9. (ii) Apply Lemma 4.18. 24 Remark 7.11. When µ is a topological measure, the functional ρ = ρµ is a quasi-linear functional (see [5, Theorem 30, Section 3]), so ρ is linear on each singly generated subalgebra. Theorem 7.10 gives an analog of this for the case when µ is a deficient topological measure: if µ is a deficient topological measure, then the functional R obtained from µ is p-conic linear, so, in particular, it is conic-linear on the cone A+(f ) for each f . The next lemma shows properties that relate R and µ. Lemma 7.12. Let µ be a finite deficient topological measure, and R, L be functionals on C0(X)obtained from µ as in Definition 7.1. z1. If U ∈ O(X) and f ∈ C(X) is such that supp f ⊆ U, 0 ≤ f ≤ 1 then R(f ) ≤ µ(U ). z2. If C ∈ C (X) and f is such that 0 ≤ f ≤ 1, f = 1 on C, then R(f ) ≥ µ(C). z3. For any f ∈ C0(X) Similarly, µ(X) · inf x∈X f (x) ≤ R(f ) ≤ µ(X) · sup x∈X f (x). µ(X) · inf x∈X f (x) ≤ L(f ) ≤ µ(X) · sup x∈X f (x). Hence, k R k, k L k≤ µ(X). z4. If U ∈ O(X) then µ(U ) = sup{R(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U }. z5. If K ∈ K (X) then µ(K) = inf{R(g) : g ∈ Cc(X), g ≥ 1K }. z6. If K ∈ K (X) then µ(K) = inf{R(g) : g ∈ Cc(X), g ≥ 1K , 0 ≤ g ≤ 1} = inf{R(g) : g ∈ C0(X), g ≥ 1K } = inf{R(g) : g ∈ C0(X), g ≥ 1K , 0 ≤ g ≤ 1}. z7. k R k= µ(X) =k L k, so R(f ) ≤k R kk f k and L(f ) ≤k L kk f k. z8. If f, g ∈ Cc(X), f, g ≥ 0, supp f, supp g ⊆ K where K is compact, then R(f ) − R(g) ≤k f − g k µ(K). If X is compact we also have: (i) If c is a constant then R(c) = cµ(X) and, hence, R(f + c) = R(f ) + cµ(X). (ii) For any functions f, g ∈ D(ρ) R(f ) − R(g) ≤ R(1) k f − g k= µ(X) k f − g k . Proof. By Theorem 6.7, if 0 ≤ f ≤ 1 then supp r ⊆ [0, 1]. z1. Using formula (7.2) and part (I) of Theorem 6.7, we have: R(f ) = RR id dr ≤ 1 · r({t : t > 0}) = µ(f −1(0, ∞)) ≤ µ(U ). 25 z2. Using part (I) of Theorem 6.7 we have: R(f ) =ZR id dr ≥ 1 · r({t : t = 1}) = µ(f −1([1, ∞)) ≥ µ(C). z3. Let a = infx∈X f (x), b = supx∈X f (x). It is enough to assume a /∈ E (see part (I) of Theorem 6.7), for otherwise we may take an ր a, an /∈ E. Then aµ(X) = aµ(f −1([a, ∞))) = ar([a, ∞)) = ar([a, b]) ≤Z b a id dr = R(f ) ≤ br([a, ∞)) = bµ(X). Because of formula (7.6), the statement for L(f ) also holds. z4. By part z1, sup{R(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U } ≤ µ(U ). For ǫ > 0 choose K ∈ K (X) such that µ(U ) − µ(K) < ǫ. Pick f ∈ Cc(X) such that 0 ≤ f ≤ 1, supp f ⊆ U, f = 1 on K. Then by part z2 R(f ) ≥ µ(K) ≥ µ(U ) − ǫ. It follows that sup{R(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U } = µ(U ). z5. Take any U ∈ O(X) containing K. Taking an Urysohn function f such that f = 1 on K, supp f ⊆ U and using part z1 we see that inf{R(g) : g ∈ Cc(X), g ≥ 1K} ≤ R(f ) ≤ µρ(U ). Taking the infimum over all open sets containing K we have: inf{R(g) : g ∈ Cc(X), g ≥ 1K } ≤ µρ(K). To prove the opposite inequality, take any g ∈ Cc(X) such that g ≥ 1K . Let 0 < δ < 1. Let U = {x : g(x) > 1 − δ}. Then U is open and K ⊆ U . Consider function h = inf{g, 1 − δ}, so h ∈ Cc(X). Since 0 ≤ h ≤ g we have R(h) ≤ R(g). Because = 1 on U , for any function h 1 − δ h f ∈ Cc(X), 0 ≤ f ≤ 1, supp f ⊆ U we have f ≤ Lemma 7.7 1 − δ and so by parts (iii) and (iv) of R(f ) ≤ R(cid:18) h 1 − δ(cid:19) = R(h) 1 − δ ≤ R(g) 1 − δ . Then µ(K) ≤ µ(U ) = sup(cid:8)R(f ) : f ∈ Cc(X), 0 ≤ f ≤ 1U , supp f ⊆ U(cid:9) ≤ R(g) 1 − δ . Thus, for any g ∈ Cc(X) such that g ≥ 1K and any 0 < δ < 1 Therefore, µ(K) ≤ inf{R(g) : g ∈ Cb(X), g ≥ 1K }. (1 − δ)µ(K) ≤ R(g). z6. In the argument for part z5 we may use, respectively, g ∈ Cc(X), g ≥ 1K , 0 ≤ g ≤ 1 or g ∈ C0(X), g ≥ 1K or g ∈ C0(X), g ≥ 1K , 0 ≤ g ≤ 1. 26 z7. By part z4 we see that µ(X) ≤k R k. Using also part z3 we have k R k= µ(X), and by formula (7.6) also k L k= µ(X). z8. It is enough to consider K = supp f ∪ supp g. For any function h ∈ Cc(X) such that h ≥ 0, h = 1 on K as in formula (3.1) in the proof of part (V) of Lemma 3.6 we have: R(f ) − R(g) ≤k f − g k R(h). Taking the infimum over functions h, by part z5 we obtain the assertion. (i) By formula (7.3), R(c) = cµ(X), and the rest of the statement follows from Theorem 7.10. (ii) Since R is monotone, an r-functional, and µ(X) =k R k= R(1) the statement follows from Remark 4.5. Remark 7.13. The proof of part z5 is similar to the one for [5, Lemma 35(p4), Sect. 4], which, in turn, follows a proof from [9]. Remark 7.14. Part z8 of Lemma 7.12 means that on C + with respect to topology of uniform convergence on compacta. 0 (X) the functional R is continuous Definition 7.15. Let f be a continuous function on X on a locally compact space X. Con- sider the σ-algebra of subsets of X Σf = f −1(B(R)), where B(R) are the Borel subsets of R. Let r, l, m be measures on R as in Theorem 6.7 . On Σf define measure nr, nl, n by setting for each E ∈ Σf nr(E) = r(f (E)), nl(E) = l(f (E)), n(E) = m(f (E)) Remark 7.16. Definition 7.15 leads to another way to represent functionals R and L. Let Σf and measures nr, nl be as in Definition 7.15. Then for any set D ∈ B(R) we have: nr ◦ f −1(D) = r(f (f −1(D))) = r(D), nl ◦ f −1(D) = l(D), i.e. By formula (7.2) i.e. Similarly, nr ◦ f −1 = r, nl ◦ f −1 = l and n ◦ f −1 = m on B(R). (7.7) R(f ) =ZR id dr =ZR id d(nr ◦ f −1) =ZX f dnr, f dnr. f dnl. R(f ) =ZX L(f ) =ZX 27 Lemma 7.17. Let µ be a finite deficient topological measure on a locally compact space X, f ∈ C0(X), and nr, nl be measures defined in Definition 7.15. (i) nr(f −1((t, ∞)) = µ(f −1((t, ∞)) for all t, nr(f −1([t, ∞)) = µ(f −1([t, ∞)) for all t /∈ E, where E ⊆ (−∞, 0] is a countable set from part V of Lemma 6.3; nl(f −1((−∞, t)) = µ(f −1((−∞, t)) for all t, nl(f −1((−∞, t]) = µ(f −1((−∞, t]) for all t /∈ E1, where E1 ⊆ [0, ∞) is a countable set from part V of Lemma 6.3. (ii) For any open or closed set A ⊆ R µ(f −1(A)) ≤ nr(f −1(A)), µ(f −1(A)) ≤ nl(f −1(A)). (iii) If µ is a finite topological measure and f ∈ C0(X), then also n(f −1(A)) = µ(f −1(A) for any open or closed set A ⊆ R \ {0}. (iv) If µ is a finite topological measure and X is compact, then n(f −1(A)) = µ(f −1(A)) for any open or closed set A ⊆ R. Proof. Follows from (7.7), Theorem 6.7, and Theorem 6.10. Remark 7.18. Definition 7.15 and Remark 7.16 were first observed for the case of the compact space X by M. Svistula, see [15, (3.4)]. 8 Representation Theorems for deficient topological measures Theorem 8.1 (Representation theorem). Let µ be a finite deficient topological measure on a locally compact space X. (i) Then there exists a unique p-conic quasi-linear functional R on C + 0 (X) of finite norm such that µ = µR and k R k= µ(X). In fact, R = Rµ. (ii) If X is compact, the unique functional R can be taken to be a real-valued r-functional on C(X). Here Rµ and µR are as in Definition 7.1 and Definition 5.1. Proof. (i) Let µ be a finite deficient topological measure on a locally compact space X, and let R = Rµ be a functional on C + 0 (X) obtained from µ as in Definition 7.1. Note that R is a d-functional by Theorem 7.10 and Remark 4.17. Then µR defined as in Definition 5.1 from R is a deficient topological measure by Theorem 5.4. From part y4 of Lemma 5.2 and part z5 of Lemma 7.12 we see that µ(K) = µR(K) for every compact K. Thus, µ = µR. By part z7 of Lemma 7.12 µ(X) =k R k . Now we shall show the uniqueness. Let η be another p-conic quasi-linear functional on C + 0 (X) of finite norm such that µη = µR. Then for some a ≥ 1 k η k≤ aµ(X). 28 Let f ∈ C + 0 (X). Both ρ and η are positive-homogeneous, so we may assume that f (X) = [0, 1]. For n ∈ N let ti = i/n, i = 0, . . . , n. For i = 1, . . . , n consider functions φi defined as follows: φi(t) =   0 t − ti−1 1 n if t ≤ ti−1 if ti−1 < t < ti if t ≥ ti . With functions fi = φi ◦ f we have f = Pn i=1 fi. Since each φi is non-decreasing, and φi(0) = 0, each fi ∈ A+(f ). Since R and η are both p-conic quasi-linear functionals, we have n R(f ) = R( Xi=1 fi) = n Xi=1 R(fi), η(f ) = n Xi=1 η(fi). (8.1) By part z1 of Lemma 7.12 R(nf1) ≤ µ(X), so 0 ≤ R(f1) ≤ 1 0 ≤ η(nf1) ≤k η k≤ aµ(X), i.e. 0 ≤ η(f1) ≤ a n µ(X). Thus, n µ(X) ≤ a n µ(X). We have R(f1) − η(f1) ≤ aµ(X) n . (8.2) For each i = 2, . . . , n let Ki = f −1([ti, ∞)), Vi = f −1((ti−1, ∞)). Choose an open set Ui ⊆ Vi such that µ(Ui) − µ(Ki) < 1 n and then pick an Urysohn function gi ∈ Cc(X) such that 0 ≤ gi ≤ 1 n on Ki and supp gi ⊆ Ui ⊆ Vi. Since ngi = 1 on K and µ = µR = µη, by part y4 of Lemma 5.2 and Definition 5.1 µ(Ki) ≤ R(ngi), η(ngi) ≤ µ(Ui), and so R(ngi) − η(ngi) ≤ µ(Ui) − µ(Ki) ≤ 1 n , gi = 1 n . Then R(gi) − η(gi) ≤ 1 n2 . (8.3) Let g = Pn i=2 fi. Since supp (g3 + . . . + gn) ⊆ K2 and g2 = 1 on K2, by part (III) of Lemma 3.6 ρ(g2 + g3 + . . . gn) = ρ(g2) + ρ(g3 + . . . gn). Similarly for η. By induction i=2 gi, h = Pn R(g) = n Xi=2 R(gi), η(g) = η(gi). n Xi=2 Then R(g) − η(g) ≤ R(gi) − η(gi) ≤ n − 1 n2 < 1 n . n Xi=2 Note that k g − h k≤ 1 n , so by part (V) of Lemma 3.6 R(g) − R(h) ≤ µ(X) n , η(g) − η(h) ≤ aµ(X) n . Using (8.2), (8.6), and (8.5) we obtain: R(f ) − η(f ) ≤ R(f1) − η(f1) + R(h) − η(h) 29 (8.4) (8.5) (8.6) ≤ R(f1) − η(f1) + R(h) − R(g) + R(g) − η(g) + η(g) − η(h) ≤ aµ(X) n + µ(X) n + 1 n + aµ(X) n ≤ 1 n (3aµ(X) + 1). Thus, R = η. (ii) Now let X be compact. We shall show that the proof for part (i) still applies, although the reasoning for some estimates is different. We define the functional R = Rµ on C(X). It is an r-functional by Theorem 7.10. Then R(1) =k R k= µ(X) < ∞, and R(−1) ≤k R k< ∞. By monotonicity, R is real-valued. Let η be another real-valued (hence, bounded by Proposition 3.7) r-functional on C(X) such that µ = µR = µη. To show the uniqueness, by Remark 4.17 it is enough to show that ρ(f ) = η(f ) for f ≥ 0. For the functions fi from the proof for part (i) note the following: supp (f2 + . . . + fn) ⊆ f −1([t1, ∞)), n on f −1([t1, ∞)), thus by part (III) of Lemma 4.19 we may apply the c-level condition to obtain f1 = 1 R(f1 + f2 + . . . + fn) = R(f1) + R(f2 + . . . + fn). Then by induction we may show that formula (8.1) holds for R and, similarly, for η. In the same manner, by part (III) of Lemma 4.19 and induction we show that formula (8.4) holds. Note that (8.2), (8.3), and (8.5) hold as in the proof of part (i). Estimations (8.6) are valid by part (II) of Lemma 4.19. Now as in the end of the proof for part (i), we show that R = η. The proof is complete now. Remark 8.2. Our inequality (8.3) is inspired by a similar estimate in the proof of [15, Theorem 9]. Definition 8.3. Let L, QI, Φs, Φ+, Φ−, Φr, Φl represent subfamilies of bounded functionals from L, QI, Φs, Φ+, Φ−, Φr, Φl respectively. We may indicate in parenthesis the domain of functionals. For example, Φ+ = Φ+(C(X)) is the collection of all bounded p-conic quasi- linear functionals on C(X). Definition 8.4. Let DTM(X), TM(X), M(X) represent, respectively, subfamilies of finite set functions from DT M (X), T M (X), M (X). Corollary 8.5. Let X be locally compact. (i) There is a bijection Γ : Φ+(C + bijection Π = Γ−1 is given by Π : DTM(X) −→ Φ+(C + µR and Rµ are according to Definition 5.1 and Definition 7.1. 0 (X)) −→ DTM(X) given by Γ(R) = µR. The inverse 0 (X)) where Π(µ) = Rµ. Here (ii) There is a bijection ∆ : DTM(X) −→ Φ−(C − 0 (X)) given by ∆(µ) = Lµ, where Lµ is according to Definition 7.1. (iii) If X is compact, there is a bijection between DTM(X) and Φr(C(X)) given by µ 7→ Rµ, and a bijection between DTM(X) and Φl(C(X)). Proof. (i) Follows from Theorem 8.1. 30 (ii) By Remark 3.5 there is a bijection between Φ+(C + bijection ∆. By formula (7.6) −Rµ(−f ) = Lµ(f ). 0 (X)) and Φ−(C − 0 (X)), so we obtain (iii) By Theorem 8.1 there is a bijection between DTM(X) and Φr(C(X)) given by µ 7→ Rµ. By Remark 4.17 there is a bijection between Φr(C(X)) and Φl(C(X)). Corollary 8.6. Suppose ρ is a finite c-functional on C0(X), and ρ = Rµ on C + 0 (X), where Rµ is a functional on C0(X) for some finite deficient topological measure µ as in Definition 7.1. Then ρ is an s-functional iff Lµ(g) = ρ(g) for all g ∈ C − 0 (X), where Lµ is a functional on C0(X) as in Definition 7.1. Proof. If ρ(g) = Lµ(g) for all g ∈ C − −ρ(g) = ρ(−g). Since µ is finite, ρ is real-valued on C + s-functional by part (i) of Proposition 4.13. The other direction can be proved similarly. 0 (X) then ρ(g) = Lµ(g) = −Rµ(−g) = −ρ(−g), i.e. 0 (X), so ρ is real-valued. Then ρ is an Theorem 8.7. Let X be locally compact. Let Φ+ = Φ+(C + 0 (X)). Consider the map Π : DTM(X) −→ Φ+ where Π(µ) = Rµ and Rµ is the functional according to Definition 7.1. Then the map Π has the following properties: (I) Π is conic-linear, i.e. Π(cµ + dν) = cΠ(µ) + dΠ(ν), c, d ≥ 0. (II) µ ≤ ν if and only if Π(µ) ≤ Π(ν) (i.e., Rµ(f ) ≤ Rν(f ) for all f ∈ C + 0 (X)). (III) µ ∈ TM(X) iff ρ is a quasi-linear functional on C0(X), and µ ∈ M(X) iff ρ is a linear functional on C0(X), where ρ(f ) = Π(µ)(f +) − Π(µ)(f −). (IV) k Rµ k= µ(X). (V) The map Π : DTM(X) −→ Φ+ is a conic-linear order-preserving bijection such that k Rµ k= µ(X). Proof. (I) Let λ = cµ+dν, c, d ≥ 0. Take any f ∈ C + 0 (X). For function R1,λ,f in Definition 6.1 we see that R1,λ,f = cR1,µ,f +dR1,ν,f . From formula (7.4) Rλ(f ) = cRµ(f )+dRν (f ), and the statement follows. (II) Let µ ≤ ν. Take any f ∈ C + 0 (X). Using Definition 6.1 we see that R1,µ,f ≤ R1,ν,f . Then by formula (7.4) we have Rµ(f ) ≤ Rν(f ). Thus, Rµ ≤ Rν. Now assume that Rµ ≤ Rν . From part z5 of Lemma 7.12 we see that µ(K) ≤ ν(K) for any compact K. By Remark 2.3 µ ≤ ν. (III) Suppose ρ(f ) = Rµ(f +) − Rµ(f −) is a quasi-linear functional on C0(X). By part (i) of Corollary 8.5 and part (vi) of Theorem 5.4 µ is a finite topological measure. Now suppose that µ is a finite topological measure. Let mf be the measure from Theorem 6.10 for f , where f ∈ C0(X). Consider functional ρ(f ) = RR id dmf on C0(X). For φ ∈ C(f (X)) and any open set W ⊆ R we have mφ◦f (W ) = µ((φ ◦ f )−1(W )) = µ(f −1(φ−1(W ))) = mf (φ−1(W )) = (mf ◦ φ−1)(W ), thus, mφ◦f and mf ◦ φ−1 are equal as measures on R, and for φ ◦ f ∈ A(f ) we obtain ρ(φ ◦ f ) =ZR id dmφ◦f =ZR id d(mf ◦ φ−1) =ZR φ dmf . 31 For φ ◦ f, ψ ◦ f ∈ A(f ) we have: (φ + ψ) dmf =ZR φ dmf +ZR ρ(φ ◦ f + ψ ◦ f ) =ZR For any const c we see that ρ(cf ) = ρ((c id) ◦ f ) = RR c id dmf = cρ(f ). Thus, ρ is a quasi-linear functional on C0(X). Since f + = (0 ∨ id) ◦ f ∈ A(f ), and f − ∈ A(f ), we know that ρ(f ) = ρ(f +) − ρ(f −). It is clear that ρ(g) = Rµ(g) for any g ∈ C + 0 , so ρ(f ) = ρ(f +) − ρ(f −) = Rµ(f +) − Rµ(f −) = Π(µ)(f +) − Π(µ)(f −). ψ dmf = ρ(φ ◦ f ) + ρ(ψ ◦ f ). If µ ∈ M(X) then ρ(f ) =R id dmf =R f dµ is the usual integral, and the last assertion is a well known fact. (IV) This is part z7 of Lemma 7.12. (V) Clear from part (i) of Corollary 8.5 and parts (I), (II), and (IV). Theorem 8.8. Let X be compact. (i) QI = Φs. (ii) Φ+ = Φr and Φ− = Φl. Here all functionals are on C(X). Proof. (i) By Remark 4.12 we need to show that Φs ⊆ QI, so let ρ ∈ Φs. Then ρ ∈ Φr by Remark 4.17. By Corollary 8.5 there is a unique deficient topological measure µ such that ρ(f ) = ρµ(f ) for all f ∈ C(X). By part (v) of Theorem 5.4, µ is a topological measure. Then ρ is a quasi-linear functional (see [1, Theorem 4.1] or [5, Theorem 42, Sect. 4]). (ii) If ρ is an r-functional, by Corollary 8.5 ρ = Rµ for a unique deficient topological measure µ. By Theorem 7.10 Rµ is a p-conic quasi-linear functional. Thus, Φr ⊆ Φ+. The other inclusion is given by Lemma 4.18. We can prove that Φ− = Φl in a similar way, using Corollary 8.5 and Lemma 4.18. Remark 8.9. From Theorem 8.1, part (iii) of Corollary 8.5, and Theorem 8.8, it follows that when X is compact, in Theorem 8.7 we may take Φ+ to be Φ+(C(X) = Φr(C(X)). Theorem 8.10. (I) Let X be locally compact. For functionals on C0(X) we have: L ⊆ QI ⊆ Φ+ ∩ Φ− ⊆ Φr ∩ Φl, L ⊆ QI ⊆ Φ+ ∩ Φ− ⊆ Φr ∩ Φl = Φs. In general, L $ QI $ Φ+. (II) Let X be compact. Then for functionals on C(X) we have: L ⊆ QI = Φ+ ∩ Φ− = Φr ∩ Φl = Φs. In general, L $ QI. 32 Proof. (I) The inclusion QI ⊆ Φ+ ∩ Φ− is given by Remark 3.4. The inclusion Φ+ ∩ Φ− ⊆ Φr ∩ Φl follows from Lemma 4.18. By Remark 4.17 Φr ⊆ Φc, so from Corollary 8.6 we see that Φr ∩ Φl ⊆ Φs. By part (ii) of Remark 4.17 we have: Φs ⊆ Φr ∩ Φl. The proper inclusion L $ QI follows from the existence of quasi-linear functionals that are not linear, or existence of topological measures that are not measures. The proper inclusion QI $ Φ+ follows from part (III) of Theorem 8.7 and the existence of deficient topological measures that are not topological measures. See Remark 2.5. For an example of a quasi-linear but not linear functional on a locally compact space see [5, Example 55, Sect. 5]. (II) Use the previous part and Theorem 8.8. Remark 8.11. Let X be compact. From Corollary 8.5 and Theorem 8.8 we see that the functionals corresponding to finite deficient topological measures can be described in four ways: as p-conic quasi-linear functionals, as r-functionals, as n-conic quasi-linear functionals, and as l-functionals. From Theorem 8.10 we see that the functionals corresponding to finite topological mea- sures can be described in four ways: as quasi-linear functionals; as s-functionals; as functionals that are both p-conic quasi-linear and n-conic quasi-linear; and as functionals that are both r- and l-functionals. Remark 8.12. Theorem 8.10 answers positively the question posed in [15, Remark 7], of whether for a compact space Φs = Φr ∩ Φl. Note that by part (I) of Lemma 4.19 our definition of r- and l-functionals in the compact case coincide with those in [15, Definition 6]. By Definition 4.9 and Theorem 8.10 our definition of an s-functional coincides with the one in [15, Definition 6], where it was first introduced. Acknowledgments. This work was conducted at the Department of Mathematics at the University of California Santa Barbara. The author would like to thank the department for its hospitality and supportive environment. References [1] J. Aarnes. Quasi-states and quasi-measures. Adv. Math., 86 (1): 41 -- 67, 1991. [2] S. Butler. Ways of obtaining topological measures on locally compact spaces. Bulletin of Irkutsk State University, Series "Mathematics", 25: 33 -- 45, 2018. [3] S. Butler. 1902.01957 Solid-set functions and topological measures on locally compact spaces. arXiv: [4] S. Butler. Deficient topological measures on locally compact spaces. arXiv: 1902.02458 [5] S. Butler. Quasi-linear functionals on locally compact spaces. arXiv: 1902.03358 [6] J. Dugundji. Topology. Allyn and Bacon, Inc., Boston, 1966. [7] R. Engelking. General topology. PWN, Warsaw, 1989. [8] D. Grubb. Signed quasi-measures. Trans. Amer. Math. Soc., 349(3): 1081 -- 1089, 1997. [9] D. Grubb. Lectures on quasi-measures and quasi-linear functionals on compact spaces. Unpub- lished, 1998. 33 [10] Ø. Johansen, A. Rustad. Construction and Properties of quasi-linear functionals. Trans. Amer. Math. Soc. 358(6): 2735 -- 2758, 2006. [11] M. Entov, L. Polterovich. Quasi-states and symplectic intersections. Comment. Math. Helv., 81: 75 -- 99, 2006. [12] L. Polterovich, D. Rosen. Function theory on symplectic manifolds. CRM Monograph series, vol. 34. American Mathematical Society, Providence, Rhode Island, 2014. [13] A. Rustad. Unbounded quasi-integrals. Proc. Amer. Math. Soc., 129(1): 165 -- 172, 2000. [14] M. Svistula. A Signed quasi-measure decomposition. Vestnik Samara Gos. Univ. Estestven- nonauchn., 62(3):192 -- 207, 2008. (in Russian) [15] M. Svistula. Deficient topological measures and functionals generated by them. Sbornik: Math- ematics, 204(5): 726 -- 76, 2013. 34
1005.1996
1
1005
2010-05-12T06:50:12
The canonical injection of the Hardy-Orlicz space $H^\Psi$ into the Bergman-Orlicz space ${\mathfrak B}^\Psi$
[ "math.FA" ]
We study the canonical injection from the Hardy-Orlicz space $H^\Psi$ into the Bergman-Orlicz space ${\mathfrak B}^\Psi$.
math.FA
math
The canonical injection of the Hardy-Orlicz space H Ψ into the Bergman-Orlicz space BΨ Pascal Lefèvre, Daniel Li, Hervé Queffélec, Luis Rodríguez-Piazza November 20, 2018 Abstract. We study the canonical injection from the Hardy-Orlicz space H Ψ into the Bergman-Orlicz space BΨ. Mathematics Subject Classification. Primary: 46E30 -- Secondary: 30D55; 30H05; 32A35; 32A36; 42B30 Key-words. absolutely summing operator -- Bergman-Orlicz space -- compact- ness -- Dunford-Pettis operator -- Hardy-Orlicz space -- weak compactness 1 Introduction and notation 1.1 Introduction There are two natural Orlicz spaces of analytic functions on the unit disk D of the complex plane: the Hardy-Orlicz space H Ψ and the Bergman-Orlicz space BΨ. It is well-known that in the classical case Ψ(x) = xp, H p ⊆ Bp and the canonical injection Jp from H p to Bp is bounded, and even compact. In fact, for any Orlicz function Ψ, one has H Ψ ⊆ BΨ and the canonical injection JΨ : H Ψ → BΨ is bounded, but we shall see in this paper that its compactness requires that Ψ does not grow too fast. We actually characterize in Section 2 the compactness: JΨ is compact if and only if limx→+∞ Ψ(Ax)/[Ψ(x)]2 = 0 for every A > 1, and the weak compactness: JΨ is weakly compact if and only if lim supx→+∞ Ψ(Ax)/[Ψ(x)]2 < +∞ for every A > 1 . We show that, if these two properties are "often" equivalent (this happens for example if Ψ(x)/x is non-decreasing for x large enough), it is not always the case. We actually show a stronger result in Section 4: there is an Orlicz function Ψ such that JΨ is weakly compact and is Dunford-Pettis, but such that JΨ is not compact. 1.2 Notation An Orlicz function is a non-decreasing convex function Ψ : [0, +∞[→ [0, +∞[ such that Ψ(0) = 0 and Ψ(∞) = ∞. One says that the Orlicz function Ψ has 1 Ψ(βx)/Ψ(x) = +∞. It is equivalent to say that, for every β > 1, Ψ(βx) ≤ CβΨ(x). property ∆2 (Ψ ∈ ∆2) if Ψ(2x) ≤ C Ψ(x) for some constant C > 0 and x large It enough. is known that if Ψ ∈ ∆2, then Ψ(x) = O (xp) for some 1 ≤ p < +∞. One says (see [6], [7]) that Ψ satisfies the condition ∆0 if, for some β > 1, one +∞ for every has lim x→∞ 1 ≤ p < ∞. Indeed, let 1 ≤ p < ∞. For every β > 1 one can find x0 > 0 such that Ψ(βx)/Ψ(x) ≥ βp for x ≥ x0; then Ψ(βnx0) ≥ βnpΨ(x0) for every n ≥ 1. That implies that Ψ(x) ≥ Cp xp for every x > 0 large enough. Since p ≥ 1 is arbitrary, we get xp = o [Ψ(x)]. We say that Ψ ∈ ∇0(1) if, for every A > 1, Ψ(Ax)/Ψ(x) is non-decreasing for x large enough. This is equivalent to say (see [7], Proposition 4.7) that log Ψ(ex) is convex. When Ψ ∈ ∇0(1), one has either Ψ ∈ ∆2, or Ψ ∈ ∆0. If Ψ ∈ ∆0, then Ψ(x)/xp −→x→∞ If (S,S, µ) is a finite measure space, one defines the Orlicz space LΨ(µ) as the set of all (classes of) measurable functions f : S → C for which there is a C > 0 such that RS Ψ(f/C) dµ is finite. The norm kfkΨ is the infimum of all C > 0 for which the above integral is ≤ 1. The Morse-Transue space M Ψ(µ) is the subspace of f ∈ LΨ(µ) for which RS Ψ(f/C) dµ is finite for all C > 0; it is the closure of L∞(µ) in LΨ(µ). One has M Ψ(µ) = LΨ(µ) if and only if Ψ ∈ ∆2. If Ψ(x)/x −→x→+∞ +∞, the conjugate function Φ of Ψ is defined by Φ(y) = supx>0(cid:0)xy − Ψ(x)(cid:1). It is an Orlicz function and [M Ψ(µ)]∗ = LΦ(µ), isomorphi- We may note that if Ψ(x)/x does not converges to infinity, we must have Ψ(x) ≤ ax for some a ≥ 1 and x large enough. Then LΨ(µ) = L1(µ) isomorphi- cally and then Φ(y) = +∞ for y > a (giving LΦ(µ) = L∞(µ) isomorphically). cally. We denote by D the open unit disk of C and by T = ∂D the unit circle. The normalized area-measure on D is denoted by A and the normalized Lebesgue measure on T is denoted by m. The Hardy-Orlicz space H Ψ is defined as {f ∈ H 1 ; f ∗ ∈ LΨ(m)}, where f ∗ is the boundary values function of f , and HM Ψ = H Ψ ∩ M Ψ(m) is the closure of H ∞ in H Ψ. The Bergman-Orlicz space BΨ is the subspace of analytic f ∈ LΨ(A), and BM Ψ = BΨ ∩ M Ψ(A) is the closure of H ∞ in BΨ. Since, for f ∈ H Ψ, kfkHΨ = sup0<r<1 kfrkHΨ (see [7], Proposition 3.1), where fr(z) = f (rz), one has: Z 2π 0 Ψ(cid:18)f (reit) kfkHΨ (cid:19) dt kfkHΨ (cid:19) dA =Z 1 Ψ(cid:18)f (reit) 2π ≤Z 2π 0 0 (cid:20)Z 2π 0 kfrkHΨ (cid:19) dt Ψ(cid:18)f (reit) kfkHΨ (cid:19) dt Ψ(cid:18)f (reit) 2π ≤ 1 ; 2π(cid:21) 2r dr ≤ 1 , hence: ZD so f ∈ BΨ and kfkBΨ ≤ kfkHΨ. It follows that H Ψ ⊆ BΨ and the canonical injection JΨ : H Ψ → BΨ is bounded, and has norm 1. Let us point out that 2 the boundedness also follows from [7], Theorem 4.10, 2), since JΨ is a Carleson embedding JΨ : H Ψ → BΨ ⊆ LΨ(A). This injection is not onto, since there are functions f ∈ BΨ with no radial limit on a subset of T of positive measure (the proof is the same as in Bp: see [4], § 3.2, Lemma 2, page 81). Note that JΨ is not an into-isomorphism: take fn(z) = zn, for every n ∈ N; it is easy to see that {fn}n tends to 0 in BΨ, but not in H Ψ. Acknowledgment. This work is partially supported by a Spanish research project MTM 2009-08934. Part of this paper was made during an invitation of the second-named author by the Departamento de Análisis Matemático of the Universidad de Sevilla. It is a pleasure to thanks the members of this department for their warm hospitality. 2 Compactness and weak-compactness In order to characterize the compactness and the weak-compactness of JΨ, we introduce the following quantity QA, A > 1: (2.1) QA = lim sup x→+∞ Ψ(Ax) [Ψ(x)]2 , which will turn out to be essential. We are going to start with the compactness. Theorem 2.1 The canonical injection JΨ : H Ψ → BΨ is compact if and only if (2.2) lim x→+∞ Ψ(Ax) [Ψ(x)]2 = 0 for every A > 1 . Remarks. 1) Condition (2.2) means that QA = 0 for every A > 1. equivalent to say that: It is (2.3) sup A>1 QA < +∞. Indeed, assume that M := supA>1 QA < +∞. Let 0 < ε ≤ 1 and A > 1; we can find xA = xA(ε) > 0 such that Ψ(Ax/ε)/[Ψ(x)]2 ≤ 2M for x ≥ xA. By convexity, one has Ψ(Ax) ≤ ε Ψ(Ax/ε), and hence Ψ(Ax)/[Ψ(x)]2 ≤ 2εM for x ≥ xA. We get QA = 0. 2) It is clear that condition (2.2) is satisfied whenever Ψ ∈ ∆2, but Ψ(x) = − 1 satisfies (2.2) without being in ∆2. However, condition (2.2) e[log(x+1)]2 implies that Ψ cannot grow too fast. More precisely, we must have Ψ(x) = o (exα ) for every α > 0 . 3 Indeed, one has Ψ(At) ≤ [Ψ(t)]2 for t ≥ tA, and, by iteration, Ψ(AntA) ≤ [Ψ(tA)]2n for every n ≥ 1. For every x > 0 large enough, taking n ≥ 1 such that AntA ≤ x < An+1tA, we get Ψ(x) ≤ C1 eC2xα , with α = log 2/ log A. Since A > 1 is arbitrary, α may be any positive number. The little-oh condition follows from the fact that the inequality is true for all α > 0. Proof of Theorem 2.1. By definition, BΨ is a subspace of LΨ(D,A); hence we can see JΨ as a Carleson embedding JΨ : H Ψ → LΨ(D,A). If S(ξ, h) = {z ∈ D ; z − ξ < h}, the compactness of JΨ implies, by [7], Theorem 4.11, that, for every A > 1, every ε > 0, and h > 0 small enough: h2 ≤ 4 A[S(ξ, h)] ≤ 4ε Ψ[AΨ−1(1/h)] , that is, setting x = Ψ−1(1/h), Ψ(Ax) ≤ 4ε [Ψ(x)]2, and (2.2) is satisfied. Conversely, one has: sup 0<t≤h sup ξ=1 A[S(ξ, t)] t ≤ sup 0<t≤h t2 t = h , which is o(cid:0)(1/h)/Ψ[AΨ−1(1/h)](cid:1) for every A > 1, if (2.2) holds; hence, by [7], Theorem 4.11, again, JΨ is compact. (cid:3) We now turn ourself to the weak compactness. Theorem 2.2 The following assertions are equivalent: (a) JΨ : H Ψ → BΨ is weakly compact; (b) JΨ fixes no copy of c0; (c) JΨ fixes no copy of ℓ∞; (d) QA < +∞, for every A > 1; (e) H Ψ ⊆ BM Ψ; (f) JΨ is strictly singular. Recall that an operator T : X → Y between two Banach spaces is said to be strictly singular if there is no infinite-dimensional subspace X0 of X on which T is an into-isomorphism. The proof will be somewhat long, and before beginning it, we shall remark that if Ψ ∈ ∆0, then condition (2.4) QA < +∞ for every A > 1 implies condition (2.2). Indeed, if lim x→+∞ Ψ(βx) Ψ(x) = +∞, we get, for every A > 1: lim sup x→+∞ Ψ(Ax) [Ψ(x)]2 = lim sup x→+∞ Ψ(Ax) Ψ(βAx) Ψ(βAx) [Ψ(x)]2 ≤ lim sup x→+∞ Ψ(Ax) Ψ(βAx) QβA = 0 . Now, if, for some A > 1, Ψ(Ax)/Ψ(x) is non-decreasing for x large enough (in particular if Ψ ∈ ∇0(1)), one has the dichotomy: either Ψ ∈ ∆2, and then JΨ is compact; or Ψ ∈ ∆0 and hence the weak compactness of JΨ implies, by the two above theorems, its compactness. Hence: 4 Proposition 2.3 If, for some A > 1, Ψ(Ax)/Ψ(x) is non-decreasing, for x large enough, then the weak compactness of JΨ is equivalent to its compactness. However, it is easy to construct an Orlicz function Ψ which satisfies condi- tion (2.4), but not condition (2.2). We do not give an axample here because we have a stronger result in Section 4. In order to prove Theorem 2.2, we shall need several lemmas. Lemma 2.4 Let Ψ be any Orlicz function. If we define Ψ1(t) = [Ψ(t)]2, t ≥ 0, then Ψ1 is an Orlicz function for which H Ψ ⊆ BΨ1 and the canonical injection of H Ψ into BΨ1 is continuous. Proof. It is enough to see that H Ψ continuously embeds into LΨ1(A), and for this we can use Theorem 4.10 in [7]. Following the notation of that theorem for the measure µ = A, it is easy to see that, as h → 0+, ρA(h) ≈ h2, and KA(h) ≈ h. Observe that, for t > 1, we have Ψ1[Ψ−1(t)] = t2, and so, for h ∈ (0, 1), 1/h 1/h2 = h (cid:23) KA(h). Using part 2) of Theorem 4.10 in [7], the lemma follows. Ψ1[Ψ−1(1/h)] 1/h = (cid:3) k=1 αkfnk (z): Lemma 2.5 Let M > δ > 0 and {fn}n be a sequence in H Ψ ∩ BM Ψ such that: (a) {fn}n tends to 0 uniformly on compact subsets of D; (b) kfnkBΨ ≥ δ, for every n ≥ 1; (c) kfnkHΨ ≤ M , for every n ≥ 1. Then there exists a subsequence {fnk}k such that Pk fnk (z) < +∞, for every z ∈ D, and for every α = (αk)k ∈ ℓ∞, one has, writing T α(z) = P∞ (δ/2)kαk∞ ≤ kT αkBΨ ≤ 2Mkαk∞. T α ∈ BΨ It is clear that, by (2.5), we are defining an operator T from ℓ∞ Remark. into BΨ which is an isomorphism between ℓ∞ and its image. In particular, the subsequence {fnk}k is equivalent, in BΨ, to the canonical basis of c0. Proof. First we are going to construct, inductively, a subsequence {fnk}k of {fn}, and an increasing sequence {rk}k in (0, 1), such that limk→∞ rk = 1 and, setting (2.5) and and Dk = {z ∈ D ; z ≤ rk} , for k ≥ 1, C1 = D1 , Ck = Dk \ Dk−1 = {z ∈ D ; rk−1 < z ≤ rk}, k ≥ 2, we have: (2.6) fnk (z) ≤ 2−k, for every z ∈ Dk−1, and every k ≥ 2 ; 5 and (2.7) kfnk 1ID\CkkLΨ < δ2−k−2 , for every k ≥ 1. Start the construction by taking n1 = 1. It is a known fact that, for every function f in the Morse-Transue space M Ψ(A), we have (2.8) lim A(A)→0kf 1IAkLΨ = 0. Now, using (2.8), with f = fn1 and considering sets A of the form A = {z ∈ D ; r < z < 1}, we get r1 ∈ (0, 1) so that, for C1 = D1 = {z ∈ D ; z ≤ r1}, we have By the uniform convergence of {fn}n to 0 on D1, we can find n2 > n1 such that kf11ID\C1kLΨ < δ2−3 . fn2 (z) ≤ 1/4, for every z ∈ D1, and kfn21ID1kLΨ < δ2−5 . Using this last inequality and (2.8) again (for f = fn2), we get r2 ∈ (r1, 1), r2 > 1 − 1/2, such that, setting C2 = {z ∈ D ; r1 < z ≤ r2}, we have kfn21ID\C2kLΨ < δ2−4 . Now that we have (2.6) and (2.7) for k = 1 and k = 2, it is clear how we must iterate the inductive construction. At the time of choosing rk ∈ (rk−1, 1), we also impose the condition rk > 1 − 1/k in order to get limk→∞ rk = 1. fnk (z) ≤ 2−k, Once the construction is achieved, let us see why the subsequence {fnk}k works. The condition (2.6) and the fact that limk→∞ rk = 1 imply that, for every compact set K in D and z ∈ D, there exists lK ∈ N such that: for every z ∈ K, and every k ≥ lK . This yields two facts. First, Pk fnk (z) < +∞, for every z ∈ D, and secondly: for every bounded complex sequence α = (αk)k ∈ ℓ∞, the series Pk αkfnk It remains to prove the estimates in (2.5) about the norm of T α in LΨ(A). By homogeneity, we may assume that kαk∞ = 1. Let us write gk = fnk 1ICk and hk = fnk 1ID\Ck , for every k ≥ 1, converges uniformly on compact subsets of D, and its sum, the function T α, is analytic on D. g = ∞Xk=1 αkgk and h = ∞Xk=1 αkhk . We have T α = g + h. By (2.7) and the fact that αk ≤ 1, we have that h ∈ LΨ(A) and khkLΨ ≤ δ/4. By the condition (c) in the statement and the definition of the norm in H Ψ we have, for every n and every r ∈ (0, 1): (2.9) 1 2π Z 2π 0 Ψ(cid:0)fn(reit)/M(cid:1) dt ≤ 1 . 6 The function gk is 0 outside of Ck, and the sequence {Ck}k is a partition of D. Therefore: ZD Ψ(g/M ) dA = ≤ ZCk ZCk ∞Xk=1 ∞Xk=1 Ψ(g/M ) dA = ∞Xk=1 Ψ(fnk/M ) dA . ZCk Ψ(αkfnk/M ) dA Integrating in polar coordinates, setting r0 = 0, and using (2.9), we get: ZD Ψ(g/M ) dA ≤ ≤ ∞Xk=1 ∞Xk=1 rk−1 Z rk Z rk rk−1 1 2π Z 2π 0 2r 2r dr = 1 , Ψ(fnk (reit)/M ) dt dr and therefore kgkLΨ ≤ M , and kT αkLΨ ≤ δ/4 + M ≤ 2M . On the other hand, for every k, we have: kgkLΨ ≥ kg 1ICkkLΨ = αkkfnk − hkkLΨ ≥ αk (δ − δ/22+k) ≥ 3δ 4 αk . Taking the supremum on k, we get kgkLΨ ≥ (3δ/4)kαk∞ = 3δ/4. Consequently, kT αkLΨ ≥ kgkLΨ − khkLΨ ≥ (3δ/4) − δ/4 ≥ δ/2 , and Lemma 2.5 is fully proved. (cid:3) In the following lemma we isolate the proof of the implication (c) =⇒ (d) in the statement of Theorem 2.2. Lemma 2.6 Assume that the Orlicz function Ψ is such that, for some A > 1, (2.10) lim sup x→+∞ Ψ(Ax) [Ψ(x)]2 = +∞ Then the injection JΨ : H Ψ → BΨ fixes a copy of ℓ∞. Proof. Let us take a sequence of positive numbers {dn}n, and a sequence {ξn}n in T, such that the disks {D(ξn, dn)}n are pairwise disjoint in D. In particular, we should have limn→∞ dn = 0. The convexity of Ψ implies the existence of some c > 0 such that Ψ(x) ≥ cx for every x ≥ 1. Given a sequence {βn}n in (4, +∞) to be fixed later, we can find, thanks to (2.10), an increasing sequence {xn} satisfying: (2.11) Ψ(Axn) > βn[Ψ(xn)]2, xn > 1, Ψ(xn) > 1, for every n ∈ N . 7 Define yn as the point in the interval (xn, Axn) such that (2.12) [Ψ(yn)]2 = Ψ(Axn) . Put now hn = 1/Ψ(yn) and rn = 1 − hn. By (2.11) and (2.12), we have [Ψ(yn)]2 > βn > 4, and therefore hn ∈ (0, 1/2). Define un(z) =(cid:16) hn 1 − rn ξnz(cid:17)2 , and fn(z) = yn un(z) . The first condition imposed to βn is βn > 16/d2 n. That gives [Ψ(yn)]2 > n and hn < dn/4. Let us write Dn for the disk D(ξn, dn). Observe that, It is easy to see that kunk∞ = 1, and that kunkH 1 ≤ hn. 16/d2 for z ∈ D \ Dn, we have 1−rn ξnz = 1−rn +rn ξnξn−rn ξnz ≥ rnξn−z−hn ≥ (1/2)dn−hn ≥ dn/4 , and therefore, since [Ψ(xn)]2 ≥ Ψ(xn) ≥ c xn, fn(z) ≤ yn(cid:16) 4hn dn (cid:17)2 16yn n[Ψ(yn)]2 ≤ d2 We also impose the condition βn > 16An2/cd2 = 16Axn nβn[Ψ(xn)]2 ≤ d2 16A nβn · c d2 n, and so we have: (2.13) fn(z) ≤ 1 n2 , for z ∈ D \ Dn . From (2.13) we deduce that {fn}n converges to 0 uniformly on compact subsets of D. Moreover (2.13) yields that, for every bounded sequence {αn}n of complex numbers, the seriesPn≥1 αnfn is uniformly convergent on compact subsets of D. Let us write f ∗ n for the boundary value (on T = ∂D) of the function fn. We claim that : S = (2.14) ∞Xn=1 From this, it is not difficult to deduce that, for every bounded sequence {αn}n of complex numbers, the function P∞ n=1 αnfn is in H Ψ and, for M = kSkLΨ(T), f ∗ n ∈ LΨ(T, m). (2.15) (cid:13)(cid:13)(cid:13) ∞Xn=1 αnfn(cid:13)(cid:13)(cid:13)HΨ ≤ Mk{αn}nk∞ . On the other hand, taking An = {z ∈ D ; constant γ ∈ (0, 1) such that A(An) ≥ γh2 z − ξn ≤ hn}, there exists a n, and, for every z ∈ An, we have: 1 − rn ξnz ≤ 1 − rn + rn ξnξn − rn ξnz = hn + rn z − ξn ≤ 2hn , 8 and consequently un(z) ≥ 1/4. If δ = γ/4A, we have, for every n, Ayn(cid:17) Ψ(cid:16) yn 4δ(cid:17) dA ≥ γh2 Ψ(cid:16)fn nΨ(cid:16) 1 ZD γ δ (cid:17) dA ≥ZAn ≥ h2 nΨ(Ayn) > h2 nΨ(Axn) = 1 . Thus kfnkBΨ ≥ δ, for every n ∈ N. We can apply Lemma 2.5. Using this lemma and (2.15), we get a subsequence {fnk}k such that, for every α = (αk)k ∈ ℓ∞, we have: (δ/2)k{αk}kk∞ ≤(cid:13)(cid:13)(cid:13) ∞Xk=1 αkfnk(cid:13)(cid:13)(cid:13)BΨ ≤(cid:13)(cid:13)(cid:13) ∞Xk=1 αkfnk(cid:13)(cid:13)(cid:13)HΨ ≤ Mk{αk}kk∞ . This clearly says that JΨ fixes a copy of ℓ∞. It remains to prove (2.14). For obtaining this we impose the last condition to the sequence {βn}n. We shall need: ∞Xn=1 (2.16) 1/pβn ≤ 1 . n 1IDn . Thanks to (2.13), S −P∞ Let us set gn = f ∗ n=1 gn is a bounded function. Thus we just need to prove that G =P∞ n=1 gn is in LΨ(T). We have kGkLΨ(T) ≤ A. Indeed, recalling that the Dn's are pairwise disjoint, and that each gn is 0 out of Dn, we have: ∞Xn=1ZDn∩T ZT A(cid:17) dm = Ψ(cid:16) G ≤ Ψ(cid:16) G A(cid:17) dm = A (cid:17) dm Ψ(cid:16) ynu∗ n Ψ(cid:16)f ∗ A (cid:17) dm n and by the convexity of Ψ, and the fact that un ≤ 1, ∞Xn=1ZDn∩T ∞Xn=1ZT ∞Xn=1ZT u∗ ∞Xn=1 ≤ ≤ ∞Xn=1 nΨ(cid:16) yn A(cid:17) dm = ∞Xn=1 Ψ(yn/A) Ψ(yn) ≤ Ψ(xn) Ψ(yn) = kunkH1Ψ(cid:16) yn A(cid:17) ∞Xn=1 pΨ(Axn) ≤ Ψ(xn) ∞Xn=1 1 √βn ≤ 1 , by the required condition (2.16), and that ends the proof of Lemma 2.6. (cid:3) We are now in position to prove Theorem 2.2. Proof of Theorem 2.2. We shall prove that: (a) =⇒ (b) =⇒ (c) =⇒ (d) =⇒ (e) =⇒ (a) , and that (b) ⇐⇒ (f). 9 The implications (a) =⇒ (b) =⇒ (c) and (f) =⇒ (b) are trivial, and we have seen in Lemma 2.6 that (c) =⇒ (d). every f in the unit ball of H Ψ, we have: (d) =⇒ (e). By Lemma 2.4, there exists a constant C > 0 such that, for (2.17) ZD [Ψ(f/C)]2 dA ≤ 1 . For every A > 0, there exist xA, such that Ψ(Ax) ≤ (QA + 1)[Ψ(x)]2, for every x ≥ xA. Thus for every x ≥ 0 we have Ψ(Ax) ≤ (QA + 1)[Ψ(x)]2 + Ψ(AxA). Then, by (2.17), we have ZD Ψ(Af/C) dA < +∞ , for every A > 0 . Therefore f ∈ BM Ψ, for every f in the unit ball of H Ψ, and thus for every f in H Ψ. (e) =⇒ (a). Let {fn}n be in the unit ball of H Ψ. We have to prove that {fn}n has a subsequence which converges in the weak topology of BΨ. By Montel's Theorem {fn}n has a subsequence converging uniformly on compact subsets of D, to a function g which, by Fatou's lemma, also belongs to the unit ball of H Ψ. If this subsequence converges to g in the norm of BΨ we are done. If not, after perhaps a new extraction of subsequence, there exist δ > 0 and a subsequence {fnk}k, such that kfnk − gkBΨ ≥ δ, and kfnk − gkHΨ ≤ 2 . Since moreover {fnk − g}k converges to 0 uniformly on compact subsets of D and, by condition (e), fnk − g ∈ BM Ψ, we may apply Lemma 2.5 and we get that {fnk − g}k has a subsequence equivalent to the canonical basis of c0 in BΨ, and is therefore weakly null. This yields that {fn}n has a subsequence converging to g in the weak topology of BΨ. (b) =⇒ (f). Suppose there exists an infinite-dimensional subspace X of H Ψ on which the norms k · kBΨ and k · kHΨ are equivalent. We shall have finished if we prove that X contains a subspace isomorphic to c0 because then JΨ will fix a copy of c0. We can assume that X is contained in BM Ψ because we already know that (b) implies (e). X being infinite-dimensional, there exists, for every n ∈ N, fn ∈ X, such that kfnkHΨ = 1, and cfn(k) = 0, for k = 0, 1, . . . , n. By the equivalence of the norms in X, there exists δ > 0 such that kfnkBΨ ≥ δ, for every n. The unit ball of H Ψ is compact in the topology of H(D). Since lim n→∞cfn(k) = 0 , for every k ≥ 0 , the only possible limit of a subsequence of {fn}n is the function 0. So {fn}n converges to 0 uniformly on compact subsets of D. As fn ∈ X ⊆ BM Ψ, for 10 every n, we can apply Lemma 2.5, and we get that {fn}n has a subsequence generating an space Y isomorphic to c0 in BΨ. This space Y is contained in X, where the norms are equivalent, so Y is also isomorphic to c0 for the norm of H Ψ. (cid:3) 3 Other properties 3.1 Dunford-Pettis Recall that an operator T : X → Y between two Banach spaces X and Y is said to be Dunford-Pettis if {T xn}n converges in norm whenever {xn}n con- verges weakly. Every compact operator is Dunford-Pettis. The next proposition shows that, in "most" of the cases, these two properties are equivalent for JΨ. Proposition 3.1 If the conjugate function of Ψ satisfies condition ∆2, then JΨ : H Ψ → BΨ is Dunford-Pettis if and only if it is compact. We shall see in Section 4 that without condition ∆2 for the conjugate func- tion, Jψ may be Dunford-Pettis without being compact. Proof. Remark first that speaking of the conjugate function of Ψ implicitly assume that Ψ(x)/x tends to +∞ as x goes to +∞. Assume that JΨ is not compact. By Theorem 2.1, there are some A > 1 and a sequence {xj}j going to +∞ such that Ψ(Axj ) ≥ [Ψ(xj)]2. Setting rj = 1−1/Ψ(xj), this is equivalent to say that AΨ−1(cid:0)1/(1−rj)(cid:1) ≥ Ψ−1(cid:0)1/(1−rj)2(cid:1). Define: 1 − rjz(cid:19)2 fj(z) = xj(cid:18) 1 − rj · One has fj ∈ HM Ψ and kfjkHΨ ≤ 1 (see [7], Corollary 3.10). Since {fj}j converges to 0 uniformly on compact subsets of D, {fj}j converges to 0 in the weak-star topology of H Ψ ([7], Proposition 3.7). But, since the conjugate func- tion of Ψ satisfies condition ∆2, H Ψ is the bidual of HM Ψ ([7], Corollary 3.3); hence {fj}j converges weakly to 0 in HM Ψ. for z ∈ Sj; hence, writing K = kfjkBΨ , one has: On the other hand, if Sj = D(1, 1 − rj ) ∩ D, one has 1 − rj z ≤ 2(1 − rj ) 1 =ZD Ψ(fj/K) dA ≥ZSj Ψ(fj/K) dA ≥ A(Sj )Ψ(xj /4K) . Since A(Sj) ≥ α(1 − rj)2, with 0 < α < 1, we get (since Ψ(αxj /4K) ≤ αΨ(xj/4K), by convexity): kfjkBΨ ≥ (α/4) xj Ψ−1(cid:0)1/(1 − rj)2(cid:1) = (α/4) Therefore JΨ is not Dunford-Pettis. On the other hand, one has: 11 Ψ−1(cid:0)1/(1 − rj)(cid:1) Ψ−1(cid:0)1/(1 − rj)2(cid:1) ≥ α 4A · (cid:3) Proposition 3.2 If JΨ is Dunford-Pettis, then JΨ is weakly compact. Proof. By Theorem 2.2, if JΨ is not weakly compact, there is a subspace X0 of H Ψ isomorphic to c0 on which JΨ is an into-isomorphism; hence JΨ cannot be Dunford-Pettis. (cid:3) We shall see in the next section that JΨ may be weakly compact without being Dunford-Pettis. 3.2 Absolutely summing Every p-summing operator is weakly compact and Dunford-Pettis; so it may be expected that JΨ is p-summing for some p < ∞. The next results show that this is never the case as soon as Ψ grows faster than all the power functions. Recall that an operator T : X → Y between two Banach spaces X and Y is called (p, q)-summing if there is a constant C > 0 such that (cid:16) nXk=1 kT xkkp(cid:17)1/p ≤ C kx∗kX∗ ≤1(cid:16) nXk=1 sup x∗(xk)q(cid:17)1/q , for every finite sequence (x1, . . . , xn) in X. If q = p, it is said p-summing. Every p-summing operator is (p, q)-summing for q ≤ p. Theorem 3.3 If JΨ : H Ψ → BΨ is p-summing, then, for every q > p, Ψ(x) = O (xq) for x large enough. Moreover, if p < 2, then JΨ is compact. In order to prove this, we need two lemmas. Lemma 3.4 If the canonical injection IΨ : A → BΨ is (p, 1)-summing, where A = A(D) is the disk algebra, then Ψ(x) = O (x2p) for x large enough. In particular, Jr : H r → Br is (p, 1)-summing for no p < r/2, and, if Ψ ∈ ∆0, then JΨ is (p, 1)-summing for no p < ∞. Recall that the disk algebra is the space of continuous functions on D which are analytic in D. We refer to [9] for a detailed study of r-summing Carleson embeddings H r → In particular, it follows from these results that Jr : H r → Br is 1- Lr(µ). summing for 1 ≤ r < 2. On the other hand, it is easy to see that J2 : H 2 → B2 for the canonical orthonormal is not Hilbert-Schmidt (i.e. not 2-summing): basis {zn}n and {√n + 1 zn}n of H 2 and B2, J2 is the diagonal operator of multiplication by {1/√n + 1}n. It also follows from [9] that, for r ≥ 2, Jr is p-summing for no finite p. Proof. Assume that we do not have Ψ(x) = O (x2p) for x large enough. Then lim supx→+∞ Ψ(x)/x2p = +∞. Given any K > 0, take y > 0 such that Ψ(y)/y2p ≥ K and such that h = 1/pΨ(y) ≤ 1/2. Let N be the integer part of (1/h) + 1. Writing ξj = e2πij/N , we set: uj(z) = h2 [1 − (1 − h) ξjz]2 · 12 We have uj ∈ A(D). By [7], Lemma 5.6, one has, since h ≥ 1/N : e2 1 − (1 − h)2N [1 − (1 − h)2][1 − (1 − h)N ]2 ≤ (1 − e)2 := C . N −1Xj=0 uj(eit) ≤ N h2 Hence: sup kx∗kA∗ ≤1 N −1Xj=0 x∗(uj) ≤ C . On the other hand, it is easy to see that uj(z) ≥ 1/9 when z−(1−h)ξj < h; 1 =ZD hence, if Sj = {z ∈ D ; z − (1 − h)ξj < h}, one has, since A(Sj) = h2: Ψ(cid:16) 1/9 kujkBΨ(cid:17) , kujkBΨ(cid:17) dA ≥ h2Ψ(cid:16) 1/9 kujkBΨ(cid:17) dA(z) ≥ZSj Ψ(cid:16) uj(z) so kujkBΨ ≥ 1/9Ψ−1(1/h2). Since y = Ψ−1(1/h2), one gets: y2p (cid:21)1/2 yp ≥ (1/9)p(cid:20) Ψ(y) ≥ kujkp BΨ ≥ (1/9)p N K 1/2 9p · N −1Xj=0 This yields that the (p, 1)-summing norm of IΨ should be greater than (cid:3) K 1/2p/9C, and, as K is arbitrary, that IΨ is not (p, 1)-summing. Remark. When IΨ : A ֒→ BΨ is p-summing, we have this shorter argument. By Pietsch's factorization theorem, this IΨ factors through H p. It follows from [7], Theorem 4.10, that α h2 ≤ ρA(h) ≤ 1/Ψ−1(A/h1/p), for some constants 0 < α < 1 and A > 0, and h small enough. That means that Ψ(x) ≤ C x2p for x large enough. Lemma 3.5 If the canonical injection IΨ : A → BΨ is 1-summing, then JΨ is compact. Proof. The canonical injection J1 : H 1 → B1 (as well as JΨ whenever Ψ ∈ ∆2) is compact. Hence we may assume that H Ψ is not H 1 and hence that Ψ(x)/x tends to +∞ as x tends to +∞. Assume that JΨ is not compact. Then, as in the proof of Proposition 3.1, there are some A > 1 and a sequence {xk}k going to +∞ such that Ψ(Axk) ≥ [Ψ(xk)]2. Setting hk = 1/Ψ(xk), we define, as in the proof of Proposition 3.4: uk,j(z) = h2 k [1 − (1 − hk)ξk,j z]2 , where ξk,j = e2πij/Nk , with Nk the integer part of (1/hk) + 1. One has uk,j ∈ A and (see the proofs of the two quoted propositions): Nk−1Xj=0 uk,j(eit) ≤ C and kuk,jkBΨ ≥ δα A 1 Ψ−1(1/hk) · 13 It follows that: Nk−1Xj=0 kuk,jkBΨ ≥ δα A Nk Ψ−1(1/hk) ≥ δα A 1/hk Ψ−1(1/hk) = δα A Ψ(xk) xk −→k→∞ +∞. Hence IΨ is not 1-summing. (cid:3) = 1 2p + 1 p (see [2], Theorem 2.22), i. e. p1 = 2 Proof of Theorem 3.3. Since JΨ : H Ψ → BΨ is p-summing and the canon- ical injection IΨ : A → BΨ factors as IΨ : A → H Ψ → BΨ, this injection is p-summing. By Lemma 3.4, Ψ(x) = O (x2p) for x large enough. Hence we have the factorization A → H 2p → H Ψ → BΨ. Since the first injection is 2p- summing and the last one is p-summing, the composition is max(1, p1)-summing, with 1 3 p. If p1 > 1, we can use p1 again Lemma 3.4 with p1 instead of 2p; we get that Ψ(x) = O (x2p1 ), for x large enough, and that the factorization IΨ : A → H 2p1 → H Ψ → BΨ is max(1, p2)- summing, with 1 p . Going on the same way, we get a decreas- p2 ing sequence {pn}n such that the canonical injection A → BΨ is max(1, pn)- summing and 2αn+1 ; hence pn −→n→∞ p · Writing pn = αnp, we get αn+1 = 2αn p/2. In particular, Ψ(x) = O (xq) for every q > p. = 1 2p1 If p < 2, one has max(1, pn) = 1 for n large enough, and Lemma 3.4 implies (cid:3) + 1 1 pn+1 = 1 2pn + 1 that JΨ is compact. Remark 1. It is not clear whether JΨ p-summing, with p ≥ 2, implies that JΨ is compact. However, when r ≥ 2, Jr : H r → Br is p-summing for no p < ∞ (see [9]). Remark 2. An operator T : X → Y between two Banach spaces is said to be finitely strictly singular (or superstrictly singular) if for every ε > 0, there is an integer Nε ≥ 1 such that, for every subspace X0 of X of dimension ≥ Nε, there is an x ∈ X0 such that kT xk ≤ ε kxk. Every finitely strictly singular operator is strictly singular. It is not difficult to see that every compact operator is finitely strictly singular and it is shown in [10] (see also [5], Corollary 2.3) that every p-summing operator is finitely strictly singular. We do not know when JΨ is finitely strictly singular. 3.3 Order boundedness Recall that an operator T : X → Y from a Banach space X into a Banach lattice Y is said to be order bounded if there is y ∈ Y+ such that T x ≤ y for every x in the unit ball of X. Since the Bergman-Orlicz space BΨ is a subspace of the Banach lattice LΨ(D,A), we may study the order boundedness of JΨ. Actually, we are going to see that the natural space for the order boundedness of JΨ is not LΨ(D,A), but the weak Orlicz space LΨ,∞(D,A), the definition of which we are recalling below (see [7], Definition 3.16). 14 Definition 3.6 Let (S,S, µ) be a measure space; the weak-LΨ space LΨ,∞ is the set of the (classes of ) measurable functions f : S → C such that, for some constant c > 0, one has, for every t > 0: µ(f > t) ≤ 1 Ψ(ct) · One has LΨ ⊆ LΨ,∞ and ([7], Proposition 3.18) the equality LΨ = LΨ,∞ implies that Ψ ∈ ∆0. On the other hand, this equality holds when Ψ grows sufficiently; for example, if Ψ satisfies the condition ∆1: xΨ(x) ≤ Ψ(αx), for some constant α > 1 and x large enough. Proposition 3.7 JΨ : H Ψ → BΨ is always order bounded into LΨ,∞(D,A). Proof. Since (see [7], Lemma 3.11): (3.1) 1 4 Ψ−1(cid:16) 1 1 − z(cid:17) ≤ sup kf kHΨ ≤1f (z) ≤ 4Ψ−1(cid:16) 1 1 − z(cid:17) , one has, denoting by S(z) the supremum in (3.1), for t large enough: A(S > t) ≤ A(cid:0){z ∈ D ; z > 1 − 1/Ψ(t/4)}(cid:1) ≤ 2 Ψ(t/4) ≤ 1 Ψ(t/8) , and the result follows. Since we also have, for t large enough: (cid:3) A(S > t) ≥ A(cid:0){z ∈ D ; z > 1 − 1/Ψ(4t)}(cid:1) ≥ 1 Ψ(4t) , we get: Corollary 3.8 JΨ is order bounded into LΨ(D,A) if and only if LΨ = LΨ,∞. This is the case if Ψ ∈ ∆1. Remark. Contrary to the compactness, or the weak compactness, which re- quires that Ψ does not grow too fast, the order boundedness of JΨ into LΨ(D,A) holds when Ψ grows fast enough. Nevertheless, for Ψ(x) = exp[(cid:0) log(x+1)(cid:1)2 ]−1, JΨ is compact and order bounded into LΨ(D,A). When JΨ is weakly compact, JΨ maps H Ψ into BM Ψ (Theorem 2.2); hence, we may ask whether JΨ may be order bounded into M Ψ(D,A); however, we have: Proposition 3.9 JΨ is never order bounded into M Ψ(D,A). Proof. If it were the case, we should have S ∈ M Ψ(D,A), and hence ZD Ψ(cid:20)4 × 1 4 Ψ−1(cid:16) 1 1 − z(cid:17)(cid:21) dA(z) < +∞ , which is false. (cid:3) 15 4 An example Theorem 4.1 There exists an Orlicz function Ψ such that JΨ is weakly compact and Dunford-Pettis, but which is not compact. Note that such an Orlicz function is very irregular: Ψ /∈ ∆2, Ψ /∈ ∆0, so, for every A > 1, Ψ(Ax)/Ψ(x) is not non-decreasing for x large enough, and the conjugate function of Ψ does not satisfies condition ∆2. The following lemma is undoubtedly well-known, but we have found no ref- erence, so we shall give a proof. Recall that a sublattice X of L0(µ) is solid if f ≤ g and g ∈ X implies f ∈ X and kfk ≤ kgk. Lemma 4.2 Let (S,S, µ) be a measure space, and let X be a solid Banach sublattice of L0(µ), the space of all measurable functions. Then, for every weakly null sequence {fn}n in X and every sequence {An}n of disjoint measurables sets, the sequence {fn1IAn}n converges weakly to 0 in X. Proof. If the conclusion does not hold, there are a continuous linear functional σ : X → C and some δ > 0 such that, up to taking a subsequence, σ(fn1IAn ) ≥ δ. Set, for every measurable set A ∈ S: µn(A) = σ(fn1IA) . Then µn is a finitely additive measure with bounded variation. By Rosenthal's lemma (see [3], Lemma I.4.1, page 18, or [1], Chapter VII, page 82), there is an increasing sequence of integers {nk}k such that: (cid:12)(cid:12)(cid:12)µnk(cid:16)[l6=k Anl(cid:17)(cid:12)(cid:12)(cid:12) ≤ µnk(cid:16)[l6=k Now, if A =Sl≥1 Anl , {fnk1IA}k is weakly null, but: ) − µnk(cid:16)[l6=k σ(fnk 1IA) ≥ σ(fnk 1IAnk Anl(cid:17) ≤ δ/2 . Anl(cid:17) ≥ δ − δ 2 = δ 2 , so we get a contradiction. Proof of Theorem 4.1. We begin by defining a sequence {xn}n of positive numbers in the following way: set x1 = 4 and, for every n ≥ 1, xn+1 > 2xn is the abscissa of the second intersection point of the parabola y = x2 with the straight line containing (xn, x2 n − 2xn (for example, x2 = 56). Define Ψ : [0, +∞) → [0, +∞) by Ψ(x) = 4x for 0 ≤ x ≤ 4, and, for n ≥ 1: (4.1) Ψ(xn) = x2 n , Ψ affine between xn and xn+1 . n); we have xn+1 = x3 n) and (2xn, x4 Ψ(2xn) = x4 n , (cid:3) Then Ψ is an Orlicz function and (4.2) x2 ≤ Ψ(x) ≤ x4 for x ≥ 4. 16 For this Orlicz function Ψ, JΨ is not compact, since Ψ(2x)/[Ψ(x)]2 does not tend to 0. However, JΨ is weakly compact, because one has the factorization H Ψ ֒→ H 2 ֒→ B4 ֒→ BΨ (by (4.2) and Lemma 2.4). Assume that JΨ is not Dunford-Pettis: there exists a weakly null sequence {fn}n in the unit ball of H Ψ which does not converges for the norm in BΨ. Then {fn}n converges uniformly to 0 on the compact subsets of D (since it is weakly null) and we may assume that kfnkBΨ ≥ δ for some δ > 0. We may also assume that kfnk∞ −→n→∞ +∞ because if {fn}n were uniformly bounded, we should have kfnkBΨ −→n→∞ We are going to show that there exist a subsequence {fnk}k and pairwise dis- joint measurable sets Ak ⊆ T such that the sequence {fnk 1IAk}k ⊆ LΨ(T, m) is equivalent to the canonical basis of ℓ1, whence a contradiction with Lemma 4.2. It is worth to note from now that the Poisson integral P maps boundedly 0 and the canonical injection is L2(T) into L4(D). bounded from H 2 into B4, by Lemma 2.4. Indeed, L2(T) = H 2 ⊕ H 2 0, by dominated convergence. We have seen in the proof of Lemma 2.5 that there exist a subsequence {fnk}k and disjoint measurable annuli C1 = {z ∈ D ; z ≤ r1} and Ck = {z ∈ D ; rk−1 < z ≤ rk}, k ≥ 2, with 0 < r1 < r2 < ··· < rn < ··· < 1, such that kfnk 1ICkkLΨ(D) ≥ δ/2. The assumptions of that lemma are satisfied here: kfnkHΨ ≤ 1, kfnkBΨ ≥ δ, {fn}n converges uniformly to 0 on the compact subsets of D, and fn ∈ BM Ψ because H Ψ ⊆ BM Ψ, since JΨ is weakly compact. Then: Fact 1. There exist two sequences {αk}k and {βk}k, with βn > αn −→n→∞ such that, if gk = f ∗ ≤βk}, then: +∞ nk 1I{αk≤f ∗ nk kP(gk)kLΨ(D) ≥ δ/3 , where f ∗ nk is the boundary value of fnk on T. nk 1I{f ∗ nk 12 Ψ−1(cid:0)1/A(Ck)(cid:1) and vk = P(cid:0)f ∗ <αk}(cid:1) 1ICk . One Ψ(cid:0)vk/(δ/12)(cid:1) dA ≤ Ψ(cid:0)αk/(δ/12)(cid:1)A(Ck) = 1 , nk ) 1ICkkLΨ(D) = nk ) = fnk , we have kP(f ∗ Proof. 1) Let αk = δ has: Ψ(cid:0)vk/(δ/12)(cid:1) dA =ZCk ZD so kvkkLΨ(D) ≤ δ/12. Since P(f ∗ kfnk 1ICkkLΨ(D) ≥ δ/2, and we get: kP(f ∗ nk1I{f ∗ nk ≥αk}) 1ICkkLΨ(D) ≥ kfnk 1ICkkLΨ(D) − kvkkLΨ(D) ≥ 5δ 12 · ≥αk}. Since P(wk 1I{wk>β}) tends to 0 uniformly on Ck when β goes to infinity, Lebesgue's dominated convergence theorem gives: 2) Let wk = f ∗ nk 1I{f ∗ nk δ 2 − δ 12 = kP(wk 1I{wk>β}) 1ICkkLΨ(D) ≤ kP(wk 1I{wk>β}) 1ICkkL4(D) −→β→+∞ 0 , 17 so there is some βk > αk such that kP(wk 1I{wk>β}) 1ICkkLΨ(D) ≤ δ/12. We then have, with gk = f ∗ nk 1I{αk≤f ∗ nk ≤βk}: kP(gk)kLΨ(D) ≥ kP(gk) 1ICkkLΨ(D) ≥ 5δ 12 − δ 12 = δ 3 , and that ends the proof of Fact 1. (cid:3) Fact 2. There are a further subsequence, denoted yet by {fnk}k, and pairwise disjoint measurable subsets Ek ⊆ {αk ≤ f ∗ nk 1IEk , then: nk ≤ βk}, such that, if hk = f ∗ kP(hk)kLΨ(D) ≥ δ/4 . Proof. First, since gk ∈ L∞(T) ⊆ M Ψ(T), there exists εk > 0 such that m(A) ≤ εk implies kgk 1IAkLΨ(T) ≤ δ/(12 kPk) (where kPk stands for the norm of P : L2(T) → L4(D)). Now, P : LΨ(T) → LΨ(D) is bounded and its norm is ≤ kPk, thanks to the factorization LΨ(T) ֒→ L2(T) ֒→ L4(D) ֒→ LΨ(D). Hence kP(gk 1IA)kLΨ(D) ≤ δ/12 for m(A) ≤ εk. Let Bk = {αk ≤ f ∗ nk ≤ βk}. Up to taking a subsequence, we may assume that Pl>k m(Bl) ≤ εk. Let Ek = Bk \ [l>k The sets Ek, k ≥ 1, are pairwise disjoint, and kP(gk 1IEk )kLΨ(D) ≥ kP(gk 1IBk )kLΨ(D) − kP(cid:0)gk 1ISl>k Bl(cid:1)kLΨ(D) ≥ Bl . so we get the Fact 2 with hk = gk 1IEk = f ∗ nk 1IEk . δ 3 − δ 12 = δ 4 ; (cid:3) Set nk (z)(cid:1) ≤ M f ∗ Fk = {z ∈ Ek ; Ψ(cid:0)f ∗ For z ∈ Ek \ Fk, one has: ZEk\Fk f ∗ M ZT Ψ(f ∗ nk2 dm ≤ nk 1IEk\FkkL2(T) ≤ 1/√M and: 1 so kf ∗ nk (z)2} . nk ) dm ≤ 1 M , kP(f ∗ nk 1IEk\Fk )kLΨ(D) ≤ kP(f ∗ nk 1IEk\Fk )kL4(D) nk 1IEk\Fk )kL2(T) ≤ kPk√M ≤ δ 8 , ≤ kPk k(f ∗ for M large enough. It follows that, for M large enough, kP(f ∗ δ/8 and nk 1IFk )kLΨ(D) ≥ (4.3) kf ∗ nk 1IFkkLΨ(D) ≥ δ/(8 kPk) . 18 Now, we may assume that, for some α > 0, ZT f ∗ nk2 1IFk dm ≥ α , because, if not, there would be a subsequence {f ∗ L2(T); but then {P(cid:0)fnkj contrary to (4.3). It follows, using (4.2), that: 1IFkj }j converging to 0 in 1IFkj(cid:1)}j would converge to 0 in B4, and hence in BΨ, nkj (4.4) ZFk Ψ(f ∗ nk) dm ≥ α . The following lemma is now the key of the proof. Lemma 4.3 Let δn = 2xn−1/xn = 2/(x2 n−1 − 2). If Ψ(x) ≤ M x2 and x ≥ xn, then, for n large enough (n ≥ N ), one has Ψ(εx) ≥ CM ε Ψ(x) for δn ≤ ε ≤ 1. Proof. We may assume that xn ≤ x < xn+1, because if xk ≤ x < xk+1 with k ≥ n, then ε ≥ δn implies ε ≥ δk. Now, remark that: , Ψ(y) y x (4.5) Ψ(y) for 2xn ≤ x ≤ y ≤ xn+1 . Ψ(x) ≤ 4 Indeed, on the one hand, Ψ(y)−Ψ(xn) Ψ(x)−Ψ(xn) = y−xn x−xn ≤ y hand, Ψ(y) − Ψ(xn) ≥ Ψ(y) − Ψ(y/2) ≥ Ψ(y) − 1 Ψ(x)−Ψ(xn) ≤ 2 Ψ(y)−Ψ(xn) Ψ(x)−Ψ(xn) ≤ 4 y x · We shall separate three cases: 1) εx ≤ xn ≤ x ≤ 2xn. Then εx ≥ εxn and hence Ψ(εx) ≥ Ψ(εxn). But 2xn−1 ≤ εxn ≤ xn, since ε ≥ δn; hence (4.5) implies that Ψ(εx) ≥ (ε/4) Ψ(xn) = (ε/4) x2 n. On the other hand, one has, by hypothesis, Ψ(x) ≤ M x2 ≤ M (2xn)2, so we get Ψ(εx) ≥ (ε/16M )Ψ(x). x ; and, on the other Ψ(x) ≤ x/2 = 2 y 2 Ψ(y) = 1 2 Ψ(y) , so Ψ(y) 2) xn ≤ εx ≤ x ≤ 2xn. Then, since 1 ≤ 1/ε: M (2xn)2 Ψ(x) Ψ(εx) ≤ M x2 Ψ(xn) ≤ = 4M ≤ 4M ε · x2 n n/√M . 3) For x ≥ 2xn, remark that the conditions Ψ(x) ≤ M x2 and x ≥ 2xn imply n, and the n ≤ M x2, i.e. x ≥ x2 n/√M ≥ δnx2 n/√M = √M . Hence (4.5) gives, for 2xn ≤ x < xn+1 Indeed, if x ≥ 2xn, then Ψ(x) ≥ Ψ(2xn) = x4 n/√M = 2(xn−1/xn)x2 that x ≥ x2 condition Ψ(x) ≤ M x2 implies x4 In this case, one has εx ≥ εx2 2xn−1xn/√M ≥ 2xn, if xn−1 ≥ (since then 2xn ≤ εx ≤ x < xn+1): Ψ(x) Ψ(εx) ≤ 4 x εx = 4 ε · n/√M . That ends the proof of Lemma 4.3. (cid:3) 19 Extract now a further subsequence of {fnk}, yet denoted by {fnk}, in order nk (z)), z ∈ Fk, that (see Fact 1) αk ≥ xN +k. Lemma 4.3 holds, with x = Ψ(f ∗ for every k ≥ 1; one has (since, by definition, Ψ(fnk) ≤ M fnk2 on Fk): ZFk Ψ(ε f ∗ nk) dm ≥ ε C/α := c ε , for δN +k ≤ ε ≤ 1 . The proof of Theorem 4.1 reaches now its end: put uk = f ∗ take an arbitrary sequence of complex numbers such that Pk≥1 λk = 1. Let δ0 =Pk≥N δk. One has δ0 < 1, because we may assume that N had been taken large enough. One gets: nk 1IFk , and ZT Ψ(cid:16)(cid:12)(cid:12)(cid:12)Xk≥1 λkuk(cid:12)(cid:12)(cid:12)(cid:17) dm =Xk≥1 ZFk ≥ Xλk≥δN +k ≥ Xλk≥δN +k ≥ c(cid:16)1 − Xk≥N Ψ(λkfnk) dm ZFk cλk + Xλk<δN +k cλk = c(cid:16)1 − Xλk<δN +k δk(cid:17) = c (1 − δ0) := c0 . Ψ(λkfnk) dm λk(cid:17) Since c0 < 1, this implies, by convexity, that (cid:13)(cid:13)(cid:13)Xk≥1 λkuk(cid:13)(cid:13)(cid:13)LΨ(T) ≥ c0 . Hence {uk}k is equivalent to the canonical basis of ℓ1, and that achieves the proof of Theorem 4.1. (cid:3) Remarks. 1) It follows from Theorem 3.3 that, for this Ψ, JΨ is not p-summing for p < 4. By modifying the definition of Ψ (taking Ψ(xn) = xr/2 and Ψ(2xn) = xr n), we get, for every 4 ≤ r < ∞, an Orlicz function Ψ such that JΨ is Dunford- Pettis and weakly compact, without being p-summing for p < r, and without being compact. We do not know whether it is possible to have JΨ p-summing for no finite p. n 2) Let us point out that the fact that JΨ is Dunford-Pettis does not trivially follows from its weak compactness: H Ψ does not have the Dunford-Pettis prop- erty. In fact, if it were the case, the weakly compact injection H Ψ ֒→ H 2 would be Dunford-Pettis, and hence also H 4 ֒→ H 2 (since H 4 ֒→ H Ψ ֒→ H 2). But it is not the case: the sequence {zn}n converges weakly to 0 in H 4, whereas it does not converges in norm to 0 in H 2. 20 Proposition 4.4 There is an Orlicz function Ψ for which JΨ is weakly compact, but not Dunford-Pettis. Proof. Let us call Ψ0 the Orlicz function constructed in Theorem 4.1, and let Ψ(x) = Ψ0(x2). Then, with β = 2, Ψ(βx) = Ψ0(4x2) ≥ 4Ψ0(x2) = (2β)Ψ(x); that means that the conjugate function of Ψ satisfies ∆2. JΨ is weakly compact (since JΨ factors as H Ψ ֒→ H 4 ֒→ B8 ֒→ BΨ), but is not compact, since [Ψ(√xn)]2 = Ψ(√2√xn). Since the conjugate function satisfies ∆2, JΨ is not Dunford-Pettis, by Proposition 3.1. (cid:3) References [1] J. Diestel, Sequences and Series in Banach Spaces, Graduate Texts in Math- ematics 92, Springer-Verlag, New York (1984). [2] J. Diestel, H. Jarchow, and A. Tonge, Absolutely Summing Operators, Cambridge Studies in Adv. Math. 43, Cambridge Univ. Press (1995). [3] J. Diestel and J. J., Jr. Uhl, Vector Measures, Mathematical Surveys, No. 15, American Mathematical Society, Providence, R.I. (1977). [4] P. Duren and A. Schuster, Bergman Spaces, Math. Surveys and Mono- graphs 100, Amer. Math. Soc. (2004). [5] P. Lefèvre, When strict singularity of operators coincides with weak com- pactness, to appear in J. Operator Theory. [6] P. Lefèvre, D. Li, H. Queffélec, and L. Rodríguez-Piazza, A criterion of weak compactness for operators on subspaces of Orlicz spaces, J. Funct. Spaces and Applications 6, No. 3 (2008), 277 -- 292. [7] P. Lefèvre, D. Li, H. Queffélec, and L. Rodríguez-Piazza, Composition op- erators on Hardy-Orlicz spaces, preprint, math.FA/0610905, to appear in Memoirs Amer. Math. Soc. (2010), DOI: 10.1090/S0065-9266-10-00580-6. [8] P. Lefèvre, D. Li, H. Queffélec, and L. Rodríguez-Piazza, Nevanlinna count- ing function and Carleson function of analytic maps, preprint, arXiv : 0904.2496, hal-00375955. [9] P. Lefèvre and L. Rodríguez-Piazza, Absolutely summing Carleson embed- dings, in preparation [10] A. Plichko, Superstrictly singular and superstrictly cosingular operators, Functional analysis and its applications, 239 -- 255, North-Holland Math. Stud., 197, Elsevier, Amsterdam (2004). 21 Pascal Lefèvre, Univ Lille Nord de France F-59 000 LILLE, FRANCE UArtois, Laboratoire de Mathématiques de Lens EA 2462, Fédération CNRS Nord-Pas-de-Calais FR 2956, F-62 300 LENS, FRANCE [email protected] Daniel Li, Univ Lille Nord de France F-59 000 LILLE, FRANCE UArtois, Laboratoire de Mathématiques de Lens EA 2462, Fédération CNRS Nord-Pas-de-Calais FR 2956, Faculté des Sciences Jean Perrin, Rue Jean Souvraz, S.P. 18, F-62 300 LENS, FRANCE [email protected] Hervé Queffélec, Univ Lille Nord de France F-59 000 LILLE, FRANCE USTL, Laboratoire Paul Painlevé U.M.R. CNRS 8524, F-59 655 VILLENEUVE D'ASCQ Cedex, FRANCE [email protected] Luis Rodríguez-Piazza, Universidad de Sevilla, Facultad de Matemáticas, Departamento de Análisis Matemático, Apartado de Correos 1160, 41 080 SEVILLA, SPAIN [email protected] 22
1610.07739
1
1610
2016-10-25T06:08:28
A classification of continuous wavelet transforms in dimension three
[ "math.FA" ]
This paper presents a full catalogue, up to conjugacy and subgroups of finite index, of all matrix groups $H < {\rm GL}(3,\mathbb{R})$ that give rise to a continuous wavelet transform with associated irreducible quasi-regular representation. For each group in this class, coorbit theory allows to consistently define spaces of sparse signals, and to construct atomic decompositions converging simultaneously in a whole range of these spaces. As an application of the classification, we investigate the existence of compactly supported admissible vectors and atoms for the groups.
math.FA
math
A CLASSIFICATION OF CONTINUOUS WAVELET TRANSFORMS IN DIMENSION THREE 6 1 0 2 t c O 5 2 ] BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Abstract. This paper presents a full catalogue, up to conjugacy and subgroups of finite index, of all matrix groups H < GL(3, R) that give rise to a continuous wavelet transform with associated irreducible quasi-regular representation. For each group in this class, coorbit theory allows to consistently define spaces of sparse signals, and to construct atomic decompositions converging simultaneously in a whole range of these spaces. As an application of the classification, we investigate the existence of compactly supported admissible vectors and atoms for the groups. . A F h t a m [ 1 v 9 3 7 7 0 . 0 1 6 1 : v i X r a 1. Introduction Continuous wavelet transforms in higher dimensions have been studied since [34]. The construction departs from the choice of a closed matrix group H < GL(d, R). The affine group generated by H and the group of translations is the semi-direct product group G = Rd × H. This group acts unitarily on functions f ∈ L2(R) by the quasi-regular representation π(x, h)f (y) = det(h)−1/2f (h−1(y − x)) . Given any pair of vectors ψ, f ∈ L2(Rd), the continuous wavelet transform of f with respect to ψ is the function Wψf : G → C , (x, h) 7→ hf, π(x, h)ψi . The continous wavelet transform associated to ψ is the linear operator Wψ : f 7→ Wψf , and ψ is called a wavelet or analyzing vector. We call ψ admissible if Wψ is a nonzero scalar multiple of an isometry into L2(G), i.e., if the equality kf k2 2 = 1 cψZHZRd Wψf (x, h)2 dx dh det(h) f = 1 cψZHZRd Wψf (x, h) π(x, h)ψ dx dh det(h) holds for all f ∈ L2(Rd). This isometry property allows to invert the wavelet transform by its adjoint, which leads to the weak-sense inversion formula We call H admissible if there exist admissible vectors ψ ∈ L2(Rd), and irreducibly ad- missible if the quasi-regular representation is in addition irreducible. Admissible groups have been studied extensively, see [4, 17, 31, 29, 24, 20, 21, 11, 5, 3] for a small sample of the existing literature. Nonabelian versions, where Rd is replaced by a suitably cho- sen noncommutative Lie group, were also considered [7, 8, 27, 36]. A particular class of irreducibly admissible matrix groups gained particular attention, the so-called shearlet group acting in dimension two [11] and its higher-dimensional relatives [12, 10, 9, 25]. 1 2 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA The interest of irreducibly admissible dilation groups is strengthened by the realization that coorbit space theory, as developed in [14, 15, 16], applies to these groups and their associated continuous wavelet transform. This observation had been made early on for the similitude groups in [15], was later shown for the shearlet cases in [11, 12, 10], and finally established in full generality in [22]. Since then, the theory of wavelet coorbit spaces has been further developed in [23, 25, 26, 38]; see also [13] for an overview of the recent developments and their historical roots. Coorbit space theory provides a unified approach to wavelet approximation theory, that allows to consistently define function spaces in terms of wavelet coefficient decay. Here consistency refers to the fact that for functions in these spaces, the fast wavelet coefficient decay is guaranteed for all choices of analyzing vectors taken from a well-understood class of nice wavelets. This property emphasizes that the approximation-theoretic properties of group-theoretically defined wavelet systems ultimately depend on the way the group G generates the system from a single mother wavelet, rather than on the specific choice of the mother wavelet. In other words, it puts the focus of the analysis of wavelet system on the role of the dilation group H. It is the main purpose of this paper to provide an overview of the available choices of H in dimension three. 2. Aims and main result of this paper In this paper we set out to determine all irreducibly admissible matrix groups H < GL(3, R) up to conjugacy. In effect, we will only classify the connected components H0 < H of the identity in H. However, from this subgroup to the full group is only a small step: By [18, Remark 4], H0 is an open subgroup of finite index. There are several motivations for studying such a classification. The chief interest of ob- taining a full list of candidate groups is to get an overview of the variety of available group actions. The three-dimensional case is arguably of interest to applications, due to the in- creasing use made of voxel data provided by imaging applications such as computerized tomography or MRI. Judging by the existing digital implementations of wavelet-like trans- forms such as curvelets and shearlets in the three-dimensional setting [6, 35], the task of numerically implementing a given continuous wavelet transform in dimension three seems challenging, but not insurmountable. Hence we believe that the list of available options presented below is potentially useful in applications. In such a description of all possible cases, it is however advisable to use suitable equivalence relations as a means of reducing the complexity of the general picture. In our context, conjugacy is the natural choice for such an equivalence relation: If two matrix groups H1, H2 are conjugate, say H1 = gH2g−1 for some invertible matrix g, then it is very straightforward to relate the associated wavelet systems via a dilation by g. Indeed, if ψ2 is an admissible vector for H2, and f2 ∈ L2(G) is arbitrary, then defining ψ1(x) = det(g)−1/2ψ2(g−1(x) , f1(x) = det(g)−1/2f2(g−1(x) we obtain for hi ∈ Hi(i = 1, 2) related via h1 = gh2g−1 and x ∈ Rd that W H1 ψ1 f1(x, h1) = W H2 ψ2 (g−1x, h2) . Here we temporarily used H1, H2 as superscripts to differentiate the wavelet transforms associated to H1, H2, respectively. Hence one can consider the wavelet transforms over CLASSIFYING WAVELET TRANSFORMS IN 3D 3 H1 and H2 as instances of the same transform, viewed with respect to different linear coordinate systems on Rd. In particular, any property of interest pertaining to wavelets arising from the action of H1 will translate to an analogous property of wavelets arising from H2, and it is clearly sufficient to study just one of the groups. More generally speaking, for most questions of interest it will be sufficient to study just one representative from each full conjugacy class. Our chief example of a property that can be checked by focusing on the representatives modulo conjugacy is the question whether there exist compactly supported admissible vectors and atoms for a given wavelet transform. It is currently not known whether every irreducibly admissible matrix group has compactly supported admissible vectors, and for the much smaller class of atoms (also called "better vectors" in the coorbit scheme [15]), this existence question is even harder to answer. We prove, via a case-by-case treatment of all possible candidates, that every irreducibly admissible matrix group in dimension three has compactly supported admissible vectors, and that the overwhelming majority of cases even allows the existence of compactly supported atoms. In the context of generalized wavelet analysis, some classification results are already known. For instance, the analogous search for all irreducibly admissible H < GL(2, R) groups in dimension two (or at least, of their connected components) produces the fol- lowing list, complete up to conjugacy [19]: • The diagonal group, given by H =(cid:26)(cid:20) a 0 H =(cid:26)(cid:20) a H =(cid:26)(cid:20) a 0 b (cid:21) : a > 0, b > 0(cid:27) ; −b a (cid:21) : a2 + b2 6= 0(cid:27) ; 0 ac (cid:21) : a > 0, b ∈ R(cid:27) . b b • the similitude group • and the shearlet groups defined as Here c ∈ R can be chosen arbitrarily. Existence of atoms for each of these cases (and thus for all irreducibly admissible matrix groups acting in dimension two) was shown in [23]. The analytic properties of the asso- ciated wavelet transforms were explored in the thesis [33]. Further (partial) classification results exist for abelian irreducibly admissible groups in arbitrary dimensions [18], and for the class of generalized shearlet dilation groups [25] . In a similar spirit to our paper, the papers [1, 2] classified a class of subgroups of the metaplectic group that allow an inversion formula under the metaplectic representation. We are not aware of further sources deal- ing with the systematical construction and classification of irreducibly admissible matrix groups or certain subclasses thereof. The following theorem summarizes the results of our paper. Throughout the paper, we use R∗ = R \ {0}. The terminology appearing in part (c) will be introduced in Section 6. Theorem 2.1. Let H < GL(3, R) be irreducibly admissible. (a) Its connected component is conjugate to precisely one group from the following list: 4 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA (1) The abelian groups. H0 = a1 0 0 a2 0 0 0 a3   : ai > 0 . 0 a 0 b cos t −b sin t 0 b sin t . . 0 a b 0 0 a 0 b cos t  : a > 0, b > 0, t ∈ R 0 0 c  : a > 0, c > 0, b ∈ R 0 0 a  : a > 0, b, c ∈ R  0 0 a  : a > 0, b, c ∈ R  a b c 0 a 0 a b c 0 a b . . (2) The nonabelian, solvable groups. eas cos(bs) −eas sin(bs) eas sin(bs) with (a, b) ∈ R × R∗. (1.a) (1.b) (1.c) (1.d) (1.e) (2.a) (2.b) (2.c) H =  H0 =  H0 = H0 = H0 =  H0 = H0 =  with a ∈ R∗. eas 0 0  es 0 0 easa 0 0 t1 t1 t1 , t2 t2 eas cos(bs) : s, t1, t2 ∈ R cos(r)  : r, s, t1, t2 ∈ R es cos(r) : r, s, t1, t2 ∈ R t2 , . cos(r) − sin(r) sin(r) es cos(r) −es sin(r) es sin(r) CLASSIFYING WAVELET TRANSFORMS IN 3D 5 (2.d) (2.e) (2.f) (2.g) (2.h) aλ 0 0 with λ ∈ R∗. with λ ∈ R∗. with λ ∈ R∗. H0 =  H0 =  H0 = exp h = H0 = exp h =  H0 =  H0 =  H0 =  H0 =  aλ 0 0 with λ ∈ R∗. aλ 0 0 0 0 (2.i) (2.j) (2.k) aλ−1 t1 0 aλ+1bc ab 0 0 aλ 0 0 t1 aλ 0 t1 aλ−c 0 aλ 0 0  aλ 0 0 aλ−1 t1 0 , , , , t2 t2 aλ t2 0 t2 0 t1 aλ 0 t1 + aλ ln a t1 aλ aλ ln a 0 aλ−1 : a > 0, t1, t2 ∈ R aλ−1 : a > 0, t1, t2 ∈ R aλ  : a > 0, t1, t2 ∈ R  : a > 0, t1, t2 ∈ R aλ−1  : a > 0, t1, t2 ∈ R aλ−2  : a > 0, t1, t2 ∈ R b : a > 0, b > 0, t ∈ R 0 aλ : a > 0, b > 0, t ∈ R 0 aλbc 0 δaλ−1 ln a t1aλ−2 t 0 0 0 t b t2 t2 . , with (λ, c) ∈ R × ([−1, 1] \ {0}). with (λ, δ) ∈ R∗ × R. with (λ, c) ∈ R∗ × (R \ {1}). 6 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA with λ ∈ R∗. with (ν1, ν2) ∈ R∗ × R. (2.l) (2.m) (2.n) b 0 ab 0 0 H0 =  H0 =  H0 =  ab 0 b 0 0 0 with λ ∈ R∗. t −ν1 ln a + ν2b ln b 0 b  : a > 0, b > 0, t ∈ R . ab 0 0 . t b 0 0 0 bλ : a > 0, b > 0, t ∈ R  : a > 0, b > 0, t ∈ R b t δ1 ln a + δ2b ln b . with (δ1, δ2) ∈ R2 \ (0, 0). (3) The nonsolvable groups. (3.a) H = R+ × SO(3). Every group H0 appearing in this list is contained in an irreducibly admissible matrix group H. (b) H has compactly supported admissible vectors. (c) If the connected component of H is not conjugate to one of the groups from case (2.l), then H has compactly supported atoms ψ ∈ Bv0, for any polynomially bounded weight v0 on G = R3 ⋊ H. The classification of the solvable cases is achieved in Section 4, whereas the classification of the non-solvable cases can be found in Section 5. Parts (b) and (c) are proved in Section 6. We count 7 isolated cases (among them the 5 abelian ones), 7 one-parameter and 6 two- parameter families, with the overwhelming majority of examples provided by the solvable cases. The cases (1.e), (2.d) and (2.e) correspond to the standard shearlet groups [11], whereas the case (2.i) contains the Toeplitz shearlet group [10]. We note that some of the parametrized families given above can be merged, e.g. (1.e), (2.d) and (2.e), which would lead to a somewhat smaller list of cases. However, the current list reflects the construction method for the different cases and thus enables an easier tracing of the argument proving the classification, which is why we have refrained from streamlining the presentation in the theorem. 3. Preliminaries 3.1. Properties of irreducibly admissible matrix groups. Our classification of irre- ducibly admissible matrix groups H departs from a characterization of said groups, that uses the dual action of H. This group action is given by the mapping H × Rd ∋ (h, ξ) 7→ CLASSIFYING WAVELET TRANSFORMS IN 3D 7 h−T ξ, where h−T denotes the inverse of the transpose of h. Now irreducibly admissible matrix groups are characterized as follows [21], [20, Corollary 5.15]. Theorem 3.1. Let H < GL(d, R) denote a closed matrix group. (a) H is irreducibly admissible if and only if there exists an open, conull orbit O ⊂ Rd with the additional property that for all ξ ∈ O, the associated stabilizer Hξ = {h ∈ H : h−T ξ = ξ} is compact. (b) If H is admissible, then G = Rd ⋊ H is nonunimodular. Note that the condition in part (b) is equivalent to the statement that the functions ∆H and det do not coincide; here ∆H is the modular function of H. By [18, Remark 4], property (a) implies that the connected component H0 < H acts with open orbits and (almost everywhere) compact stabilizers. Furthermore, the fact that the stabilizers in H associated to the open orbits are compact implies that H/H0 must be finite. Also by [18, Remark 4], the dimension of admissible matrix groups H < GL(d, R) lies between d and d(d+1)/2, which for d = 3 means that dim(H) ∈ {3, 4}. To summarize, our search in the following concentrates on connected matrix groups H0 < GL(3, R) with open orbits and associated compact stabilizers; necessarily of dimension 3 or 4. Our proof strategy consists in carefully studying the dual action of connected groups H < GL(3, R), and to extract conjugation-invariant features of this action that allow to systematically separate the general class into a various subcases giving rise to the classification in Theorem 2.1. For this purpose, we rely on various notions from differential geometry and Lie group theory. These are introduced next. 3.2. Linear actions of Lie groups. Let V be a finite-dimensional vector space over R, and let H be a locally compact topological group. A representation of H in V is a continuous homomorphism α : H → GL(V); a representation of H can be regarded as a linear action of H on V. If W is the complexification of V, then α defines a linear action of H on W in the obvious way. Given a representation α of H in V , the Cartesian product V × H is endowed with the semi-direct product operation defined by (x, h) · (x′, h′) = (x + α(h)x′, hh′) and the resulting group is denoted by V ⋊α H. Let α : H → GL(V) and β : K → GL(W) be representations of groups H and K. We say that α and β are isomorphic if there is a group isomorphism c : H → K and a vector space isomorphism g : V → W, such that β(c(h)) ◦ g = g ◦ α(h). Note that in this situation (v, h) 7→ (gv, c(h)) is an isomorphism of the corresponding semi-direct products. Given a representation α, its dual representation α∗ : H → GL(V ∗) is defined by α∗(h)f = f ◦ α(h)−1. If α and β are isomorphic representations, then α∗ and β∗ are isomorphic. The groups of interest in this paper are matrix groups. We identify an element X ∈ gl(d, R) with a d × d real matrix in the usual canonical way. With the canonical inclusion Rd ⊂ Cd, we have gl(d, R) ⊂ gl(d, C). The conjugacy class of an element X ∈ gl(d, R) is 8 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA the set {sXs−1 : s ∈ GL(d, C)}; note that X ′ belongs to the conjugacy class of X if and only if X ′ is the matrix representation of X with respect to some basis of Cn. If V = Rd, then α = idH is called the standard Let H be a subgroup of GL(d, R). representation of H. The representation of interest in this paper is the dual of the standard representation: identifying Rd with its dual via the standard inner product, the dual of the standard representation is given by h 7→ (hT )−1, and is called the transpose representation. If H and K are subgroups of GL(d, R), then the standard representations of H and K, or their duals, are isomorphic if and only if H and K are conjugate. If H is a closed subgroup of GL(d, R), then H is a Lie subgroup of GL(d, R). For the present purposes, this means that each of the following hold. (1) There is a subspace h of the space gl(d, R) of all endomorphisms of Rd that is closed under the usual bracket operation [X, Y ] = XY − Y X, and such that where h = {X ∈ gl(d, R) : exp RX ⊂ H} exp X = I + X + 1 2! X 2 + 1 3! X 3 + · · · ; h is called the Lie algebra of H. (2) If n = dim h, then there is an open neighborhood U of the identity 1 in H, and an open neighborhood V of 0 ∈ h, such that exp V : V → U is a homeomorphism. More generally, if K is a closed subgroup of H with Lie algebra k, then there is an open neighborhood U of K in the quotient space H/K, and an open neighborhood V of a complementary space s for k in h, such that the map X 7→ exp XK is a homeomorphism of V onto U. (3) If α : H → GL(V) is a representation of H, then there is a unique Lie algebra homomorphism dα : h → gl(V) such that α(exp X) = exp dα(X) holds for all X ∈ h. One has dα(X) = , X ∈ h d dt α(exp tX)(cid:12)(cid:12)(cid:12)t=0 and the Lie algebra of ker α is the kernel of dα in h. A Lie algebra homomorphism of h into gl(V) is called a Lie algebra representation of h. Just as a representation of H is a continuous linear group action of H, a Lie algebra homomorphism of h into gl(V) is said to define a linear action of the Lie algebra h on V. If H and K are closed subgroups of GL(d, R) and c : H → K is a continuous isomorphism, then dc is an isomorphism of their Lie algebras, and if λ is a linear form on h then λc := λ(dc−1·) is a linear form on k and λ 7→ λc is a vector space isomorphism of h∗ and k∗. In particular if K = sHs−1 for some s ∈ GL(n, R), then the Lie algebras h and k of H and K satisfy k = shs−1. From now on H will denote a closed subgroup of GL(d, R) with Lie algebra h, acting on Rd (and Cd) by the transpose representation. A subspace Z of Rd is H-invariant if hT Z = Z for all h ∈ H, and an invariant subspace Z is called H-irreducible if Z 6= {0} and the only proper subspace of Z that is invariant is {0}. Similar notions of invariance and irreducibility exist for representations of a Lie algebra. CLASSIFYING WAVELET TRANSFORMS IN 3D 9 Denote by H0 the connected component of the identity in H; then H0 is generated by elements of the form exp X, X ∈ h. It follows that Z is H0-invariant if and only if Z is h-invariant, and Z is H0-irreducible if and only if Z is h-irreducible. Given an invariant subspace Z of Rd, let π : V → Rd/Z be the canonical map; H acts linearly on the quotient Rd/Z by h · π(x) = π((hT )−1x), h ∈ H, x ∈ Cd; there is an associated quotient action of h as well. The complexification of the quotient space Rd/Z is Cd/ZC. Given any linear action of h in a real vector space V, a complex linear form λ on h is a characteristic function for this action if the subspace Eλ = {u ∈ VC : X · u = λ(X)u, ∀X ∈ h} is non-zero. The non-zero elements of Eλ are called characteristic vectors for λ. Suppose that λ is a characteristic function for a linear action, and choose a characteristic vector e for λ. If λ is real-valued, then we can choose e ∈ V and Z = Re is h-invariant. If λ is not real-valued, then write e = e1 + ie2 with e1, e2 ∈ V , and put Z = span{e1, e2}. Then dimR Z = 2, Z is h-invariant, and the linear action of h in Z has the characteristic functions λ and λ. Moreover, the subspace Z defined in each case above is both invariant and irreducible. A complex linear form on h is a weight if there is an invariant subspace Z of VC such that λ is a characteristic function for the quotient action of h on VC/Z. The following is easily verified. Lemma 3.2. Let h act linearly on a real vector space V. (a) For any weight λ, [h, h] ⊂ ker λ. (b) If λ is a weight, then λ is also a weight, and Eλ = E λ. (c) If there are d independent weights, then h is commutative. The linear action of primary interest here is the transpose action, that is, the dual of the standard representation. With Rd identified with its dual via the dot product, we have h · v = (hT )−1v, h ∈ H, v ∈ Rd and the corresponding action of the Lie algebra h is X · v = −X T v, X ∈ h, v ∈ Rd. For v ∈ Rd, let H(v) be the stabilizer of v for the transpose action, and let h(v) = {X ∈ h : X T v = 0}) be the annihilator of v in h. Then H(v) is a closed subgroup of H (though not necessarily connected), and h(v) is a Lie subalgebra of h. There is a natural continuous bijection from the quotient space H/H(v) onto the H T -orbit of v given by hH(v) 7→ (hT )−1v, and this map is a homeomorphism if and only if the orbit is a locally compact subset of Rd. We leave the proof of the following to the reader. Lemma 3.3. A linear form λ on h is a weight for the transpose action of h if and only if −λ is a weight for the standard action. 10 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Define φ = φv : H → V by φ(h) = (hT )−1v. The differential of φ at the identity e ∈ H is a map dφe : h → V given by f (exp tX T )−1v) = h▽f (v), dφe(X)i d dt(cid:12)(cid:12)(cid:12)t=0 where f is any smooth function on V. The following is an immediate consequence of the chain rule. Lemma 3.4. One has dφe(X) = −X T v. Hence h(v) = ker dφe and coincides with the Lie algebra of H(v). Let O be an H T -orbit in Rd. Given v ∈ O and w = hv, then h(w) = hh(v)h−1, and hence the map v 7→ dim(h/h(v)) is constant on O. In fact it can shown that O is a submanifold of Rd whose tangent space at each v ∈ O is the image of dφv, so that dim O = dim h/h(v). In particular, for any v ∈ Rd, dim h/h(v) ≤ d. The following lemma translates the open orbit property for H to a condition on h. Lemma 3.5. Let O = H T v ⊂ Rd. Then dim h/h(v) = d if and only if O is open. Proof. Suppose that dim h/h(v) = d. Then for each v ∈ O, (dφv)e is an isomorphism of h/h(v) onto Rd. Now the inverse mapping theorem says that there is an open neigh- borhood U of e in H and an open neighborhood E of v in Rd such that φU is a home- omorphism of U onto E. In particular the orbit O contains an open neighborhood of v. Conversely, suppose that O is open and fix v ∈ O. We first show that the quotient space H/H(v) is homeomorphic with O. Let K be a compact neighborhood of H(v) in H/H(v). The canonical image of K in O is a compact continuous image of K. Now H/H(v) is a countable union of translates of K, and hence O is a countable union of translates of the image of K. It follows that the image of K has non-empty interior in O, hence in Rd. We conclude that there is an open neighborhood U of H(v) in H/H(v) and an open neighborhood V of v ∈ O that are homeomorphic. But now any open subset of H/H(v) is a union of translates of some open subset U0 of U, and hence its image in O is the union of the same translates of the image V0 of U0. Since V0 is open this shows that the canonical map H/H(v) is open, hence a homeomorphism. By the Lie group property (2) above, we have a subspace e of h, having dimension dim(h/h(v)), that contains an open subset homeomorphic with an open subset of Rd, so dim(h/h(v)) = d. (cid:3) Put Ω = {v ∈ V : dim h/h(v) = d}. Fixing a basis {Xj : 1 ≤ j ≤ n} for h and {ei : 1 ≤ i ≤ d} for Rd, then Ω = {v ∈ V : rankM(v) = d} where M(v) is the d × n matrix Thus Ω is Zariski-open in Rd. Of course there are many examples of Lie subalgebras h of gl(d, R) for which Ω is empty. If Ω is nonempty, then we say that h is an open orbit subalgebra of gl(d, R). If H is a closed subgroup of GL(d, R) with Lie algebra h, then M(v) =(cid:2)hei, X T j vi(cid:3) . CLASSIFYING WAVELET TRANSFORMS IN 3D 11 the transpose representation of H produces open orbits if and only if h is an open orbit subalgebra. In this case we say that H is an open orbit group. Lemmas 3.3, 3.4, and 3.5 hold for any representation of a Lie algebra in a real vector space. Given a real subspace V that is invariant for the transpose action, we will also consider the restriction of the transpose action to V, as well as the quotient action on Rd/V. 4. Classifying the solvable groups The classification of the solvable connected subgroups H < GL(3, R) acting with open orbits and associated compact stabilizers constitutes a large portion of this paper, and they contribute all cases except one in the list from Theorem 2.1. Let us briefly outline the arguments and results of this paper, and how they combine to provide the major part of the proof of 2.1(a). We proceed by analyzing weights on the Lie algebra that are associated to the dual action. While one such weight, associated to the dual action of a particular matrix group H, is not a conjugation invariant feature of said group, the existence of weights with certain additional properties turns out to be such a feature. Similarly, the subspace of nilpotent matrices contained in the Lie algebra of H is not a conjugation invariant, but its dimension is. These features of H, or rather its Lie algebra, allows to separate the class of solvable groups into several subclasses, each of which can then be classified further, to the point of computing a list of representatives modulo conjugacy. More precisely, we obtain the following subclasses: • H is abelian; see Proposition 4.3. This contributes the cases (1.a) -- (1.e). Admis- sible groups containing these groups are given in 4.4. • H is nonabelian, and has a nonreal weight; see Proposition 4.7. This contributes the cases (2.a) -- (2c), with the admissible groups containing the connected ones provided by Corollary 4.8. • H is nonabelian, all its weights are real-valued, and the dimension of the nilpo- tent part is two; see Proposition 4.9. This contributes the cases (2.d) -- (2.i); note that case (3) from the Proposition corresponds to the cases (2.f) and (2.g) from Theorem 2.1. The admissible groups containing the connected ones are provided by Corollary 4.10. • H is nonabelian, all its weights are real-valued, and the dimension of the nilpotent part is one. This case is treated in Proposition 4.13, which contributes the cases (2.j) -- (2.n). Note that case (1) from the Proposition corresponds to the cases (2.j) and (2.k) in Theorem 2.1. This time, the admissible groups containing the connected ones are given in Proposition 4.13. We now start with the classification. For subspaces h and k of gl(d, R), define [h, k] = span{[X, Y ] : X ∈ h, Y ∈ k}. For any Lie subalgebra h of gl(d, R), put h(0) = h(0 = h and define h(k), h(k), k = 1, 2, 3, . . . recursively by h(k) = [h(k−1), h(k−1)] and h(k) = [h, h(k−1). Recall that h is said to be solvable (resp. nilpotent) if there is some positive integer n such that h(n) = {0} (resp. h(n) = {0}. 12 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Theorem 4.1. (Lie's Theorem) Any linear action of of a solvable Lie algebra on a complex vector space has a characteristic function. The following consequence is almost immediate. Corollary 4.1. Let h be a solvable Lie subalgebra of gl(d, R). Then h is conjugate with a Lie subalgebra of gl(d, C) that consists of upper triangular matrices. Moreover, the set of weights for the standard representation of h is precisely the set of forms occurring on the diagonal of an upper triangular conjugate of h. Now suppose that h is a solvable Lie subalgebra of gl(3, R), and let W (h) denote the set of weights for the standard representation of h . It is instructive to first treat the case where h is commutative. The following standard result shows that in this case every weight is in fact a characteristic function. Lemma 4.2. Let h be a commutative Lie subalgebra of gl(3, R). (a) Given λ ∈ W (h), the subspace Zλ = {v ∈ C3 : ∀A ∈ h, ∃p ∈ N, (AT − λ(A)I)pv = 0} of C3 is hT -invariant, and there is a basis for Zλ for which the restriction of AT to Zλ has the form λI + N where N is strictly lower-triangular. (b) C3 = ⊕Zλ where the sum is taken over the set W (h) of (distinct) weights. The preceding leads quickly to a classification of connected, abelian, open orbit, admissible subgroups of GL(3, R). Let H be such a group, with Lie algebra h. First suppose that there is a weight λ that is not real-valued. Then W (h) = {γ, λ, λ} where γ is a real- valued form. Recall that Zλ = Z λ, and dim Zλ = dim Zγ = 1, so that h is simultaneously diagonalizable in C3. Write α = Re λ, β = Im λ. Observe that the restriction of the transpose action to Zλ + Z λ has open orbits. Hence α and β are R-linearly independent and α, β, γ is a basis for the real linear dual of h. Choose a basis for h dual to this basis, and choose a basis e1, e2, and e3 for R3 such that Zγ = Ce1 and Zλ = C(e2 + ie3). Thus h is conjugate to the span of 0 0 0 0 0 −1 0 1 0  , 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0  0 0 1 , H ′ =  and H is conjugate to 0 a 0 b cos t −b sin t 0 b sin t 0 b cos t  : a > 0, b > 0, t ∈ R . CLASSIFYING WAVELET TRANSFORMS IN 3D 13 Note that a covering group of H ′ for which there is one co-null open orbit is then H ′F where Now suppose that W (h) consists only of real-valued forms and let n = W (h). If n = 3 then H is conjugate with and a covering group of H ′ with one open orbit is H ′F where If n = 2, then h conjugate to the span h′ of and the corresponding connected subgroup of GL(3, R) is . 0 0 ǫ3 0 0 a3 ǫ1 0 0 0 ǫ2 0 a1 0 0 0 a2 0  ±1 0 0 0 0 0 0 0 0 F = H ′ =  : ai > 0  F =  : ǫi = ±1  0 0 0 ,  0 0 0 , 0 0 1 , H ′ = 0 0 c  : a > 0, c > 0, b ∈ R,  F =  : ǫi = ±1  a b 0 0 a 0 0 0 0 0 0 0 1 0 0 0 1 0 0 1 0 0 0 0 ǫ1 0 0 0 ǫ1 0 0 0 ǫ2 . . . A covering group for which there is one co-null open orbit is then H ′F where Finally, suppose that there is exactly one real-valued weight λ. Choose a basis e1, e2, e3 for R3 such that Re3 and V2 = Re2 + Re3 are hT -invariant. Here n is codimension one in h, and H T acts in R3/Re3 with open orbits. Hence there must be X ∈ n such that X T e1 = e2 + ce3. Now Xe2 may or may not be zero, but for Y ∈ n such that Y T e1 ∈ Re 3, XY = Y X implies that Y T e2 = 0. It follows that dim n = 2, and if Xe2 = 0 also, then it is immediate that h is conjugate to the Lie algebra h′ spanned by the elements I, X, Y where and H is conjugate to X T = H ′ =  0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 , Y T = 1 0 0 , 0 0 a  : a > 0, b, c ∈ R a b c 0 a 0 . 14 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA If X T e2 = δe3 with δ 6= 0, then replacing e3 by e′ spanned by I, X, and Y where 3 = δe3, we find that h is conjugate to h′ and H is conjugate to the corresponding connected subgroup of GL(3, R) given by X T = H ′ =  0 0 0 1 0 0 0 0 0 0 0 0 0 1 0 , Y T = 1 0 0 , 0 0 a  : a > 0, b, c ∈ R a b c 0 a b . In both cases, a covering group of H ′ for which there is a single open orbit is H ′F where F = ±I. Thus we have proved the following; see also [18, Remark 20]. Proposition 4.3. Let H be a closed, connected abelian admissible subgroup of GL(3, R) for which H T has open orbits. Then H is conjugate to exactly one of the following sub- groups. (1) The diagonal group: (2) The block diagonal group H = (R+ · SO(2)) × R∗ +, that is H =  a1 0 0 0 a2 0 0 0 a3  : ai > 0 . . 0 a 0 b cos t −b sin t 0 b sin t H =  H =  H = H = . 0 a b 0 0 a 0 b cos t  : a > 0, b > 0, t ∈ R 0 0 c  : a > 0, c > 0, b ∈ R 0 0 a  : a > 0, b, c ∈ R  0 0 a  : a > 0, b, c ∈ R  a b c 0 a 0 a b c 0 a b . . (3) The group (4) The group (5) The group Of special interest are the irreducibly admissible groups. For abelian case, we have the following. Proposition 4.4. Let H be an abelian irreducibly admissible subgroup of GL(3.R). Then H is conjugate to exactly one of the following subgroups. CLASSIFYING WAVELET TRANSFORMS IN 3D 15 (1) The diagonal group: H =  a1 0 0 0 a2 0 0 0 a3  : ai 6= 0 . (2) The block diagonal group H = (R+ · SO(2)) × R∗ +, that is (3) The group (4) The group (5) The group . 0 0 a b 0 0 a 0 a 0 b cos t −b sin t 0 b sin t H = b cos t  : a 6= 0, b 6= 0, t ∈ R  H = 0 0 c  : a 6= 0, c 6= 0, b ∈ R  H = 0 0 a  : a 6= 0, b, c ∈ R  H = 0 0 a  : a 6= 0, b, c ∈ R, c 6= 0  a b c 0 a 0 c a b 0 a b . . . We now turn to the case where H is non-abelian. Let H be a closed, admissible, solvable, open orbit subgroup of GL(3, R) that is not abelian and let h be its Lie algebra. Then h is not commutative, and we have observed that dim h = 3 or 4. Recall that if λ ∈ W (h) then −λ is a weight for the transpose representation. We will be especially interested in those λ ∈ W (h) for which −λ is a characteristic function for the transpose representation; denote the set of these weights by C(h). Thus for λ ∈ C(h), we have f ∈ C3 such that AT f = λ(A)f, A ∈ h and we let Eλ denote the subspace of all f satisfying the preceding. The characteristic space for h is defined by E(h) = ⊕λ∈C(h)Eλ; let d(h) denote the dimension (over C) of E(h). If λ ∈ C(h), then λ ∈ C(h) also and Eλ = E λ. It follows that E(h) is the complexification of E(h) ∩ R3. Denote by c(h) (resp. w(h)) the dimension of the span of the characteristic functions (resp. weights) for the transpose action of h. If h′ is conjugate to h, then c(h′) = c(h), w(h′) = w(h), and d(h′) = d(h). Since h is not commutative and open orbit, then 1 ≤ d(h) ≤ 2, and 1 ≤ c(h) ≤ w(h) ≤ 2. 16 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Let n ⊂ h be the subspace consisting of those elements for which all of the weights are zero. Since [h, h] ⊆ n, then n is an ideal in h, and since we are assuming that h is non-commutative, then n 6= {0}. Note that dim n = 3 − w(h). Let λ ∈ C(h) and suppose that λ is not real-valued. Choose f ∈ Eλ. The subspace spanned by f, f in C3 is the complexification of a two-dimensional subspace Z of R3 that is hT -invariant. Since R3/Z is one-dimensional, then it is immediate that we have an ordered basis (e, f , f ) for C3 with respect to which AT is given by the matrix (1) 0 γ(A) τ (A) λ(A) τ (A) 0  0 0 λ(A) for all A ∈ h. Here γ is a real linear form, and τ is a complex linear form on h. Note that in this case d(h) ≥ 2. Proposition 4.5. Let h be a non-commutative solvable, open orbit Lie algebra. (1) If C(h) contains a real-valued form, then every form in W (h) is real-valued. (2) Suppose that C(h) contains a form λ that is not real-valued, and write α = Re λ, β = Im λ. Then dim n = 2, and {α, β, γ} is R-linearly dependent in h∗. Proof. To prove (1), suppose that γ is a real-valued characteristic function and that there is a weight λ that is not real-valued. As above, we have e ∈ R3, and f, f ∈ C3, such that the matrix of AT , A ∈ h with respect to the ordered basis (f, f , e) is given by λ(A) 0 τ (A)  0 0 0 λ(A) τ (A) γ(A) where τ is a complex linear form on h. Now since hT is of open orbit type, then hT acts with open orbits in the quotient R3/Re; hence λ has real rank 2, that is, α and β are independent. Choose B ∈ h such that λ(B) = i, and observe that the transpose action of any element of n is determined by its action on f . Given a non-zero element X ∈ n, [B, X] ∈ n also, but [BT , X T ]f = (γ(B) − i)τ (X)e. Since γ(B) − i /∈ R, it follows that dim n = 2. This means that τ and τ are linearly independent over R. Let X1, X2 ∈ n such that τ (X1) = 1, τ (X2) = i. Write f = e1 + ie2 with e1, e2 ∈ R3. Now (e1, e2, e) is an ordered basis for R3 and for v ∈ R3 write v = v1e1 + v2e2 + v3e. If v2 1 + v2 2 6= 0, then it is straightforward to check that (v1X2 − v2X1)v = 0, and hence that exp R(v1X2 − v2X1) is contained in the stabilizer of v. Thus almost all elements of R3 have a non-compact stabilizer, and H is not admissible. Thus (1) is proved. CLASSIFYING WAVELET TRANSFORMS IN 3D 17 To prove (2), choose an ordered basis (e, f , f ) of C3 with e ∈ R3, such that AT is given by 0 γ(A) τ (A) λ(A) τ (A)  0 0 λ(A) 0 for all A ∈ h. Choose B ∈ h such that λ(B) /∈ R. Since h is non-commutative, we have X ∈ n, X 6= 0. Then X T e = τ (X)f + τ (X)f , and since X T f = 0, then X T BT e = γ(B)X T e. Hence if Y T = [BT , X T ], then Y ∈ n and Y T e = (BT X T − X T BT )e = λ(B)τ (X)f + λ(B)τ (X)f − γ(B)(τ (X)f + τ (X)f = (λ(B) − γ(B))τ (X)f + (λ(B) − γ(B))τ (X)f so τ (Y ) = (λ(B) − γ(B))τ (X). Since λ(B) /∈ R, then τ (Y ) and τ (X) are linearly independent over R, and hence Y and X are independent in n. Thus dim n ≥ 2, and since n is determined by the values of τ and τ , then dim n = 2. Finally, since dim h ≤ 4 while dim n ≥ 2, then dim(spanR{α, β, γ}) ≤ 2. (cid:3) As a consequence of the preceding, we have the following corollary. Corollary 4.6. Let h be a non-commutative solvable, open orbit Lie subalgebra of gl(3, R). (1) R3 contains a transpose-invariant subspace of codimension one. (2) Let V be a transpose-invariant subspace of codimension one, let λ be the weight for the quotient action in R3/V, and let h1 = ker λ. Then the transpose action of h1 in V is an open orbit action. As in the case where H is abelian, we describe the conjugacy classes of Lie algebras for solvable admissible H. First suppose that C(h) contains a form λ that is not real-valued so that W (h) = {γ, λ, λ} where γ is real-valued. By Lemma 4.5, λ and λ are characteristic functions, and we have an ordered basis (e, f , f ) for C3 and a complex linear form τ on h, such that the matrix of AT with respect to the basis (e, f , f ) is given by the form (1) for all A ∈ h. Clearly we can take e ∈ R3, and writing e = e1, f = e2 + ie3 with e2, e3 ∈ R3, then (e1, e2, e3) is a basis for R3. Recall that also by Lemma 4.5, dim(spanR{α, β, γ}) = 1 or 2. As before put α = Re λ, β = Im λ. Since λ is not real-valued, then β 6= 0. Hence there are three possibilities: (1) both α and γ are multiples of β, (2) α is a multiple of β but γ is not, and (3) γ is multiple of β but α is not. Observe that these cases are if h′ is a solvable Lie algebra in gl(3, R) that is conjugate to h, "conjugate invariant": then both h and h′ fall within the same case. (1) Suppose that both α and γ are multiples of β. Put µ = Re τ, ν = Im τ . Since the dimension of the span of α, β, γ, µ, and ν is at most 3. We have 3 ≤ dim h = dim span{α, β, γ, µ, ν} ≤ 3 18 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA so dim h = 3, and µ and ν are linearly independent. Choose X1, X2 in n such that with respect to the basis (e1, e2, e3), X T 0 0 0 1 0 0 0 0 0 , X T 1 = 2 = 1 0 0 . 0 0 0 0 0 0 Choose A ∈ h such that µ(A) = ν(A) = 0 while λ(A) 6= 0; using the form (1) for AT , we compute that [A, X1 + iX2] = (γ(A) − α(A) + iβ(A))(X1 + iX2). Since [X1, X2] = 0 we get ∆H (exp tA) = e2t(α(A)−γ(A)) while det(exp tA) = et(2α(A)+γ(A)). Since H is admissible, then ∆H 6= det on H so γ 6= 0. It follows that {γ, µ, ν} is a basis for h∗ and we fix the dual (ordered) basis {A, X1, X2} for h. With λ(A) = a + ib, we now have [A, X1 + iX2] = (1 − a + ib)(X1 + iX2) as well as [X1, X2] = 0, characterizing the Lie algebra. To sum up, we have that h is conjugate to the Lie algebra spanned by {A, X1, X2} where AT = 1 0 0 0 a −b 0 b a , X T 1 = 0 0 0 1 0 0 1 0 0 . 2 = 0 0 0 , X T 0 0 0 0 0 0 (2) Suppose that α is a multiple of β, while γ and β are independent. Choose B, C ∈ h such that γ(B) = 0, λ(B) = 1, γ(C) = 1, λ(C) = 0. Note that β(B) 6= 0, and that n = ker γ ∩ ker λ, being codimension-two in h, is non-trivial. Now by Lemma 4.5, dim n = 2, and using now the real basis (e1, e2, e3) for R3, a basis X1, X2 for n is given by X T 0 0 0 1 0 0 0 0 0 , X T 1 = 2 = 1 0 0 . 0 0 0 0 0 0 Modifying B and C by linear combinations of X1 and X2 if necessary, we can assume that Be1 = Ce1 = 0. We compute that [B, X1 + iX2] = −λ(X1 + iX2), [C, X1 + iX2] = X1 + iX2 characterizing the Lie algebra. It follows that ∆H (exp tB) = e2tα, ∆H(exp tC) = e2t while det(exp tB) = e2tα, det exp tC = et. Thus H satisfies the modularity condition. Whether h is the Lie algebra of an admissible group now depends only upon the stabilizers, CLASSIFYING WAVELET TRANSFORMS IN 3D 19 which are generated by the annihilators in h: annihilator h(v) for generic v ∈ R3 is let v = v1e1 + v2e2 + v3e3. Then the h(v) = R(cid:18)B + Re (cid:18)λ(v2 + iv3) 2v1 (cid:19) X1 + Im (cid:18)λ(v2 + iv3) 2v1 (cid:19) X2(cid:19) By direct computation we verify that exp h(v) is compact if and only if λ is purely imag- inary, that is, α = 0. To sum up, if α is a multiple of β and γ is independent of β, and furthermore h is the Lie algebra of an admissible group, then h is conjugate with the Lie algebra spanned by A, B, X1, X2 where BT = 0 0 0 0 0 −1 0 1 0  , C T = 1 0 0 0 0 0 0 0 0 , X T 0 0 0 , X T 1 = 2 = 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 (3) Here we suppose that α is not a multiple of β, while γ belongs to the real span of α and β. We choose A, B ∈ h such that α(A) = 1, β(A) = 0, α(B) = 0, β(B) = 1. Until the end of this case, we use the complex matrices with respect to the basis (e, f, f ) for C3. Again dim n = 2 and dim h = 4. Let X1, X2 ∈ n be given by τ (X1) = 1 and τ (X2) = i. Using (1) we compute [AT , X T 2 − γ(B)X T 2 . This can be expressed in the complexification of h more simply: with X = X1 + iX2, we get j ] = (1 − γ(A))X T j and [BT , X T 1 − γ(B)X T 1 , [BT , X T 2 ] = −X T 1 ] = X T [AT , X T ] = (1 − γ(A))X T , [BT , X T ] = (−γ(B) − i)X T . To compute h(v), we first observe that h(π(v)) = ker γ = k + R(γ(B)A − γ(A)B). Now 0 0 0 γ(B) − iγ(A) 0 0 0 0 γ(B) + iγ(A) . γ(B)AT − γ(A)BT = a(v) = Re (cid:18) (γ(B) − iγ(A))(v2 + iv3) 2v1 Let v ∈ R3 such that v1(v2 + iv3) 6= 0. Then we have a(v), b(v) ∈ R such that h(v) = R(γ(B)A − γ(A)B − a(v)X1 − b(v)X2) and we compute that (cid:19) , b(v) = Im (cid:18)(γ(B) − iγ(A))(v2 + iv3) 2v1 (cid:19) This generates a torus if and only if γ(B) = 0. Coming back to the real basis (e1, e2, e3), we conclude that if the weights of the transpose action of h satisfy the conditions of 2.2, then there is a basis (e1, e2, e3) for R3 such that h is spanned by elements A, B, X1, X2, which are given with respect to this basis by A = γ(A) 0 0 1 0 0 0 γ(B) 0 0 1 , B = 0 0 0 0 −1 1 0  , X1 = 0 0 0 1 0 0 0 0 0 , X2 = 0 0 0 0 0 0 1 0 0 20 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA and h is admissible if and only if γ(B) = 0. Thus h is a span of elements A, B, X1, X2 and we have a basis for R3 such that these elements are given by AT = a 0 0 0 1 0 0 0 1 , BT = 0 0 0 0 0 −1 0 1 0  , X T 1 = 0 0 0 , X T 2 = 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 where a ∈ R, a 6= 0. We summarize the preceding analysis. Proposition 4.7. Let h be the Lie algebra of a connected, non-abelian admissible, open orbit subgroup H of GL(3, R). Assume that h is solvable and has a weight that is not real valued. Then W (h) = {λ, λ, γ} where γ is real valued, and both λ and λ are characteristic functions. Moreover, exactly one of the following holds. (1) λ has real rank one, and γ is a linear combination of λ and λ. Then there is a unique a + ib ∈ C \ R such that h is conjugate with the Lie algebra spanned by  H =   1 0 0 0 a −b 0 b a ,  es 0 0 t1 eas cos(bs) −eas sin(bs) eas sin(bs) . t2 1 0 0 0 0 0 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 ,  0 0 0 . eas cos(bs) : s, t1, t2 ∈ R 0 0 0 ,  0 0 0 ,  0 0 0 cos(r)  : r, s, t1, t2 ∈ R 0  , X T 0 0 0 , X T 1 = 2 = 0 0 0 . 0 1 0 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 0 0 1 0 0 0 . t1 t2 cos(r) − sin(r) sin(r) 0 0 0 0 0 −1 0 1 The group H is given by Here we have es 0 0 0 0 0 0 0 −1 0 1 0  ,  H =  0 0 1 , BT = a 0 0 0 1 0 AT = (2) λ has real rank one, and γ is not a linear combination of λ and λ. Then h is conjugate to the Lie algebra spanned by (3) λ has real rank two. Then there is a unique non-zero real number a such that h is conjugate with the Lie algebra spanned by CLASSIFYING WAVELET TRANSFORMS IN 3D 21 In this case eas 0 0 H =  t1 es cos(r) −es sin(r) es sin(r) t2 es cos(r) : r, s, t1, t2 ∈ R . In each of the above cases we see that H has two open orbits, the orbits of [±1, 0, 0]T in R3. Hence the following is immediate. Corollary 4.8. Let h be the Lie algebra of a connected, non-abelian admissible, open orbit subgroup H of GL(3, R). Assume that h is solvable and has a weight that is not real valued. Let E denote the identity matrix. Then {−E, E}H is irreduciablly admissible. We now turn our attention to the second main case. Suppose now that h is the Lie algebra of a closed, solvable, admissible, open orbit subgroup of GL(3, R) and that every weight for the standard representation of h is real-valued. By Lemma 4.1, h is contained in the algebra of real upper-triangular 3 × 3 matrices, so the matrix exponential is injective on h. The image H = exp h is the (unique) closed connected subgroup of GL(3, R) corresponding to h, and is an exponential Lie group. Now a one-parameter subgroup of an exponential group is non-compact, and for each v ∈ R3, the stabilizer H(v) is connected. Since H is open orbit admissible, then dim h = 3. Recall that E = E(h) is the span of all characteristic vectors for the transpose action; since h is not commutative, then dim E ≤ 2. In the preceding analysis, we saw that if λ ∈ W (h) is not real-valued, then both λ and λ belong to C(h), and hence dim E = 2. When all weights are real, then dim E may be either 1 or 2, and this will lead a greater number of conjugacy classes. As always, n = ∩λ∈W ker λ denotes the ideal of nilpotent elements in h. Since h is an open orbit Lie algebra, then n 6= h; indeed dim n = 1 or 2. Another reason that the complex weight case is simpler is that by Lemma 4.5, the presence of a weight that is not real implies that dim n = 2. In what follows we will distinguish conjugacy classes by these two characteristics: the dimension of E, and the dimension of n. It will also be helpful to identify the isomorphism classes of three-dimensional solvable Lie algebras that are represented here. In addition to the Heisenberg Lie algebra, several non-nilpotent Lie algebras will occur. They are defined by writing the non-vanishing Lie brackets between basis elements. s0 = span{A, C, Y } with [A, Y ] = Y . s1,c = span{A, X, Y } with [A, X] = cX, [A, Y ] = Y . Here 0 < c ≤ 1. s2 = span{A, X, Y } with [A, X] = X + Y, [A, Y ] = Y . In each case the connected group H = exp h is easily obtained via the exponential map. An irreducibly admissible group is a finite extension of H. 22 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Suppose first that dim n = 2. Since H is an open orbit group, there is a non-zero weight, of which every weight is a multiple. We describe the conjugacy classes of Lie algebras satisfying these requirements by identifying a basis of n and an element A /∈ n. (1) Suppose that dim E = 2. Since h is of open orbit type then the quotient of the transpose action of h on the one-dimensional space R3/E is given by a non-zero weight. Moreover, n acts trivially on E. It follows immediately from these observations that h is conjugate to a Lie algebra spanned by A, X, Y where A is diagonal, and where X and Y are given by X T = 0 0 0 1 0 0 0 0 0 , Y T = 0 0 0 0 0 0 1 0 0 . Since h is non-commutative, then there are at least two distinct weights, that is, A cannot If the weight for R3/E coincides with a characteristic be a multiple of the identity. function, then h belongs to the class s0; indeed, h is conjugate to the Lie algebra spanned by basis A, X, Y where AT = 0 0 λ 0 0 λ 0 0 λ − 1 and where λ 6= 0. Note that here C = X is central and [A, Y ] = Y . If dim E = 2 and there is a weight that is not a characteristic function, then h is conjugate to the Lie algebra spanned by A, X, Y where AT = 0 λ 0 λ − 1 0 0 0 0 λ − b , X T = 0 0 0 1 0 0 0 0 0 , Y T = 0 0 0 0 0 0 1 0 0 where λ 6= 0 and b ≥ 1. Thus [A, X] = X, [A, Y ] = bY . (2) Suppose that dim n = 2, and that dim E = 1, E = Re3. Since E is invariant under the transpose representation, we have a representation α of h in R3/E; let k be the kernel of this quotient representation. Since the quotient map is open, then α is an open orbit action, and dim k = 1. Since h is open orbit, then n + k 6= h, and hence k ⊂ n. Choose X ∈ n \ k and Y ∈ n ∩ k, Y 6= 0. The subspace V = X R3 + E is invariant. With e1 ∈ R3 such that e2 = X T e1 /∈ E, then V = span{e2, e3} and it is easily seen that every element of h is upper triangular with respect to the ordered basis (e1, e2, e3). We can then modify this basis {X, Y } of n so that they are given with respect to the basis (e1, e2, e3) by X T = 0 0 0 1 0 0 0 δ 0 , Y T = 0 0 0 0 0 0 1 0 0 Now compute AT = [AT , X T ] = 0 c2 − 1 δ(A) 0 c2 0 1 0 0 0 δ(A) c3  . (c3 − c2)δ 0 0 0 0 0 CLASSIFYING WAVELET TRANSFORMS IN 3D 23 where δ = 0 or 1. We then have A ∈ h given by and since [AT , X T ] belongs to nT , it follows that [AT , X T ] = (c2 − 1)X T + δ(A)Y T , (2) [AT , Y T ] = (c3 − 1)Y T , and that (3) (c3 − c2)δ = (c2 − 1)δ. The constants c2 and c3 express the relations among the weights of the transpose action of h, and are therefore conjugate-invariant. Let n be the number of distinct weights. In light of (2) and (3), we observe the following. • n = 1 if and only if c2 = c3 = 1, whence δ(A) 6= 0, and h is a Heisenberg Lie algebra. • n = 2 and there are two weights for the quotient action in R3/E, if and only if c2 = c3 6= 1. Here δ = 0 and since V 6= E, then δ(A) 6= 0. • If n = 2 and c2 6= c3, then either c2 = 1 or c3 = 1, and hence δ = 0. But c2 6= c3 also implies that A is diagonalizable, whence dim E = 2, which is not the case here. The preceding observations lead to several isomorphism classes of Lie algebra. (a) Suppose that n = 1, that is, c2 = c3 = 1. As observed h is a Heisenberg Lie algebra and there are two one-parameter families of conjugacy classes. If δ = 0, that is X 2 = 0, then h is conjugate with the Lie algebra spanned by A, X, Y , given by where λ 6= 0. If δ = 1 , then h is conjugate with the Lie algebra spanned by A, X, Y given with respect to this basis by AT = AT = λ 0 0 0 λ 0 0 1 λ , X T = 0 1 λ , X T = λ 0 0 0 λ 0 0 0 0 1 0 0 0 0 0 , Y T = 0 1 0 , Y T = 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 1 0 0 0 0 0 0 0 0 Here again λ 6= 0 but now X 2 6= 0, so these two families of conjugacy classes are distinct. 24 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA (b) Suppose that n = 2, but c2 = c3. The relation (3) shows that δ = 0. Here there is a two-parameter family of conjugacy classes each of which has a basis B, U, V of h such that Namely h is conjugate with the Lie algebra spanned by B, U, V given by [B, U] = U + V, [B, V ] = V. BT = λAT = 0 λ 0 λ + 1 0 ν where λ = 1/(c2 − 1), and ν 6= 0. 0 0 λ + 1 , U T = 0 0 0 1 0 0 0 0 0 , V T = 0 0 0 0 0 0 ν 0 0 The third observation above shows that there is only one more case: (c) n = 3, so that c2 6= c3 and c2 6= 1, c3 6= 1. The basis (e2, e3) for V can be chosen so that A is diagonal, and since dim E = 1, we must have δ = 1. Putting λ = 1/(1 − c2), then the relations (3) show that h is conjugate with the Lie algebra spanned by B, X, Y where BT = λA = 0 λ 0 λ − 1 0 0 and [B, X] = X, [B, Y ] = 2Y . 0 0 λ − 2 , X T = 0 0 0 1 0 0 0 1 0 , Y T = 0 0 0 0 0 0 1 0 0 . We sum up as follows. Proposition 4.9. Let H be a closed, connected, admissible, open orbit subgroup of GL(3, R) whose Lie algebra h is non-commutative solvable with real weights. Suppose further that the dimension of the subalgebra of nilpotent matrices in h is two. Then h is conjugate to exactly one of the following Lie algebras. (1) If dim E = 2 and every weight is also a characteristic function, then h is conjugate to the Lie algebra spanned by basis A, C, Y where and where λ 6= 0. Here C is central and [A, Y ] = Y , so h belongs to the isomorphism class s0. Computing the matrix exponential one finds 0 0 λ 0 0 λ 0 0 λ − 1 , C = A = H = exp h =  aλ 0 0 t1 aλ 0 0 0 1 0 0 0 0 1 0 0 0 0 0 0 0 , Y T = 0 0 0 . aλ−1 : a > 0, t1, t2 ∈ R t2 0 . CLASSIFYING WAVELET TRANSFORMS IN 3D 25 (2) If dim E = 2 and there is a weight that is not a characteristic function, then h is conjugate to the Lie algebra spanned by A, X, Y where where λ 6= 0 and 0 < c ≤ 1. (Note that (1) is a limiting case of the above.) Thus [A, X] = cX, [A, Y ] = Y , so h belongs to the isomorphism class s1,b. Here one finds that H = exp h is given by (3) If dim E = 1 and there is only one weight, then h is conjugate with the Lie algebra spanned by A, X, Y , given either by (4) If dim E = 1 and there are two weights for the quotient action in R3/E, then h is conjugate with the Lie algebra spanned by or t2 0 . 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 1 0 0 1 0 0 0 0 0 1 0 0 0 aλ 0 0 t1 aλ−c 0 λ 0 0 0 λ 1 λ 0 0 0 λ 1 0 λ 0 λ − c 0 0 A = 0 0 1 0 0 0 0 0 0 λ − 1 , X = 0 0 0 , Y = aλ−1 : a > 0, t1, t2 ∈ R 0 0 0 0 0 0 , Y = 0 0 0 , Y = 0 0 0 aλ  : a > 0, t1, t2 ∈ R  : a > 0, t1, t2 ∈ R 0 0 0 , Y = 0 0 0 . aλ−1  : a > 0, t1, t2 ∈ R H =  0 0 λ , X = A = 0 0 λ , X = A = H = exp h =  H = exp h =  λ − 1 , X = A = H =  t2 t1 aλ aλ ln a 0 λ 0 λ − 1 0 0 aλ−1 δaλ−1 ln a 0 1 0 0 0 0 aλ 0 0 t1 aλ 0 aλ 0 0 t1 0 0 0 δ t1 + aλ ln a t2 aλ . . 0 0 1 0 0 0 . where λ 6= 0. In both cases h is a Heisenberg Lie algebra. In the first case aλ 0 0 while in the second case, Here h belongs to the isomorphism class s2. In this case H = exp h is given by t2 where λ, δ ∈ R∗. 26 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA (5) If dim E = 1 and there are three distinct weights, then h is conjugate with the Lie algebra spanned by A, X, Y where with λ 6= 0. Here h belongs to the isomorphism class s1,1/2. Exponentiating and simplifying gives 0 0 0 0 0 0 1 0 0 0 0 1 0 0 0 1 λ 0 λ − 1 0 A = λ − 2 , X = H = exp h =  0 0 0 , Y = 0 0 0 , aλ−2  : a > 0, t1, t2 ∈ R aλ 0 0 t1 0 aλ−1 t1aλ−2 t2 . The same argument as for Corollary 4.8 is applicable here as well: Corollary 4.10. Let H be a closed, connected, admissible, open orbit subgroup of GL(3, R) whose Lie algebra h is non-commutative solvable with real weights. Suppose further that the dimension of the subalgebra of nilpotent matrices in h is two. Let E denote the identity matrix in GL(3, R). Then {E, −E}H is irreducibly admissible. Remark that the group F H is also irreducibly admissible, where F =  ǫ 0 0 0 1 0 0 0 1 : ǫ ∈ {−1, 1} . Now we analyze the situation where dim n = 1. The following shows that if dim n = 1, then only one isomorphism class of Lie algebras occurs. Lemma 4.11. Suppose that dim n = 1. Then h belongs to the class s0, and [h, h] = n. Proof. Since h is non-commutative and dim n = 1, then [h, h] = n. Write n = RY ; it is enough to show that Y is not central in h: since n = [h, h], then we have A ∈ h such that [A, Y ] = Y , and if B is independent of A and Y , then [A, B] = cY for some c ∈ R and C = B − cY is central. To see that Y is not central in h, suppose the contrary; then h is Heisenberg. Let h be the image of h in gl(2, R) given by the quotient action of hT on R3/Re3. Since gl(2, R) does not contain a Heisenberg algebra, then the image of Y in h is zero and h is commutative. Now there are exactly two conjugacy classes of commutative Lie algebras in gl(2, R) that are of open orbit type: h is conjugate to the Lie algebra spanned either by or by (cid:20)1 0 0 0(cid:21) , (cid:20)1 0 0 1(cid:21) , (cid:20)0 0 0 1(cid:21) (cid:20)0 0 1 0(cid:21) . CLASSIFYING WAVELET TRANSFORMS IN 3D 27 In the first case h is conjugate to a Lie algebra spanned by elements of the form P = 1 0 ν1 0 0 δ1 0 0 c , Q = 0 0 ν2 0 1 0 0 δ2 d , Y = 0 0 ν3 0 0 0 0 δ3 0 , where Y commutes with P and Q and [P, Q] ∈ RY ; a simple computation shows that this is impossible. A similar argument shows that the second case also leads to a contradiction. (cid:3) We continue to assume that n = [h, h] = RY for some Y 6= 0. Observe that E is hT - invariant and Y T E = 0. Recall that dim E ≤ 2 since h is non-commutative. First suppose that dim E = 2, and that there are two linearly independent characteristic functions. Here we have an ordered basis (e1, e2, e3) for R3 and a basis A, B, Y for h such that with respect to the basis (e1, e2, e3), The fact that [A, Y ] ∈ RY and [B, Y ] ∈ RY implies that µ(Y )ν(Y ) = 0. We may then assume that the basis (e1, e2, e3) and A, B, Y for h are chosen so that b 0 a 0 0 µ(A) 1 0 0 0 µ(B) 0 0 AT = ν(A) 0 0 , BT = ν(A) 0 0 , BT = AT = ν(B) 0 1 , Y T = ν(B) 0 1 , Y T = 0 0 1 0 0 0 µ(Y ) 0 0 ν(Y ) 0 0 . 0 0 0 . 0 0 0 1 0 0 0 0 0 0 a 0 b 0 Here [A, Y ] = (a − 1)Y, [B, Y ] = bY ; since Y is not central, then either a 6= 1 or b 6= 0 must hold. Since [A, B]e1 ∈ Re3 but also [A, B] ∈ RY , then [A, B] = 0. Hence the center of h is RC where C is given by If a + b 6= 1, then (again modifying the basis of R3) we may assume that ν(C) = 0, and since [A, C] = [B, C] = 0, then ν(A) = ν(B) = 0 also. If a 6= 1 then replacing A by λA where λ = 1/(a − 1) gives 0 b b 0 0 0 C T = bAT + (1 − a)BT = ν(C) 0 1 − a . 0 0 0 0 1/b . A′ = A′ = (1/b)B = λ + 1 0 0 λ 0 1 0 0 0 0 0 0 0 while if a = 1 and b 6= 0, then we can take 28 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA These cases are conjugate-invariant since the rank of C is either 1 or 3 when a 6= 1, while if a = 1 and b 6= 0 then the rank of C is two. Suppose now that a + b = 1; both a 6= 1 and b 6= 0 hold. Here the center is RC where for some ν(C) ∈ R. If a = 0 and b = 1, then we get that h is conjugate to the Lie algebra spanned by A, C, Y where and ν(A) ∈ R. Observe that in this case, for v ∈ R3 such that v1 6= 0, the matrix M(v) has rank 3 if and only if ν(A) 6= 0. Thus the assumption that h is an open orbit subalgebra means that ν(A) 6= 0. If a 6= 0, then again replacing A by λA where l = 1/(a − 1) we get that h is conjugate to the Lie algebra spanned by A, C, Y where C = I and C T = AT = 1 0 0 0 1 0 ν(C) 0 1 ν(A) 0 0 0 0 1 0 0 0 λ + 1 0 0 λ 0 0 0 AT = 0 0 . Note that in this case both λ and λ + 1 are non-zero. Continuing the analysis when dim n = 1, suppose now that dim E = 2, but that the characteristic functions are linearly dependent. Here again Y E = 0, and of course there are two linearly independent weights. Hence we can choose an ordered basis (e1, e2, e3) for R3 such that e2, e3 spans E and Y e1 = e2. We then have a basis A, B, Y for h given by AT = 0 0 0 0 1 0 ν(A) 0 a , BT = 1 0 0 0 0 0 ν(B) 0 0 , Y T = 0 0 0 1 0 0 0 0 0 with respect to the chosen basis. Replacing e1 by e1 − ν(A)e3 we may assume that ν(A) = 0. One finds that [A, Y ] = −Y and [B, Y ] = Y , and since [A, B] ∈ RY , then [A, B] = 0. So with C = A + B, we have RC is the center of h. Note that the relation [A, B] = 0 implies that ν(A) + aν(B) = 0. If a 6= 1, then C is diagonalizable and since A commutes with C, then A is simultaneously diagonalizable; thus h is conjugate to the Lie algebra spanned by A, C, Y where AT = 0 0 0 0 1 0 0 0 a , C T = 1 0 0 0 1 0 0 0 a , Y T = 0 0 0 1 0 0 0 0 0 . CLASSIFYING WAVELET TRANSFORMS IN 3D 29 If a = 1, then ν(A) + ν(B) = 0 and C is the identity matrix. Thus h is conjugate to the Lie algebra spanned by A, C, Y where AT = 0 0 0 0 1 0 0 0 1 , C T = 1 0 0 0 1 0 0 0 1 , Y T = 0 0 0 1 0 0 0 0 0 . Next we consider the situation where dim n = dim E = 1; write E = Re3 for some e3 ∈ R3, and we consider the quotient action in R3/E. As before we consider the quotient of the transpose representation of h acting in R3/E, and let k denote its kernel. As before we have dim k = 1. Lemma 4.12. Let h be a solvable open orbit Lie subalgebra of gl(3, R) and suppose that dim n = dim E = 1. Then the action of h in R3/E is diagonalizable. Proof. Since dim n = 1 also, then either n = k, or n ∩ k = {0}. If n = k, then h/k is a commutative Lie aubalgebra of gl(2, R), and this case there are only two possibilities for the quotient of the transpose action in R3/E: either it has two independent characteristic functions (i.e., is diagonalizable), or just one weight. Now suppose that the lemma is false. Then either n = k and the action of h in R3/E has only one weight, or n 6= k. We dispense with these cases one at a time. Suppose that n = k and that the action of h in R3/E has only one weight. We have an ordered basis (e1, e2, e3) for R3 with respect to which each A ∈ h is given by where again µ, ν, and δ are real linear forms on h. The open orbit assumption implies that λ 6= 0 and that µ is independent of λ. Hence we have A, B, Y ∈ h of the form 0 0 0 0 ν(Y ) δ(Y ) 0 , 0 0 1 0 AT = ν(A) δ(A) a , BT = [BT , Y T ] = 0 1 0 0 0 λ(A) µ(A) λ(A) ν(A) 0 0 0 0 δ(A) λ2(A) b , Y T = bδ(Y ) 0 , 0 0 0 0 0 0 0 0 0 1 ν(B) δ(B) bν(Y ) − δ(Y ) AT = Since dim n = 1 then b 6= 0. Now we compute that and since [B, Y ] ∈ RY this leads to δ(Y ) = 0. We choose Y so that ν(Y ) = 1, and A and B so that ν(A) = ν(B) = 0. Since b 6= 0, then the basis (e1, e2, e3) can be chosen so that δ(B) = 0. Finally, observe that since [A, B] ∈ RY then δ([A, B]) = 0; but a direct computation shows that δ([A, B]) = −dδA, so δ(A) = 0. but now e2 and e3 are characteristic vectors for h, a contradiction to our assumption that dim E = 1. 30 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Next assume that n 6= k, that is, that the quotient action of h on R3/E is not commutative. Writing E = Re3, Choose C ∈ k so that Ce3 = e3. Since Y 6= 0, we may choose an ordered basis (e1, e2, e3) for R3 so that h is upper triangular, and so that Y e1 = e2. Now [C, Y ] ∈ n ∩ k = {0} so [C, Y ] = 0 and Y C = 0, so 0 = Y Ce1 = CY e1 = Ce2. Similarly [A, C] = 0. Thus the center of h is RC. It follows that both e2 and e3 are characteristic vectors for h, again contradicting the assumption that dim E = 1. (cid:3) We proceed to describe the distinct conjugacy classes of h when dim E = 1 and dim n = 1. By Lemma 4.12 we have an ordered basis (e1, e2, e3) for R3 with respect to which each A ∈ h is given by AT = λ1(A) 0 ν(A) 0 0 0 λ2(A) δ(A) λ3(A) where λi, µ, ν, δ are real linear forms on h, and where the characteristic function λ3 is dependent upon λ1 and λ2. Now choose A, B ∈ h such that λ1(A) = λ2(B) = 1, λ2(A) = λ1(B) = 0. Given Y so that n = RY , then a computation of [AT , Y T ] ∈ RY T shows that either ν(Y ) = 0 or δ(Y ) = 0. Since the basis elements e1, e2 are interchangeable, then we may assume that ν(Y ) = 1, δ(Y ) = 0, and then we may assume that ν(A) = ν(B) = 0. Thus we have a basis (e1, e2, e3) for R3 and A, B, Y ∈ h given with respect to this basis by AT = 1 0 0 0 0 0 0 δ(A) a , BT = 0 1 0 0 0 δ(B) 0 0 b , Y T = 0 0 0 0 0 0 1 0 0 . We compute that [A, Y ] = (1 − a)Y, [B, Y ] = −bY ; since Y is not central then either a − 1 or b is non-zero, and since [A, B] ∈ RY while [A, B]e1 = 0, then [A, B] = 0. Hence the center of h is spanned by the element C = bA + (1 − a)B and with respect to our basis for R3, C T is given by (4) C T = 0 b 0 0 1 − a 0 0 δ(C) b . If a + b 6= 1, then C is diagonalizable on the span of e2 and e3; but then, C being central in h, both e2 and e3 would be characteristic vectors for all of h. Since we assume now that dim E = 1, this is not the case. Hence a + b = 1, and we may assume that C T = 1 0 0 0 0 1 0 δ(C) 1 . Now with AT as above, [A, C] = 0 implies aδ(C) = 0. If a 6= 0, then δ(C) = 0 and C = I, but also (modifying the element e2) we get δ(A) = 0, again contradicting our assumption CLASSIFYING WAVELET TRANSFORMS IN 3D 31 that dim E = 1. Thus a = 0, and with respect to the basis (e1, e2, e3), AT = 1 0 0 0 0 0 0 δ(A) 0 , C T = 1 0 0 1 0 0 0 δ(C) 1 , Since dim E = 1, then either δ(A) 6= 0 or δ(C) 6= 0. Y T = 0 0 0 0 0 0 1 0 0 . We sum up the preceding. Proposition 4.13. Let H be a closed, connected, admissible, open orbit subgroup of GL(3, R) whose Lie algebra h is non-commutative solvable with real weights. Suppose further that the dimension of the subalgebra of nilpotent matrices in h is one. Then h is conjugate to a Lie algebra spanned by matrices A, C, and Y where [A, Y ] = Y , RC is the center of h, and where the matrices A, C, and Y are exactly one of the following. (1) If there are two linearly independent characteristic functions and three distinct weights, then there are two families of conjugacy classes. If the rank of a non-zero central element of h is odd, then while if the rank of a non-zero central element of h is 2, then where λ 6= 0, c 6= 1. In the first case the group H = exp h is 0 0 0 while in the second, 0 0 aλ+1bc 0 0 0 1 0 0 1 0 0 0 1 0 0 0 0 1 0 0 c 0 0 0 c 0 1 0 0 0 0 0 λ + 1 0 0 λ 0 AT = 0 0 , C T = AT = 0 0 λ , C T = H =  H =  F =  0 0 1 , Y T = 0 0 0 0 0 0 , Y T = 0 0 0 b : a > 0, b > 0, t ∈ R 0 aλ : a > 0, b > 0, t ∈ R  : ǫ1, ǫ2 ∈ {−1, 1} ǫ1 0 0 1 0 0 0 0 ǫ2 t 0 aλbc 0 ab 0 0 t b 0 0 An irreducibly admissible group is here given by F H where . 32 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA (2) If there are two linearly independent characteristic functions and two distinct weights, then again there are two families of conjugacy classes. Here ν1 6= 0. In this case connected group H is given by AT = H =  0 0 0 0 −1 0 ν1 0 0 , C T = t −ν1 ln a + ν2b ln b b 0 ab 0 0 0 b 1 0 0 0 1 0 0 0 0 1 0 0 0 0 0 ν2 0 1 , Y T =  : a > 0, b > 0, t ∈ R . An irreducibly admissible group is here given by F H where (3) If the dimension of the characteristic space E is two while the characteristic functions are linearly dependent, then (4) If the dimension of the space E generated by the characteristic vectors is one, then In this case H is given by with λ 6= 0. An irreducibly admissible group is here given by F H where . , t b 0 0 ab 0 0 ǫ 0 0 0 1 0 1 0 0 0 0 0 1 0 0 0 1 0 0 0 0 1 0 0 0 0 0 . 0 0 1 : ǫ ∈ {−1, 1} 0 0 λ , Y T = 0 bλ : a > 0, b > 0, t ∈ R  : ǫ1, ǫ2 ∈ {−1, 1} 0 δ2 1 , Y T = 1 0 0 ,  : a > 0, b > 0, t ∈ R F =  0 0 0 , C T = AT = H =  F =  AT = 0 δ1 0 , C T = H =  F =  0 1 : ǫ1, ǫ2 ∈ {−1, 1} ǫ1 0 0 1 0 0 0 0 ǫ2 δ1 ln a + δ2b ln b 0 0 0 0 0 0 0 0 ǫ2 0 2 6= 0. In this case ab 0 b 0 0 0 1 0 0 0 1 0 0 1 0 0 ǫ1 0 0 . . t b 0 0 . where δ2 1 + δ2 An irreducibly admissible group is here given by F H where CLASSIFYING WAVELET TRANSFORMS IN 3D 33 5. The non-solvable case It remains to show that the only non-solvable connected matrix group H < GL(3, R) acting with open orbits and associated compact stabilizers is conjugate to R+ × SO(3). Referring to the definition in [30], a real linear connected reductive group is a closed connected group of real matrices which is invariant under transpose. A real linear connected semisimple group is a connected linear real reductive group with finite center. It is well-known that if G is connected and semisimple, then its algebra must be semisimple (is simple or is a direct sum of simple algebras) and if G is a connected reductive Lie group then its Lie algebra can be written as a direct sum of the central ideal with the commutator ideal which is necessarily semisimple. Let G be a semisimple connected Lie group of real matrices with Lie algebra g. Let Θ be the inverse transpose which defines an automorphism on G. Θ is the so called Cartan involution. Next, let T = {g ∈ G : Θ (g) = g} . Then T is a maximal compact subgroup of G. We define an involution ι : g → g such that ι (X) = −X T . This involution is the derivative of Θ at the identity of the group. It is also a Lie algebra automorphism. As a linear operator defined over g, the map ι has two distinct eigenvalues: 1, −1. Let k be the eigenspace corresponding to the eigenvalue 1 and let p be the eigenspace corresponding to −1. Next, the Cartan decomposition is defined as g = k + p. It is worth noting that the algebra k contains skew-symmetric matrices, while p consists of symmetric matrices. The following facts will be important. (1) The map T × p → G given by (z, X) 7→ z exp X is a diffeomorphism onto G. For a proof, we refer to [30, Page 362]. (2) Let h be a four-dimensional Lie algebra which is not solvable. Let h = r + s be the Levi decomposition of the algebra where s is the semisimple part of the decom- position and r is the solvable part. In this case, the dimension of the semisimple part must be equal to three. Next let k + p be the Cartan decomposition of s. If the dimension of the algebra k is equal to one then the algebra s is isomorphic to sl(2, R). Otherwise, s is isomorphic to so(3, R). In fact it is well-known that any non-solvable four dimensional Lie algebra is Lie isomorphic to either sl(2, R) ⊕ R or it is Lie isomorphic to so(3, R) ⊕ R; see the class labelled U3 in [32]. (3) Up to isomorphism there is one (m + 1)-dimensional irreducible representation of sl(2, R) for each natural number m, and every finite-dimensional representation of sl(2, R) is completely reducible (thanks to Weyl's theorem.) Let E3 be the identity matrix of order three. We shall now prove that if H is a four- dimensional non-solvable Lie group of matrices of order three which acts with open orbits, then up to conjugation H is one of the following groups • SO0(2, 1) · exp(RE3) • SO(3, R) · exp(RE3) The first group will then be excluded on the grounds that one of its open orbits has noncompact stabilizers. 34 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Case (A) Let us choose the following basis for sl(2, R) satisfying [H, X] = 2X, [H, Y ] = 2Y, [X, Y ] = H. Let π be an irreducible representation of this algebra in gl (V ) where V is a three- dimensional vector space. Let α be a real eigenvalue for the element π (H) (there must be one since the dimension of V is odd), and let v be a corresponding eigenvector. By the irreducibility of the representation, the set {v, π (X) v, π (Y ) v} necessarily spans V. More- over, using the fact that [H, X] = 2X, [H, Y ] = 2Y, it is easy to see that the endomorphism π(H) is diagonalizable with eigenvalues α, α + 2, α − 2. Next, let w be an eigenvector with corresponding eigenvalue α − 2 for π(H). Using the structure of the Lie algebra, with straightforward calculations, we check that the span of the set {w, π(X)w, π(X)2w} is invariant under the action of π(h). Appealing to the irreducibility of π the elements w, π(X)w, 1 2π(X)2w form a basis for the vector space V. With respect to this basis, we obtain α − 2 0 α for some real numbers µ1, µ2. Next since [X, Y ] = H, it immediately follows that 0 0 π (H) = π (H) = 0 0 0 0 0 1 0 0 0 α + 2  , π (X) = 0 2  , π (X) = 0 2 0  , π (Y ) = 0 2 0  , π (Y ) = 0  0 0 0  . 0 0 µ2 0 0 µ1 0 0 0 0 0 1 0 0 0 2 0 0 0 1 −2 0 0 0 0 0 0 Furthermore, up to isomorphism the irreducible representation π is unique. Let τ be a three-dimensional faithful representation of a Lie algebra with basis H, X, Y, Z satisfying the following non-trivial Lie brackets [H, X] = −2X, [H, Y ] = 2Y, [Y, X] = H such that the real span of H, X, Y is sl (2, R) . This is a reductive Lie algebra with a one- dimensional center spanned by the element Z. Next, it is clear that if the representation τ is irreducible then up to isomorphism 1 0 0 0 1 0 0 0 1  . Let h be the linear span of the matrices above. We shall show that this algebra is isomorphic to so(2, 1) ⊕ R where τ (H) = Put 0 0 2 0 0 0 0 0 0 1 0 0 0 0 −2  , τ (X) = so(2, 1) = M ∈ sl(3, R) : M J = 0 2 0  , τ (Y ) = 0 0 −1  +  = 0 1 0 2 0 1 − 1 2 0 1 1 2 2 1 0 0 1 0 0 0 −1 1 1 0 0 0 1 0 2 0 0 0 1 0 0 0  , τ (Z) = 0 0 −1  M T = 0 1  1 0 0 1 0 0 −1 . . CLASSIFYING WAVELET TRANSFORMS IN 3D 35 Indeed, with straightforward computations, we verify that We now define h1 to be the algebra spanned by τ (H), τ (X), τ (Y ). It follows from the computations above that h1 is isomorphic to so(2, 1) which is spanned by J(cid:18)τ(cid:18)− and Y 2(cid:19) + τ(cid:18) X 0 0 0 −1 0  0 0 0 0 1 0 0 0 0 0 0 −1 0 −1 0 1 0 0 2(cid:19)(cid:19) J −1 = J(cid:18)τ(cid:18)Y 2(cid:19) + τ(cid:18)X 0  , A = 0 0 −1  + 2(cid:19) J −1 = 1 0 0  . 0  , Jτ(cid:18)H 2(cid:19)(cid:19) J −1 = 1 0 0  , B = 0  . 0 0 −1  (tK + rA + sB)T 0 0 0 −1 0 0 1 0 0 0 0 0 0 −1 1 0 0 1 1 0 0 1 0 0 0 0 0 0 0 −1 0 1 0 0 K = (tK + rA + sB) Indeed it is easy to check that for any real numbers t, r, s is equal to the zero matrix. Thus, h is isomorphic to so(2, 1) ⊕ R. Next, we aim to compute the dual action of SO0(2, 1) exp RE3 (its Lie algebra is spanned by K, A, B, E3) where E3 is the identity matrix of order three. The corresponding Cartan decomposition is given by so (2, 1) = k + p where k = RK and p = RA + RB. Let T be a maximal torus subgroup with Lie algebra k. Since the map T × p → SO0 (2, 1) where (z, X) 7→ z exp X is a diffeomorphism then every element of SO0 (2, 1) can be written as where (θ, a, b) ∈ [0, 2π) × R × R. Writing (a, b) in polar coordinates such that a = r cos s, and b = r sin s, we obtain that 0 0 cos θ cos θ − sin θ 0 0 sin θ  exp (aA + bB) = exp 1  exp (aA + bB) 0 0 0 0 r cos s r sin s r cos s r sin s 0  is equal to   cos2 s cosh r + (sin s)2 − sin 2s sinh2( r 2) cos s sinh r − sin 2s sinh2( r 2 ) cos s sinh r cos2 s + cosh r(sin s)2 − sin s sinh r − sin s sinh r cosh r  . Thus, every element of SO0(2, 1) exp RE3 can be uniquely written as cos θ − sin θ 0 sin θ 0 cos θ 0 0 cos2 s cosh r + (sin s)2 − sin 2s sinh2( r 2) cos s sinh r − sin 2s sinh2( r 2) cos s sinh r cos2 s + cosh r sin2 s − sin s sinh r − sin s sinh r cosh r 1   et 0 0 et 0 0 0 0 et  36 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA and the dual action of the group can now be computed in a very precise manner. Let be a Zariski open subset of the dual of V. We shall prove that Ω⋆ is SO0 (2, 1) exp RE3- invariant and Σ =(cid:8)ω, ξ+, ξ−(cid:9) where ω = 0 1 is a cross-section for the dual orbits in Ω⋆. First, we check that the dual orbit of ξ± is equal to 0 0 0 0 x y Ω⋆ =  −1  z  : x2 + y2 − z2 6= 0 0  , ξ+ = 1  , ξ− =  : t, r, s ∈ R = z  : x2 + y2 − z2 < 0  z  : x2 + y2 − z2 > 0 et sin s sinh r −et cos s sinh r Ω± =  hz  : x2 + y2 − z2 = −1 Ω′ =  ±et cosh r hx hy x y x y . . . Next, let Ω = Ω+ ∪ Ω−. Then Ω = [h>0   The corresponding stabilizer subgroup for ξ± in SO0(2, 1) exp RE3 is equal to a maxi- mal torus subgroup of SO0(2, 1). Thus the stabilizer subgroup for every element in Ω is compact. Finally, the orbit of ω is given by However, every element of the one-parameter group exp (RA) stabilizes ω. Finally, our claim follows from the fact that Ω⋆ = Ω ∪ Ω′. Next, let π be a two-dimensional irreducible representation of sl(2, R) acting on a vector space V. Since the algebra generated by π(H), π(X) is solvable then there exists a basis for the vector space V such that π (H) =(cid:20) α 0 α + 2 (cid:21) , π (X) =(cid:20) 0 0 1 0 (cid:21) 0 and v is an eigenvector for π (H) with corresponding eigenvalue α. Now, since {v, π (X) v} is a linearly independent set, it is a basis. With respect to this basis, we obtain and clearly 0 π (H) =(cid:20) −1 0 π (Y ) − π (X) =(cid:20) 0 1 (cid:21) , π (X) =(cid:20) 0 0 −1 0 (cid:21) ,(cid:20) 1 1 0 (cid:21) , π (Y ) =(cid:20) 0 1 0 0 (cid:21) 0 −1 (cid:21) ,(cid:20) 0 1 0 0 (cid:21) 1 0 The stabilizer subgroups corresponding to the elements ξ+, ξ− are not trivial and not contained in the maximal torus subgroup of H (and thus not compact.) Indeed the one-parameter group 1 0 −1  . Σ =(cid:8)ξ+, ξ−(cid:9) where ξ+ = 1 0 1  , ξ− = 0 0 1  : t ∈ R 1 0 0 t 1 0   CLASSIFYING WAVELET TRANSFORMS IN 3D 37 form a basis for π (sl(2, R)) . Let τ be a three-dimensional reducible representation of sl (2, R) ⊕ R. Then up to conjugation, the image of sl (2, R) ⊕ R via this representation is spanned by matrices K = 0 1 0 −1 0 0 0 0 0  , A = 0 1 0 0 −1 0 0 0 0  , N = 0 1 0 0 0 0 0 0 0  , D = λ 0 0 0 λ 0 0 0 β  where λ, β are some real numbers. We observe that the matrices K, A, N span an algebra which is isomorphic to sl (2, R). Next, we consider the dual action of H = SL (2, R) exp (RD) . Using the Cartan decomposition of sl (2, R) , we write elements of H as exp (θK) exp (t1A + t2N + t3D) where θ ∈ [0, 2π), and t1, t2, t3 are some real num- bers. Assuming that β 6= 0, there are two open orbits, and a cross-section for these orbits is given by stabilizes ξ+ and ξ−. Thus the dilation group SL (2, R) exp (RD) is not irreducibly ad- missible. Assume that β = 0 (thus λ 6= 0.) There are no open orbits associated with this dilation group. Thus the corresponding dilation groups are not irreducibly admissible. Case (B) Let h be a Lie algebra of matrices of order three which is isomorphic to the direct sum of so (2, R) ⊕ R. Up to conjugation, h is spanned by matrices X1 = 0 −1 0 1 0 0 0 0 0  , X2 = 0 0 0 0 0 −1 0 1 0  , X3 = 0 0 −1 0 0 0 1 0 0  , X4 = 1 0 0 0 1 0 0 0 1  Now, let H be a Lie group with Lie algebra h and let H0 be the connected compo- nent of the identity. Then H0 = SO (3, R) exp (RX4) . Next, let v be a unit vector. Clearly, there exists a pair of angles (θ, φ) such that v = exp (θX1) exp (φX2) e3. Thus, for an arbitrary matrix M in the orthogonal group SO (3, R) such that Me3 = v, Me3 = exp (θX1) exp (φX2) e3 and there exists an angle ω such that M = exp (θX1) exp (φX2) exp (ωX1) . Thus an element of SO (3, R) exp (RX4) can be written as follows:  cos θ − sin θ 0 sin θ 0 cos θ 0 0 1  0 0 1 0 cos φ − sin φ 0 sin φ cos φ  cos ω − sin ω 0 sin ω 0 cos ω 0 0 1  et 0 0 et 0 0 0 0 et  . 38 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Computing the dual action of SO (3, R) exp (RX4), we obtain a single open orbit (the punctured space) and the stabilizer for every element in V ∗ is compact. Thus, this dilation group is irreducibly admissible. 6. Compactly supported wavelets and atoms In this section we wish to establish the existence of compactly supported admissible wavelets for all irreducibly admissible subgroups H < GL(3, R). For most groups, we will be able to show a much stronger statement, namely the existence of compactly supported atoms ψ associated to H. These functions are distinguished by the fact that for all suffi- ciently dense lattices Γ ⊂ Rd and all sufficiently uniformly dense and uniformly discrete subsets Λ ⊂ H, the family (π(hx, h)ψ)x∈Γ,h∈Λ is a frame, and furthermore the decay of the frame coefficients of f ∈ L2(Rd) characterizes whether f belongs to a suitable coorbit space Co(Y ), where Y is a Banach space of functions on G. It is known that suitably chosen bandlimited Schwartz functions can serve as atoms [22], however the existence of atoms that are compactly supported in space is not generally established. In fact atoms are admissible vectors, i.e., Wψ is a well-defined isometry for each admissible vector ψ, and it is currently not even known whether there exist compactly supported vectors for every irreducibly admissible matrix group. The recent papers [23, 25] contain sufficient criteria for the existence of such atoms, and below we will prove that these criteria are fulfilled for all cases except case (2.l). Note that this does not exclude the existence of compactly supported atoms for this case, just that the methods from [23, 25] do not suffice for a positive answer. For the exceptional case (2.l), we will however establish the existence of a compactly supported admissible vector. 6.1. Coorbit spaces and atomic decompositions. Coorbit space theory addresses two important questions in generalized wavelet analysis: The consistent definition of spaces of sparse signals, and discretization of the continuous wavelet transforms. By definition, sparse signals should be described by fast decay of their wavelet transforms. One possible way of quantifying this decay behavior is to impose a weighted mixed Lp- norm on the wavelet coefficients: Nice signals are those for which the wavelet coefficient decay is sufficiently fast to guarantee weighted integrability. One of the starting points of coorbit space theory was the realization that the scale of Besov spaces can be understood in precisely these terms [15]. Coorbit space theory provides the same type of results for a much larger setup, including continuous wavelet transforms associated to irreducibly admissible matrix groups H. We will now briefly describe the main ingredients of this theory, for a general irreducibly admissible matrix group H < GL(d, R). We start out by fixing a Banach function space Y on G = Rd ⋊ H. A relevant choice from the approximation-theoretic point of view is Lp(G), with 1 ≤ p < 2; see Remark 6.4. Furthermore, we assume that Y is strongly invariant under both left and right translations of G, and that it is solid: For every pair F, H of measurable functions on G with F ∈ Y and H ≤ F pointwise a.e., it follows that H ∈ Y and kHkY ≤ kF k. CLASSIFYING WAVELET TRANSFORMS IN 3D 39 Definition 6.1. (a) A weight on G is a continuous function v : G → R+ satisfying the submultiplicativity condition v((x1, h1)(x2, h2)) ≤ v(x1, h1)v(x2, h2) . (b) A weight v on G is called polynomially bounded if it satisfies v(x, h) ≤ C(1 + x)s(1 + khk)r , for suitable r, s ≥ 0. (c) Let Y be a solid Banach function space on G. A weight v0 is called control weight for Y if it is bounded from below, and satisfies v0(x, h) = ∆G(x, h)−1v0((x, h)−1) , as well as max(cid:0)kL(x,h)±1kY →Y , kR(x,h)kY →Y , kR(x,h)−1kY →Y ∆G(x, h)−1(cid:1) ≤ v0(x, h) where L(x,h) and R(x,h) denote left and right translation by (x, h) ∈ G. Note that two different notions of weights exist in this paper. Each is established ter- minology within its context, and for the remainder of the paper, only the definition of weights according to 6.1 will play a role. For these reasons, we have refrained to further differentiate between the two, and prefer to live with a slight conflict of terminology as a consequence. We next define the set of atoms. The role of the weight v0 in the definition can be read as a compatibility condition incurred by the choice of the function space Y : The faster v0 grows, the more restrictive the definition of Bv0 becomes. This is comparable to the vanishing moment and smoothness conditions imposed on wavelet ONB's characterizing homogeneous Besov spaces Bα p,q, which also depend on the parameters p, q, α. Definition 6.2. Let Y denote a solid Banach function space on the locally compact group G, U ⊂ G a compact neighborhood of the identity, and F : G → C. We let (cid:0)MR U F(cid:1) (x) = sup y∈U F (yx−1) denote the right local maximum function of F with respect to U. Given a weight v0 on G, let Now 0 6= ψ ∈ L2(Rd) is called a v0-atom, if L1 v0(G) = {F : G → C : F v0 ∈ L1(G)} . MR U (Wψψ) ∈ L1 v0(G) , and we denote the set of all such ψ by Bv0. Remark 6.3. We note that Wψψ ≤ MR U (Wψψ) implies that for every ψ ∈ Bv0, Wψψ ∈ L1 v0(G) ∩ L∞(G) ⊂ L2(G) , since v0 is assumed bounded from below. Hence every ψ ∈ Bv0 is an admissible vector. 40 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA Now the central results of [14, 15, 16, 28] with respect to consistency and discretization can be briefly summarized as follows: (1) Given a suitable Banach function space Y on G with control weight v0, fix a v0}, and define H1,v0 = {f ∈ 1,v0 denote the space of conjugate-linear bounded 1,v0 can be defined nonzero ψ ∈ Av0 := {ψ ∈ L2(Rd) : Wψψ ∈ L1 L2(Rd) : Wψf ∈ L1 functionals on H1,v0. Then the wavelet transform Wψf of f ∈ H∼ by canonical extension of the formula for L2-functions. v0}. Let H∼ (2) Fix a nonzero ψ ∈ Bv0, and define CoY = {f ∈ H∼ 1,v0 : Wψf ∈ Y } , with the obvious norm kf kCoY = kWψf kY . Then CoY is a well-defined Banach space, and it is independent of the choice of ψ ∈ Bv0 \ {0}: Changing ψ results in an equivalent norm, and in the same coorbit space CoY . (3) Fix a nonzero ψ ∈ Bv0. Then the coorbit space norm is equivalent to the discretized , for all suitably dense and discrete subsets Z ⊂ G, with a suitably norm kWψf ZkYd defined Banach sequence space Yd. This also gives rise to atomic decompositions, i.e., discrete systems of wavelets that provide frame decompositions converging not just in L2, but also in the coorbit space norms. Remark 6.4. Readers who prefer explicitness over generality may take the spaces Y = Lp(G) as template for the general setup, with p ∈ [1, ∞]. Here the above statements can be further motivated by non-linear approximation. To begin with, the associated control weight can be taken as v0(x, h) = max(1, ∆G(h)), valid for all p ∈ [1, ∞] [23, Lemma 2.3 and subsequent remarks]. The benefit of using a common weight is that it gives rise to atomic decompositions that converge simultaneously in a whole range of coorbit spaces. We will now shortly explain the relevance of Co(Lp(G)) from the point of view of non- linear approximation: Let 1 ≤ p < 2, and let ψ ∈ Bv0. Then there exists a discrete subset Λ ⊂ G such that (π(λ)ψ)λ∈Λ is a frame of L2(G), which means that there exist constants 0 < A ≤ B, depending on Λ and ψ, such that for all f ∈ L2(R3), Akf k2 2 ≤ X(x,h)∈Λ hf, π(hx, h)ψi2 ≤ Bkf k2 2 . In this case, there exists a dual frame that can be used to write down a discrete version of the wavelet inversion formula, converging unconditionally in L2. But atoms allow much more than just frame expansions converging in L2. As a matter of fact, the elements f ∈ Co(Lp) are characterized by any of the following equivalent statements: (a) The discrete coefficient sequence (hf, π(λ)ψ)λ∈Λ is p-summable. (b) f has an expansion f = X(x,h)∈Λ cλπ(λ)ψ , with p-summable coefficients (cλ)λ∈Λ. CLASSIFYING WAVELET TRANSFORMS IN 3D 41 The second property has interesting consequences for nonlinear approximation: If we let : cλ ∈ C , Λ′ ≤ n) . (5) En(f ; (π(λ)ψ)λ) = inf((cid:13)(cid:13)(cid:13)(cid:13)(cid:13) f −Xλ∈Λ′ cλπ(λ)ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H then approximating f using the n largest coefficients in (b) leads to the following estimate for the approximation error En(f, (π(λ)ψ)λ) ≤ Cǫkf kCo(Lp)n−p/2+ǫ , valid for any ǫ > 0 and f ∈ Co(Lp). We refer to [23, Section 1.1] for more details. These observations clearly justify the interpretation of Co(Lp) as a space of sparse signals. It should also be mentioned that the coorbit scheme extends to the quasi-Banach setting; see [37, 38]. In particular, Y = Lp(G) with p < 1 can also be employed, which leads to even sharper decay estimates for nonlinear approximation. Note however that the vanishing moment conditions for atoms that we will employ below have not yet been established for quasi-Banach spaces. Note that the discussion in this subsection is devoid of meaning if the set Bv0 is empty. Clearly, checking the condition in Definition 6.2 poses a significant challenge, since it re- quires estimating Wiener amalgam space norms of continuous wavelet transforms. How- ever, for the setting of generalized wavelet transforms, the following result is obtained by combining [22, Theorem 2.1] with [22, Lemma 2.7]. c (O). Then ψ ∈ Bv0. Theorem 6.5. Assume that 0 6= bψ ∈ C ∞ 6.2. Vanishing moment conditions for admissible vectors and atoms. The pre- vious subsection was mainly intended to motivate the interest in the set Bv0 of atoms. At this point the only easily accessible atoms are compactly supported in frequency, and the existence of atoms that are compactly supported in space is open. Sufficient criteria for this latter class have been formulated in [23, 25], and we will now briefly recount these conditions. We denote the unique open dual orbit of H by O ⊂ Rd, and its complement by Oc. We aim for statements that a function ψ is admissible (or an atom) provided it has sufficient decay, sufficient smoothness, and sufficiently many vanishing moments in Oc. Only the last condition actually reflects our specific choice of the matrix group H, and it is rigorously defined as follows: Definition 6.6. Let r ∈ N be given. f ∈ L1(Rd) has vanishing moments in Oc of order r if all distributional derivatives ∂αbf with α ≤ r are continuous functions, and all derivatives of degree α < r are vanishing on Oc. Note that under suitable integrability conditions on ψ, the vanishing moment conditions are equivalent to ∀α < r, ∀ξ ∈ Oc xαψ(x)e−2πihξ,xidx = 0 . : ZRd It is important to note that [22, Lemma 4.1] guarantees, for any r > 0, the existence of ψ ∈ C ∞ c (Rd) with vanishing moments of order r. Hence sufficient vanishing moment criteria for atoms guarantee the existence of compactly supported atoms. 42 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA The central tool for the formulation of vanishing moment conditions for admissible vectors and atoms is an auxiliary function A : O → R+ defined as follows: Given any point ξ ∈ O, let dist(ξ, Oc) denote the minimal distance of ξ to Oc, and define A(ξ) := min dist(ξ, Oc) 1 +pξ2 − dist(ξ, Oc)2 , 1 1 + ξ! . By definition, A is a continuous function with A(·) ≤ 1, and it is extended continuously to all of Rd by setting it to zero on Oc. If η ∈ Oc denotes an element of minimal distance to ξ, the fact that R+ · η ⊂ Oc then entails that η and ξ − η are orthogonal with respect to the standard scalar product on Rd, and we obtain the more transparent expression A(ξ) = min(cid:18)ξ − η 1 + η , 1 1 + ξ(cid:19) . We can now formulate sufficient criteria for the existence of compactly supported ad- missible functions and atoms. Note that part (a) provides access to compactly supported atoms for coorbit spaces associated to weighted-mixed Lp spaces with polynomial weights. Proposition 6.7. Assume that v0 is a polynomially bounded weight on G. Define AH : H → R+ by AH (h) = A(hT ξ) . (a) If there exists an exponent e > 1 and a constant C > 0 such that, for all h ∈ H, then ψ ∈ C ∞ c (Rd) with vanishing moments of order r > (d + e)/2 is admissible. (b) If there exists an exponent e > 1 and a constant C > 0 such that, for all h ∈ H, ∆G(h)AH (h)e ≤ C kh±1kAH (h)e ≤ C then there exists an r > 0, explicitly computable from v0 and d, such that every compactly supported smooth function sufficiently many vanishing moments of order r is contained in Bv0. Note that the condition in Part (b) does not depend on the precise choice of matrix norms. Proof. For part (a), we recall the general admissibility condition derived in [17]: Fixing a vector ξ0 ∈ O, the fact that the continuous homomorphism ∆G restricted to the compact stabilizer Hξ0 is constant implies that letting Φ : O → R+, with Φ(hT ξ0) = ∆G(h), yields a well-defined continuous function. Furthermore, Φ has the property that 0 6= ψ is admissible iff Our assumption on e and C yield ZO bψ(ξ)2Φ(ξ)dξ < ∞ . Φ(ξ)A(ξ)e ≤ C . If ψ is a compactly supported C ∞ functions with vanishing moments of order r, we get by [22, Lemma 3.4] that bψ(ξ) ≤ C ′bψr,mA(ξ)r . CLASSIFYING WAVELET TRANSFORMS IN 3D 43 Here we used the Schwartz norm f r,m = sup (1 + x)m∂αf (x) . x∈Rd,α≤r Hence we obtain ZO ψ(ξ)2Φ(ξ)dx ≤ C ′ψr,mZO ≤ CC ′ψr,mZO ≤ CC ′ψr,mZRd A(ξ)2rΦ(ξ)dξ A(ξ)2r−edξ (1 + ξ)e−2rdξ < ∞ , as soon as 2r − e > d. Part (b) follows combining [23, Theorem 3.4] with [25, Lemma 2.8, Theorem 2.12]. (cid:3) 6.3. Proof of Theorem 2.1 (b), (c). We will verify the conditions of Proposition 6.7. First note that the existence of compactly supported admissible vectors or atoms is clearly conjugation invariant: If the admissible matrix groups H1 and H2 are related via H2 = gH1g−1, with g ∈ GL(3, R), then every such vector for H2 is obtained from a vector with analogous properties by dilation with g, which preserves compact support. Furthermore, we may restrict attention to the connected subgroup H0. sufficient to establish the following: For each open H0-orbit O′, pick ξ′ Assume that there exists a map σ : O′ → H0 satisfying, for all ξ ∈ O′, In fact, it is 0 ∈ O′ arbitrary. (i.e., σ is a cross-section with respect to the dual action), as well as A(ξ)ekσ(ξ)±1k ≤ C , σ(ξ)T ξ′ 0 = ξ , with constants e, C independent of ξ. We then get for all h ∈ H0 that hσ(hT ξ′ 0)−1 ∈ Hξ0 , and by assumption, the stabilizer is compact. Hence there exists a constant K such that But this yields, for all h ∈ H0, that K := sup h∈H(cid:13)(cid:13)(cid:13)(cid:0)hσ(hT ξ′ A(hT ξ′ 0)ekh±1k = A(σ(hT ξ′ ≤ A(σ(hT ξ′ 0))ekh±1k 0)ekσ(hT ξ′ ≤ CK , 0)−1(cid:1)±1(cid:13)(cid:13)(cid:13) < ∞. 0)k(cid:13)(cid:13)(cid:13)(cid:0)hσ(hT ξ′ 0)−1(cid:1)±1(cid:13)(cid:13)(cid:13) by assumption on σ. Now if H is an irreducibly admissible matrix group containing H0, then H/H0 is finite, and the single open H-orbit is just the union of open H0-orbits O′. I.e., each h ∈ H can be written as h = gih0 with gi ∈ {g1, . . . , gk}, the latter being a fixed system of representatives mod H0, and h0 ∈ H0. Using a bound on the norms of the g±1 , it is now easy to extend the desired estimate from H0 to H. i 44 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA A similar reasoning allows to verify the condition of Proposition 6.7(a) by focusing on the connected component, as well; just replace all occurrences of k · k in the above argument by det(·). In view of these observations, it remains to go through the list in 2.1(a) and verify the conditions of Proposition 6.7. In the following, we will argue via 6.7 (a) for case (2.l), and via 6.7(b) for the remaining cases, and obtain 2.1(b),(c) (recall Remark 6.3). 6.3.1. Known results: Cases (1) and (3). The existence of compactly supported atoms for abelian admissible groups acting in arbitrary dimensions was shown in [25]. This takes care of the cases (1.a) through (1.e). For the similitude group in arbitrary dimensions, this fact was established in [23, Theorem 3.2]. 6.3.2. The cases (2.a) through (2.i). In all these cases, one first verifies easily that O± = H T 0  A(ξ) = min(cid:18) ±1 0 0  = R± × R2 1 + ξ(cid:19) . ξ1 are the only open orbits, and we obtain the envelope function 1 , 1 + [ξ2, ξ3] Furthermore, the dual action on the open orbits is free for all cases involved, and thus the cross-section σ : O± → H0 associated to ξ0 = [±1, 0, 0]T is uniquely determined. Since the entries of hT [1, 0, 0]T are just the entries of the first row of h, the computation of σ(ξ) is equivalent to the task of computing the remaining entries of h ∈ H0 from its first row entries. In each of the cases (2.d)-(2.i), one easily reads off the formulas for the group elements of H0 that the entries of σ(ξ) depend polynomially on ξ2, ξ3 and on arbitrary nonzero powers of ξ1 and ln(ξ1). Hence the estimates A(ξ)ξ1±1 ≤ 1 , A(ξ)ξ2 ≤ 1 , A(ξ)ξ3 ≤ 1 as well as limx→0 x ln(x) = 0 yield boundedness of A(ξ)ekσ(ξ)k ≤ C for sufficiently large e, C. For the estimate of the norm of the inverse, we need the additional observation that det(σ(ξ)) = ξ1α for a suitable real number α, which follows from the fact that σ(ξ) is upper triangular, with diagonal entries depending only on ξ1. By Cramer's rule, the entries of σ(ξ)−1 are polynomials in det(σ(ξ))−1 and the entries of σ(ξ). Both can be controlled by suitable powers of A(ξ), hence we get A(ξ)e′ The treatment of case (2.a) follows similar lines. Again, the action is free, and thus the cross-section is unique. Here, the moduli of the remaining entries of σ(ξ) are less than or equal to ξ1, and thus A(ξ)kσ(ξ)k ≤ C follows. The estimate for the inverse again follows from this, Cramer's rule, and the dependence of det(σ(ξ)) on ξ1. For the remaining cases (2.b) and (2.c), note that the cross-section for (2.a) also works for these cases; to treat case (2.b), set a = 1 in case (2.a). Hence we are done here as well. kσ(ξ)−1k ≤ C ′ as well. CLASSIFYING WAVELET TRANSFORMS IN 3D 45 6.3.3. The cases (2.j)-(2.k) and (2.m). For each of these groups, we obtain the four open orbits and the associated envelope function Again the action is free, and thus the associated cross-section is unique. An arbitrary matrix h ∈ H0 has the structure 0  ±1 0 ±1  = R± × R × R± , 1 + ξ(cid:19) . 1 + [ξ1, ξ2] ξ3 1 , , ξ1 1 + [ξ2, ξ3] O±,± = H T A(ξ) = min(cid:18) h = hT ǫ1 0 ǫ2 h1,1 h1,2 0 h2,2 0 0 0 0 h3,3  = ǫ1h1,1 ǫ1h1,2 ǫ2h3,3   . and thus, with ǫ1, ǫ2 ∈ {±1}, we get Thus the computation of σ(ξ) boils down to computing h2,2 from h1,1, h1,2, h3,3, and in each of the cases under consideration, h2,2 is seen to be a product of powers of h1,1 and h3,3. Now the estimates A(ξ)ξ1±1 ≤ 1 , A(ξ)ξ3±1 ≤ 1 , A(ξ)ξ2 ≤ 1 yield A(ξ)ekσ(ξ)k ≤ C for suitably large e, C. The estimate for kσ(ξ)−1k again follows from this fact, Cramer's rule, and the fact that the determinant of h is a product of powers of h1,1 and h3,3. 6.3.4. The case (2.n). We obtain the four open orbits and the associated envelope function O±,± = H T 0  ±1 ±1 0  = R± × R± × R , 1 + ξ(cid:19) . 1 + [ξ1, ξ3] ξ2 1 , , ξ1 1 + [ξ2, ξ3] ξ3 − δ1 ln(ξ1/ξ2) − δ2ξ2 ln(ξ2) δ1 ln(ξ1/ξ2) + δ2ξ2 ln(ξ2) ξ2  . A(ξ) = min(cid:18) σ(ξ) = ξ1 0 0 0 ξ2 0 Again the action is free, and the associated cross-section is determined as Again, all entries of σ(ξ) are controlled by suitable positive powers of ξ1±1, ξ2±1, ξ3, and A(ξ)ekσ(ξ)k ≤ C for suitably large e, C follows. The estimate for the inverse now follows in the usual way from an appeal to Cramer's rule and the observation that det(σ(ξ)) = ξ1 ξ22. 46 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA 6.3.5. The exceptional case (2.l). We obtain the two open orbits and the associated envelope function O±,± = H T 0  A(ξ) = min(cid:18) ±1 0 0  = R± × R2 , 1 + ξ(cid:19) . ξ1 1 , 1 + (ξ2, ξ3) This action is free, and the associated cross-section is determined as σ(ξ) = ξ1 0 0 ξ2 exp(−ξ3/ν1)ξ1 exp(ν2ξ1 ln(ξ1)/ν1) 0 ξ3 0 ξ1  . Now one easily verifies that, for all e ≥ 0, A(σ(1, 0, ξ3))ekσ(1, 0, ξ3)k ≥ CA(σ(1, 0, ξ3))e exp(−ξ3/ν1) → ∞ , as sign(ν1)ξ3 → −∞. This shows that the required condition for vanishing moment criteria are in fact violated by this group. Hence the criteria for the existence of compactly supported atoms are not applicable. We now show the existence of compactly supported admissible vectors. For this purpose, we need to compute the modular function of H0. Writing h(a, b, t) = b 0 ab 0 0 t −ν1 ln a + ν2b ln b 0 b  ∈ H0 for (a, b, t) ∈ R+ × R+ × R, we see that the subsets H1 = {h(1, b, 0) : b ∈ R+} , H2 = {h(a, 1, t) : a ∈ R+, t ∈ R} are two closed subgroups that commute elementwise. isomorphic to R. Furthermore, the computation I.e., H0 ∼= H1 × H2, and H1 is h(a, 1, t)h(a′, 1, t′) = h(t′ + a′t, tt′) shows that H2 is (anti)-isomorphic to the well-known ax + b-group. Hence we get that the left Haar measure of H2 is given by dt da a , and the modular function of H2 is ∆H2(h(a, 1, t)) = a. In summary, this gives and thus ∆H0(h(a, b, t)) = ∆H1(h(1, b, 0))∆H2(h(a, 1, t)) = a , ∆G(h(a, b, t)) = ∆H0(h(a, b, t)) det(h(a, b, t)) = a a b3 = b−3 . But this implies for the function Φ entering the admissibility condition that Φ(ξ) = ∆G(σ(ξ1, ξ2, ξ3)) = ∆G(h(exp(−ξ3/ν1) exp(ν2ξ1 ln(ξ1)/ν1), ξ1, ξ2)) = ξ1−3 . In particular, A(ξ)3Φ(ξ) ≤ 1, which proves the existence of compactly supported admis- sible vectors. CLASSIFYING WAVELET TRANSFORMS IN 3D 47 References [1] G. S. Alberti, L. Balletti, F. De Mari, and E. De Vito. Reproducing subgroups of Sp(2, R). Part I: Algebraic classification. J. Fourier Anal. Appl., 19(4):651 -- 682, 2013. [2] Giovanni S. Alberti, Filippo De Mari, Ernesto De Vito, and Lucia Mantovani. Reproducing subgroups of Sp(2, R). Part II: admissible vectors. Monatsh. Math., 173(3):261 -- 307, 2014. [3] Didier Arnal, Bradley Currey, and Vignon Oussa. Characterization of regularity for a connected Abelian action. Monatsh. Math., 180(1):1 -- 37, 2016. [4] David Bernier and Keith F. Taylor. Wavelets from square-integrable representations. SIAM J. Math. Anal., 27(2):594 -- 608, 1996. [5] J. Bruna, J. Cufi, H. Fuhr, and M. Mir´o. Characterizing abelian admissible groups. J. Geom. Anal., 25(2):1045 -- 1074, 2015. [6] Emmanuel Candes, Laurent Demanet, David Donoho, and Lexing Ying. Fast discrete curvelet trans- forms. Multiscale Modeling & Simulation, 5(3):861 -- 899, 2006. [7] Brad Currey and Tom McNamara. Decomposition and admissibility for the quasiregular representa- tion for generalized oscillator groups. In Radon transforms, geometry, and wavelets, volume 464 of Contemp. Math., pages 51 -- 73. Amer. Math. Soc., Providence, RI, 2008. [8] Bradley N Currey. Admissibility for a class of quasiregular representations. Canadian Journal of Mathematics, 59(5):917 -- 942, 2007. [9] Wojciech Czaja and Emily J. King. Isotropic shearlet analogs for L2(Rk) and localization operators. Numer. Funct. Anal. Optim., 33(7-9):872 -- 905, 2012. [10] Stephan Dahlke, Soren Hauser, and Gerd Teschke. Coorbit space theory for the Toeplitz shearlet transform. Int. J. Wavelets Multiresolut. Inf. Process., 10(4):1250037, 13, 2012. [11] Stephan Dahlke, Gitta Kutyniok, Gabriele Steidl, and Gerd Teschke. Shearlet coorbit spaces and associated Banach frames. Appl. Comput. Harmon. Anal., 27(2):195 -- 214, 2009. [12] Stephan Dahlke, Gabriele Steidl, and Gerd Teschke. Multivariate shearlet transform, shearlet coorbit spaces and their structural properties. In Shearlets, Appl. Numer. Harmon. Anal., pages 105 -- 144. Birkhauser/Springer, New York, 2012. [13] Hans Feichtinger and Felix Voigtlaender. From Frazier-Jawerth characterizations of Besov spaces to wavelets and decomposition spaces. Preprint, available under http://arxiv.org/abs/1606.04924, 2016. [14] Hans G. Feichtinger and Karlheinz Grochenig. A unified approach to atomic decompositions via integrable group representations. In Function spaces and applications (Lund, 1986), volume 1302 of Lecture Notes in Math., pages 52 -- 73. Springer, Berlin, 1988. [15] Hans G. Feichtinger and Karlheinz Grochenig. Banach spaces related to integrable group represen- tations and their atomic decompositions. I. J. Funct. Anal., 86(2):307 -- 340, 1989. [16] Hans G. Feichtinger and Karlheinz Grochenig. Banach spaces related to integrable group represen- tations and their atomic decompositions. II. Monatsh. Math., 108(2-3):129 -- 148, 1989. [17] Hartmut Fuhr. Wavelet frames and admissibility in higher dimensions. J. Math. Phys., 37(12):6353 -- 6366, 1996. [18] Hartmut Fuhr. Continuous wavelet transforms with abelian dilation groups. J. Math. Phys., 39(8):3974 -- 3986, 1998. [19] Hartmut Fuhr. Continuous wavelets transforms from semidirect products. Cienc. Mat. (Havana), 18(2):179 -- 190, 2000. [20] Hartmut Fuhr. Abstract harmonic analysis of continuous wavelet transforms, volume 1863 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2005. [21] Hartmut Fuhr. Generalized Calder´on conditions and regular orbit spaces. Colloq. Math., 120(1):103 -- 126, 2010. [22] Hartmut Fuhr. Coorbit spaces and wavelet coefficient decay over general dilation groups. Trans. Amer. Math. Soc., 367(10):7373 -- 7401, 2015. [23] Hartmut Fuhr. Vanishing moment conditions for wavelet atoms in higher dimensions. Adv. Comput. Math., 42(1):127 -- 153, 2016. 48 BRADLEY CURREY, HARTMUT F UHR, VIGNON OUSSA [24] Hartmut Fuhr and Matthias Mayer. Continuous wavelet transforms from semidirect products: cyclic representations and Plancherel measure. J. Fourier Anal. Appl., 8(4):375 -- 397, 2002. [25] Hartmut Fuhr and Reihaneh Raisi-Tousi. Simplified vanishing moment criteria for wavelets over general dilation groups, with applications to abelian and shearlet dilation groups. To appear in Appl. Comp. Harm. Anal., doi:10.1016/j.acha.2016.03.003, 2016. [26] Hartmut Fuhr and Felix Voigtlaender. Wavelet coorbit spaces viewed as decomposition spaces. J. Funct. Anal., 269(1):80 -- 154, 2015. [27] Daryl Geller and Azita Mayeli. Continuous wavelets and frames on stratified Lie groups. I. J. Fourier Anal. Appl., 12(5):543 -- 579, 2006. [28] Karlheinz Grochenig. Describing functions: atomic decompositions versus frames. Monatsh. Math., 112(1):1 -- 42, 1991. [29] Karlheinz Grochenig, Eberhard Kaniuth, and Keith F. Taylor. Compact open sets in duals and projections in L1-algebras of certain semi-direct product groups. Math. Proc. Cambridge Philos. Soc., 111(3):545 -- 556, 1992. [30] Anthony W. Knapp. Lie groups beyond an introduction, volume 140 of Progress in Mathematics. Birkhauser Boston, Inc., Boston, MA, 1996. [31] Richard S. Laugesen, Nik Weaver, Guido L. Weiss, and Edward N. Wilson. A characterization of the higher dimensional groups associated with continuous wavelets. J. Geom. Anal., 12(1):89 -- 102, 2002. [32] Malcolm AH MacCallum. On the classification of the real four-dimensional lie algebras. In On Ein- steins path, pages 299 -- 317. Springer, 1999. [33] Margarida Mir´o. Funcions admissibles i ondetes ortonormals a Rn. PhD thesis, Universitat Aut´onoma de Barcelona, 2010. [34] Romain Murenzi. Wavelet transforms associated to the n-dimensional Euclidean group with dilations: signal in more than one dimension. In Wavelets (Marseille, 1987), Inverse Probl. Theoret. Imaging, pages 239 -- 246. Springer, Berlin, 1989. [35] P. S. Negi and D. Labate. 3-d discrete shearlet transform and video processing. IEEE Transactions on Image Processing, 21(6):2944 -- 2954, June 2012. [36] Vignon Oussa. Admissibility for quasiregular representations of exponential solvable Lie groups. Colloq. Math., 131(2):241 -- 264, 2013. [37] Holger Rauhut. Coorbit space theory for quasi-Banach spaces. Studia Math., 180(3):237 -- 253, 2007. [38] Felix Voigtlaender. Embedding Theorems for Decomposition Spaces, with Applications to Wavelet Coorbit Spaces. PhD thesis, RWTH Aachen, 2015. Department of Mathematics and Statistics, Saint Louis University, St. Louis. MO 63103, U.S.A. E-mail address: [email protected] Lehrstuhl A fur Mathematik, RWTH Aachen University, D-52056 Aachen, Germany E-mail address: [email protected] Department of Mathematics, Bridgewater State University, Bridgewater, MA 02324, U.S.A. E-mail address: [email protected]
1208.2288
1
1208
2012-08-10T21:01:43
Norm-constrained determinantal representations of polynomials
[ "math.FA" ]
For every multivariable polynomial $p$, with $p(0)=1$, we construct a determinantal representation $$p=\det (I - K Z),$$ where $Z$ is a diagonal matrix with coordinate variables on the diagonal and $K$ is a complex square matrix. Such a representation is equivalent to the existence of $K$ whose principal minors satisfy certain linear relations. When norm constraints on $K$ are imposed, we give connections to the multivariable von Neumann inequality, Agler denominators, and stability. We show that if a multivariable polynomial $q$, $q(0)=0,$ satisfies the von Neumann inequality, then $1-q$ admits a determinantal representation with $K$ a contraction. On the other hand, every determinantal representation with a contractive $K$ gives rise to a rational inner function in the Schur--Agler class.
math.FA
math
Norm-constrained determinantal representations of multivariable polynomials Anatolii Grinshpan, Dmitry S. Kaliuzhnyi-Verbovetskyi and Hugo J. Woerdeman Mathematics Subject Classification (2010). 15A15; 47A13; 47A20; 47A48. Keywords. Determinantal representation; multivariable polynomial; d- variable Schur–Agler class; Agler denominator; semi-stable polynomial. Abstract. For every multivariable polynomial p, with p(0) = 1, we con- struct a determinantal representation p = det(I − KZ), where Z is a diagonal matrix with coordinate variables on the diagonal and K is a complex square matrix. Such a representation is equivalent to the existence of K whose principal minors satisfy certain linear rela- tions. When norm constraints on K are imposed, we give connections to the multivariable von Neumann inequality, Agler denominators, and stability. We show that if a multivariable polynomial q, q(0) = 0, sat- isfies the von Neumann inequality, then 1 − q admits a determinantal representation with K a contraction. On the other hand, every determi- nantal representation with a contractive K gives rise to a rational inner function in the Schur–Agler class. 1. Introduction Our object of study is determinantal representations p(z) = det(In − KZn), (1.1) ··· + nd, Zn = Ld for a d-variable polynomial p(z), z = (z1, . . . , zd), with p(0) = 1. Here n = (n1, . . . , nd) is in the set Nd 0 of d-tuples of nonnegative integers, n = n1 + i=1 ziIni , and K is a complex square matrix. It is of interest of how and to what extent, the algebraic and operator-theoretic properties of the polynomial correspond to the size and norm of the matrix K of its representation. The authors were partially supported by NSF grant DMS-0901628. 2 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman Various determinantal representations of polynomials have been stud- ied, often for polynomials over the reals: see a recent overview article [27], together with bibliography, and also [21]. The particular form of (1.1) has ap- peared before, for instance, in [6]. An important early result on two-variable polynomials was obtained by A. Kummert [19, Theorem 1]. Given a d-variable polynomial p(z), p(0) = 1, we consider the question of whether it can be represented in the form (1.1) for some n ∈ Nd 0 and some n×n complex matrix K, possibly subject to a constraint. It will be shown that the unconstrained version of this question can always be answered in the affirmative (Section 2). The problem of minimizing the operator norm of K over all representations (1.1) of p will be seen to be more involved (Section 3). d of p. For m, n ∈ Nd We will say that the multi-degree deg p of a polynomial p is m = (m1, . . . , md) if mi = degi p is the degree of p as a polynomial of zi, i = 1, . . . , d. The total degree tdeg p of p is the largest k over all monomials zk = zk1 1 · . . . · zkd 0, the inequality m ≤ n will be meant in the usual component-wise sense: mi ≤ ni, i = 1, . . . , d. For a matrix K, the principal submatrix determined by an index set α will be denoted by K[α]. Given a collection of complex numbers cα, in- dexed by nonempty subsets α of {1, . . . , d}, the Principal Minor Assignment Problem (see, e.g., [25, 15]) consists of finding a d × d matrix K such that det K[α] = cα for all α. This problem is, in general, overdetermined since the number of independent principal minors grows exponentially with the ma- trix size, d, while the number of free parameters, the matrix entries, is d2. It becomes well-posed under additional assumptions on K or d. For theoretical and computational advances, see [12, 20, 16]. A polynomial of multi-degree (1, . . . , 1), is said to be multi-affine. For such a polynomial, the problem of finding a representation (1.1) with n = (1, . . . , 1) is equivalent to the Principal Minor Assignment Problem. This follows by comparing the expansion det(Id − KZ(1,...,1)) = 1 + Xα6=∅ (−1)card α det K[α]Yi∈α zi, to the general form p(z) = 1 + Xα6=∅ (−1)card αcαYi∈α zi of a d-variable multi-affine polynomial p, p(0) = 1. For a general polynomial p, finding a determinantal representation (1.1) with n = deg p may not be possible by the same dimension count as above. It is clear that (1.1) implies that n ≥ deg p. If n is prescribed, one may view (1.1) as the Principal Minor Relation Problem formulated in Section 2. This paper is largely motivated by our study [13] of the multivariable von Neumann inequality and the discrepancy between the Schur and Schur– Agler norms of analytic functions on the unit polydisk Dd = {z = (z1, . . . , zd) ∈ Cd : zi < 1, i = 1, . . . , d}. Norm-constrained determinantal representations of polynomials 3 Dd such that The Schur class, consisting of analytic L(U,Y)-valued functions f on (1.2) kfk∞ := sup z∈Dd kf (z)k ≤ 1, will be denoted by Sd(U,Y). Here L(U,Y) is the Banach space of bounded linear operators from a Hilbert space U to a Hilbert space Y. The Schur– Agler class, introduced in [1], will be denoted by SAd(U,Y). It consists of analytic L(U,Y)-valued functions on Dd such that T kf (T )k ≤ 1, kfkA := sup (1.3) where the supremum is taken over all d-tuples T = (T1, . . . , Td) of commuting strict contractions on a common Hilbert space. In the scalar case U = Y = C and in the case U = Y, we will use respective shortcuts Sd, SAd, and Sd(U), SAd(U). For a bounded analytic function f : Dd → L(U,Y), the von Neumann inequality is the inequality between its Schur and Schur–Agler norms: kfkA ≤ kfk∞. (1.4) It is valid when d = 1 [28] and d = 2 [2], and not always valid when d ≥ 3 [26, 8, 14]. Thus a Schur function f is Schur–Agler if and only if (1.4) holds. One has the inclusion SAd(U,Y) ⊆ Sd(U,Y). The two classes coincide when d = 1 and d = 2, and the inclusion is proper when d ≥ 3. See, e.g., [13] for details. A d-variable polynomial is said to be stable if it has no zeros in D , and semi-stable if it has no zeros in Dd. A rational function in Sd(CN ) is said to be inner if its radial limits are unitary (unimodular, in the scalar case) almost everywhere on the d-torus. Every scalar-valued rational inner function is necessarily of the form f (z) = zn ¯p(1/z)/p(z) for some n ∈ Nd 0 and a semi-stable polynomial p [24, Theorem 5.5.1], where ¯p(z) := p(¯z). A rational inner function f ∈ Sd is said to have a transfer-function realization (of order m ∈ Nd 0) if there exists a unitary matrix d U = (cid:20)A B C D(cid:21) ∈ C(1+m)×(1+m) so that f (z) = A + BZm(I − DZm)−1C. (1.5) Such a realization for a scalar-valued rational inner f exists if and only if f ∈ SAd [1],[17, Theorem 2.9]. In Section 4, we explore the Schur–Agler class in the context of exterior products, proving, in particular, that if S is a matrix-valued Schur–Agler function, then so are its determinant det S and permanent per S. Following [18], we will say that a semi-stable polynomial p is an Agler denominator if the rational inner function zdeg p ¯p(1/z)/p(z) is Schur–Agler. Extending this notion, we will call a semi-stable polynomial p an eventual Agler denominator of order n ∈ Nd 0 if zn ¯p(1/z)/p(z) is Schur–Agler. 4 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman Representations (1.1) may allow for a fresh approach to the study of the multivariable von Neumann inequality (1.4). In Section 5, we examine the discrepancy between the Schur and Schur–Agler classes via (eventual) Agler denominators. It is shown that (i) not every (semi-)stable polynomial is an Agler denominator, (ii) if q is a polynomial in SAd with q(0) = 0, then 1 − q admits a representation (1.1) for some n ∈ Nd 0 and K a contraction, and (iii) (building on results in Section 4) if p is representable in the form (1.1) for some n ∈ Nd 0 and contractive K, then p is an eventual Agler de- nominator of order n. As a corollary, we deduce that every semi-stable linear polynomial is an Agler denominator, thus solving a problem suggested in [18]. To illustrate a possible advantage of our approach, we compare a minimal determinantal representation (1.1) to a minimal transfer-function realization (1.5) in Remark 5.10. In Section 6, we revisit the Kaijser–Varopoulous–Holbrook example to build a family of polynomials in Sd \ SAd, for every odd d ≥ 3. If d = 3, this leads to a slightly improved bound for the von Neumann constant. 2. Unconstrained determinantal representations Theorem 2.1. Every p ∈ C[z1, . . . , zd], with p(0) = 1, admits a representation (1.1) for some n ∈ Nd 0 and some K ∈ Cn×n. Note that the d-tuple n is not prescribed in the statement, although a bound on n will be deduced in the proof. If n is specified, as in the Princi- pal Minor Assignment Problem, Theorem 2.1 ensures the solvability of the following problem for sone n (see also Remark 3.8). The Principal Minor Relation Problem. Let m ∈ Nd 0, and let pk, 0 ≤ k ≤ m, be a collection of complex numbers. Given n = (n1, . . . , nd) ∈ Nd 0, n ≥ m, find a matrix K ∈ Cn×n whose principal minors K[α1 ∪ ··· ∪ αd], indexed by α1 ⊆ {1, . . . , n1}, α2 ⊆ {n1 + 1, . . . , n1 + n2}, . . . , αd ⊆ {n1 +··· + nd−1 + 1, . . . ,n}, satisfy the relations det K[α1 ∪ ··· ∪ αd] = pk, 0 ≤ k ≤ m. (2.1) (−1)k Xαi=ki, i=1,...,d When m = n = (1, . . . , 1), this is the classical Principal Minor Assignment Problem mentioned in Section 1. The following standard result (see, e.g., [23, Theorem 3.1.1]) will occa- sionally be used. Lemma 2.2. Let P = (cid:20)A B C D(cid:21) be a block matrix with square matrices A and D. If det A 6= 0, then det P = det A det(D−CA−1B). Similarly, if det D 6= 0, then det P = det D det(A − BD−1C). The proof of Theorem 2.1 will be based on the next two lemmas. Norm-constrained determinantal representations of polynomials 5 Lemma 2.3. For every q ∈ Ca×b[z1, . . . , zd], there exist natural numbers s0 = a, s1, . . . , st−1, st = b, matrices Ci ∈ Csi×si+1 , and diagonal si × si matrix functions Li with the diagonal entries in {1, z1, . . . , zd}, such that q(z) = C0L1(z)··· Ct−1Lt(z)Ct. (2.2) The factorization can be chosen so that t = tdeg q. Proof. We apply induction on t. If t = 0, then (2.2) holds trivially with C0 = q(z). Suppose a representation (2.2) exists for every matrix polynomial in z1, . . . , zd of total degree t − 1. Then a polynomial q ∈ Ca×b[z1, . . . , zd] of total degree t can be represented in the form q(z) = q0 + z1q1(z) + ··· + zdqd(z) = (cid:2)q0 Ib 0 q1(z) . . . qd(z)(cid:3)   qd(z)(cid:3) ∈ Ca×(d+1)b[z1, . . . , zd] is a polynomial of total . . . . . . . . . 0 z1Ib . . . . . .   Ib Ib ... Ib 0 ... 0 0 ... 0 zdIb ,     where(cid:2)q0 degree t − 1. By assumption, we have q1(z) . . . (cid:2)q0 q1(z) . . . qd(z)(cid:3) = C0L1(z)··· Ct−2Lt−1(z)Ct−1, which gives q(z) = C0L1(z)··· Ct−1Lt(z)Ct, with Lt = Ib 0 ... 0   0 z1Ib . . . . . . . . . . . . . . . 0 0 ... 0 zdIb   , Ct = Ib Ib ... Ib     . (cid:3) Lemma 2.4. Let Ai ∈ Csi×si+1 , i = 0, . . . , t − 1, and At ∈ Cst×s0 , where s0 = a. Then det   Ia −A0 0 Is1 ... . . . 0 −At 0 0 . . . . . . . . . . . . 0 ... . . . . . . . . . . . . −At−1 0 Ist 0   = det(Ia − A0 ··· At). (2.3) Proof. We apply induction on t ≥ 1. For t = 1, Lemma 2.2 gives det(cid:20) Ia −A1 Is1 (cid:21) = det(Ia − A0A1). −A0 6 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman Suppose (2.3) holds for t − 1 in the place of t. Then, again by Lemma 2.2, det = det   Ia −A0 0 Is1 ... . . . . . . 0 −At 0   Ia −A0 Is1 0 ... . . . ... 0   . . . = det   − 0 . . . . . . . . . . . . −At−2 0 . . . 0 ... 0 Ist−1 −At−1 0  Ist 0 ... 0 . . . . . . . . . . . . −At−2 Ist−1 −A0 Is1 . . .  Ia 0 0 ... 0 0 . . . . . . . . . . . .   −At−1At 0   0 ... 0   (cid:2)−At −At−1  0 . . . 0(cid:3)  0 . . . . . . . . . . . .  0 ... 0 . . . . . . . . . . . . −At−2 0 = det(Ia − A0 ··· At−1At).  Ist−1 Proof of Theorem 2.1. Applying Lemma 2.3 to q = 1 − p, we obtain (here a = b = 1). So, by Lemma 2.4, p(z) = 1 − C0L1(z)··· Ct−1Lt(z)Ct where N = 1 + s1 + ··· + st and p(z) = det(IN − Q(z)), Q(z) = . . . 0 C0L1(z) ... ... 0 Ct . . . 0   0 . . . . . . . . . . . . 0 ... 0 . . . . . . . . . 0 Ct−1Lt(z) . . . 0   . Since Q(z) factors as C · L(z), where   0 C0 ... . . . ... 0 Ct . . . 0  0 . . . . . . . . . . . . . . . 0 ... . . . . . . 0 0 Ct−1 0 . . .  C = , L(z) = 1 0 0 L1(z) ... . . . . . . 0   0 ... 0 . . . . . . . . . 0 Lt(z)   (cid:3) , (2.4) Norm-constrained determinantal representations of polynomials 7 it may be written in the form Q(z) = T GT −1 · T (cid:20)IN−n 0 0 Zn(cid:21) T −1 where G = T −1CT ∈ CN×N and T is a permutation matrix. Representing G as a 2 × 2 block matrix, we obtain that p(z) = det(cid:20)IN−n − G11 In − G22Zn(cid:21) . −G12Zn −G21 Hence det(IN−n − G11) = p(0) = 1 and, in particular, the matrix IN−n − G11 is invertible. Therefore, by Lemma 2.2, p(z) = det(In − G22Zn − G21(IN−n − G11)−1G12Zn) = det(In − KZn) with K = G22 + G21(IN−n − G11)−1G12. (cid:3) 3. Constrained determinantal representations We will now look into the existence of a determinantal representation (1.1) with a norm constraint on the matrix K. First, we give norm-constrained versions of Lemma 2.3 and Theorem 2.1. Lemma 3.1. For every polynomial q ∈ SAd(Cb, Ca), a factorization (2.2) exists with constant contractive matrices Ci, i = 0, . . . , t, where t ≥ tdeg q. Proof. Let q ∈ SAd(Cb, Ca) be a polynomial. By [22, Corollary 18.2], q can be written as a product of constant contractive matrices and diagonal matrices with monomials on the diagonal. Every such diagonal matrix is, in turn, a product of matrices Li as in (2.2) (interlacing with Ci = I). (cid:3) Theorem 3.2. Let p be a polynomial of the form p(z) = det(IN−q(z)), where q is a Schur–Agler polynomial with matrix coefficients, i.e., q ∈ CN×N [z1, . . . , zd] and q ∈ SAd(CN ). If p(0) = 1, then (1.1) holds with K a contraction. Proof. By Lemma 3.1, the matrices Ci in the factorization (2.2) can be chosen contractive. Then the matrix G as in the proof of Theorem 2.1 is also contrac- tive, and by the standard closed-loop mapping argument, K is contractive as well. For reader's convenience, we include this argument. Given u ∈ Cn, the vector equation G21 G22(cid:21)(cid:20)x (cid:20)G11 G12 u(cid:21) = (cid:20)x y(cid:21) in x ∈ CN−n and y ∈ Cn has a unique solution x = (IN−n − G11)−1G12u, y = (G22 + G21(IN−n − G11)−1G12)u = Ku. Since G is a contraction, we have kxk2 + kyk2 ≤ kxk2 + kuk2, i.e., kyk ≤ kuk. Since u ∈ Cn is arbitrary, K is a contraction as claimed. (cid:3) Corollary 3.3. Let p ∈ C[z1, . . . , zd], with p(0) = 1, be such that q = 1 − p ∈ SAd, then (1.1) holds with K a contraction. 8 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman Remark 3.4. The converse to Corollary 3.3 is false. Indeed, if d = 1 and √2 − 1 < a ≤ 1, then p(z) = (1 − az)2 satisfies (1.1) with K = aI2, obviously a contraction. However, k1 − pkA = k1 − pk∞ = 2a + a2 > 1. Define the stability radius s(p) of a d-variable polynomial p to be s(p) = sup(cid:26)r > 0 : p(z) 6= 0, z ∈ rDd(cid:27). Clearly, p is semi-stable if s(p) ≥ 1, and stable if s(p) > 1. It is easy to see that kKk ≥ 1/s(p) whenever p admits (1.1). Remark 3.5. With respect to a given subalgebra ∆ ⊆ Cd×d, the structured singular value µ∆(K) of a matrix K ∈ Cd×d is defined to be µ∆(K) := (cid:0)inf(cid:8)kZk : Z ∈ ∆ and det(I − KZ) = 0(cid:9)(cid:1)−1 . The theory of structured singular values was introduced in [9] to analyze linear systems with structured uncertainties; for an overview, see for instance i=1 ziIni : z ∈ Cd}, [29, Chapter 10]. If p satisfies (1.1) and ∆ = {Zn = Ld we recognize that µ∆(K) = 1/s(p). The next theorem gives a way of constructing a representation (1.1) with a certain upper bound on the norm of K. Theorem 3.6. Given a polynomial p(z) = 1 +Pk∈S pkzk, where S ⊆ Nd 0 \{0} and the coefficients pk, k ∈ S, are nonzero, let t = tdeg p, n = Pk∈S k, and β = (cid:0)Pk ∈ S pk(cid:1) 1 t+1 . Then p admits a representation (1.1) with K ∈ Cn×n, and kKk ≤ β maxnp(β2 − 1)(1 + β + ··· + βκ−1)2 + 1, 1o (3.1) for some integer κ, 1 ≤ κ ≤ t. Remark 3.7. If β ≤ 1 and s(p) = 1/β, which is the case for semi-stable linear polynomials, the norm bound asserted in the theorem is sharp. In general, it is not sharp, even in the univariate case. Remark 3.8. Theorem 3.6 implies that the Principal Minor Relation Problem Proof of Theorem 3.6. Form the matrices (2.1) with data {pk 6= 0 : k ∈ S} is solvable for n = Pk∈S k. 2(cid:21), C1 = . . . = Ct−1 = βIS, Ct = β C0 = −β all of equal norm β. Relative to the standard ordering of factors, k∈S(cid:20)pk row 1−t 1 1−t 2 2 col 2 (cid:21), k∈S(cid:20) pk pk 1 write each monomial zk, k ∈ S, as an expanded product zk = zi1(k) ··· zit(k), where zij (k) 6= 1, for 1 ≤ j ≤ k, and zij (k) = 1, for k + 1 ≤ j ≤ t. Let zk = z1 ··· z1 {z } k1 times · z2 ··· z2 {z } k2 times , · . . . · zd ··· zd {z } kd times Lj(z) = diag k∈S [zij (k)], j = 1, . . . , t, Norm-constrained determinantal representations of polynomials 9 and observe that L1(z) contains no unit entries by construction. It is then easy to check that (2.2) holds for q = 1 − p. Thus, by Lemma 2.4, we obtain p(z) = det  I1+St −  = det(I1+St − C · L(z)), . . . 0 C0L1(z) ... ... 0 Ct . . . 0   0 . . . . . . . . . . . . 0 ... 0 . . . . . . . . . 0 Ct−1Lt(z) . . . 0     where C and L(z) are as in (2.4). Using an appropriate permutation T , we can bubble-sort L(z) so that all diagonal ones are stacked in the left upper corner block: T L(z)T −1 = (cid:20)Iℓ 0 Zn(cid:21) , 0 where ℓ = 1 + St − n. Then, partitioned accordingly, T CT −1 = (cid:20)G11 G12 G21 G22(cid:21) =: G has the following structure: G =   0 . . . 0 ... ... 0 . . . 0 ∗ ∗ . . . 0 ... ... 0 . . . 0 ∗ 0 ∗ . . . . . . . . . 0 . . . . . . . . . . . . . . . 0 ... . . . . . . 0 0 ∗ . . . 0 . . . 0 ... . . . . . . 0 0 ∗ . . . 0 ∗ 0 ... ... 0 0 ... ... ... 0 0 0 . . . . . . ∗ . . . 0 ∗ . . . . . . . . . 0 . . . . . . . . . . . . . . . 0 ... . . . . . . 0 . . . 0 ∗ 0 . . . 0 ... . . . . . . 0 . . . . . . ∗ 0   . We observe that G11 is nilpotent of index κ, where κ is the number of nonzero blocks Li(0) (counting L0 ≡ 1). Since, L1(0) = 0, we necessarily have 1 ≤ κ ≤ t. Thus, we obtain that 0 Zn(cid:21)(cid:17) = det(In−KZn), p(z) = det(I1+St−C·L(z)) = det(cid:16)I1+St−G·(cid:20)Iℓ where K = G22 + G21(I − G11)−1G12 by the same argument as in the proof of Theorem 3.2. 0 10 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman The norm bound on K is obtained as follows. For a fixed u ∈ Cn, the solution to the vector equation G21 G22(cid:21)(cid:20)x (cid:20)G11 G12 u(cid:21) = (cid:20)x y(cid:21) in x ∈ Cℓ and y ∈ Cn is given by Observing that kGk = β, we have x = (Iℓ − G11)−1G12u = (Iℓ + G11 + ··· + Gκ−1 11 )G12u, y = Ku. If β ≤ 1, then kyk ≤ βkuk and thus kKk ≤ β. If β > 1, then the estimate kxk2 + kyk2 ≤ β2(kxk2 + kuk2). kxk ≤ (1 + β + ··· + βκ−1)βkuk implies that kyk2 ≤ (β2−1)kxk2+β2kuk2 ≤ β2(cid:0)(β2 − 1)(1 + β + ··· + βκ−1)2 + 1(cid:1)kuk2, which yields (3.1). Remark 3.9. Given a d-variable polynomial p, p(0) = 1, and a d-tuple n ≥ deg p such that (1.1) holds, one may consider the set Kn(p) of n×n matrices K such that det(In − KZn) = p(z). It is then of interest to determine the constant (cid:3) α(p) := inf n min K∈Kn(p)kKk. In particular, it is unclear whether α(p) < 1 (α(p) ≤ 1) for p stable (semi- stable). 4. The Schur–Agler class and wedge powers We will now examine the Schur–Agler norm of tensor and exterior products of operator-valued functions. The results are preceded by some definitions. For a background on tensor and exterior algebras see, e.g., [4, 10, 11]. Let V⊗k be the k-fold tensor power of a vector space V. The k-th an- tisymmetric tensor power V∧k of V may be viewed as a subspace of V⊗k, generated by elementary antisymmetric tensors v1 ∧ . . . ∧ vk = Xσ (sign σ)vσ(1) ⊗ . . . ⊗ vσ(k), where the summation is taken over all permutations σ of 1, 2, . . . , k. Given a linear map A : U → V of vector spaces, the linear operator A∧k(u1 ∧ . . . ∧ uk) = Au1 ∧ . . . ∧ Auk, A∧k : U∧k → V∧k, determined by the equalities is the compression πV ∧k A⊗k(cid:12)(cid:12)U ∧k of the tensor power A⊗k : U⊗k → V⊗k. Here If e1, . . . , en form a basis for V, then ei1∧. . .∧eik , 1 ≤ i1 < . . . < ik ≤ n, form a basis for V∧k of cardinality(cid:0)n k(cid:1). Relative to a choice of bases for U and πM denotes the orthogonal projection onto a subspace M . Norm-constrained determinantal representations of polynomials 11 V, the matrix entry for A∧k in row-column position ((i1, . . . , ik), (j1, . . . , jk)) is the minor of the matrix of A built from rows i1, . . . , ik and columns j1, . . . , jk. If U and V are normed vector spaces and if S(z) = Pr∈Nd Srzr is a power series with coefficients in L(U,V), then, for any tuple T = (T1, . . . , Td) of commuting operators on some normed vector space H, we may consider the operator 0 S(T ) = Xr∈Nd 0 Sr ⊗ T r 0 . 0 k   and its compression acting from U ⊗ H to V ⊗ H, provided the series converges. More gener- ally, starting with power series Sj(z) = Pr∈Nd (Sj)rzr, j = 1, . . . , k, with coefficients in L(U,V), and T = (T1, . . . , Td) as above, form the operator Oj=1 (S1 ⊗ ··· ⊗ Sk) (T ) = Xr1,...,rk∈Nd  ⊗ T r1+...+rk , (S1 ∧ ··· ∧ Sk) (T ) = (πV ∧k ⊗ IH)(S1 ⊗ ··· ⊗ Sk)(T )(cid:12)(cid:12)U ∧k⊗H (Sj)rj The objective of this section is to establish the following theorem. Theorem 4.1. Let U and V be Hilbert spaces and let S1, . . . , Sk belong to SAd(U,V). Then S1 ⊗ ··· ⊗ Sk belongs to SAd(U⊗k,V⊗k) and S1 ∧ ··· ∧ Sk belongs to SAd(U∧k,V∧k). Proof. Let T = (T1, . . . , Td) be a tuple of commuting strict contractions on some Hilbert space H. Then the mapping (Sj)rj     IV ⊗ ··· ⊗ IV ⊗ (Sj)rj ⊗ IV ⊗ ··· ⊗ IV ⊗ T rj Sj  (T ) = Xr1,...,rk∈Nd Yj=1 Xrj∈Nd (cid:18)IV ⊗ ··· ⊗ IV ⊗ Sj ⊗ IV ⊗ ··· ⊗ IV(cid:19)(T ) Yj=1  ⊗ T r1+...+rk Oj=1 Oj=1 k k = = 0 0 k k is contractive as a product of contractive factors. Hence Nk SAd(U⊗k,V⊗k). Consequently, kS1 ∧ . . . ∧ SkkA ≤ kS1 ⊗ . . . ⊗ SkkA ≤ 1, j=1 Sj belongs to which gives the second assertion. Corollary 4.2. Let S be a n × n matrix-valued Schur–Agler function, i.e., S ∈ SAd(Cn). Then, for every k = 1, . . . , n, the k-th compound matrix-valued function of S is also Schur–Agler. In particular, det S(z) is a Schur–Agler function. (cid:3) 12 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman Proof. The matrix of S∧k is the k-th compound matrix of S. The case k = n corresponds to det S(z). (cid:3) Similarly, in the setting of k-th symmetric tensor powers, one may con- sider the operators (S1 ∨ ··· ∨ Sk)(T ). The proof of the following theorem is omitted as it parallels the preceding development. Theorem 4.3. Let U and V be Hilbert spaces and let S1, . . . , Sk belong to SAd(U,V). Then S1 ∨ ··· ∨ Sk belongs to SAd(U∨k,V∨k). Corollary 4.4. Let S be a n × n matrix-valued Schur–Agler function, i.e., S ∈ SAd(Cn). Then, for every k = 1, . . . , n, the k-th permanental compound matrix-valued function of S is also Schur–Agler. In particular, the permanent of a Schur–Agler function is also Schur–Agler. We note that a permanental analog of (1.1) features in [7]. 5. Agler denominators We are in a position to discuss (eventual) Agler denominators and stability in relation to (1.1). It will first be shown that there exist stable polynomials in three or more variables that are not Agler denominators. Example 5.1. Let p(z) be a d-variable polynomial, with kpk∞ = 1 and multi- degree m, violating the von Neumann inequality (1.4). Let there exist a tuple T = (T1, . . . , Td) of commuting contractions such that T k = T k1 d = 0, for some k ∈ Nd 0, and kp(T )k > 1. Examples of such a scenario can be found in [26, 14, 8]; see also Section 6. For 0 < r < 1, the polynomial q(z) = 1 + rzk+m ¯p(1/z) is stable and so 2 ··· T kd 1 T k2 the rational function f (z) = zk+m + rp(z) 1 + rzk+m ¯p(1/z) is inner. However, since f (T ) = rp(T ), f does not belong to SAd whenever r > 1/kp(T )k. In particular, if the multi-degree of zm ¯p(1/z) is also m, then f (z) = zk+m ¯q(1/z)/q(z), so that q is not an Agler denominator. To give a concrete example, we specialize to the Kaijser–Varopoulos– Holbrook setting. The polynomial p(z1, z2, z3) = 1 + z2 2 + z2 1 5(cid:18)z2 3 − 2z1z2 − 2z2z3 − 2z3z1(cid:19) satisfies kpk∞ = 1, and there exist commuting contractions T1, T2, T3 such that kp(T1, T2, T3)k = 6/5 and T1T2T3 = 0. The corresponding rational inner function f (z1, z2, z3) = z3 1z3 2z3 5 z1z2z3(z2 3 + r 1z2 5 (z2 2 + z2 1 + z2 2z2 2 + z2 3 − 2z1z2 − 2z2z3 − 2z3z1) 2z3 − 2z2 1 − 2z1z2z2 3z2 3 − 2z1z2 3 + z2 1 + r , 1z2z3) Norm-constrained determinantal representations of polynomials 13 is not Schur–Agler for 5/6 < r < 1. For these values of r, the stable polyno- mial q(z1, z2, z3) = 1+ r 5 z1z2z3(cid:18)z2 1z2 2 +z2 2z2 3 +z2 3z2 1 −2z1z2z2 3 −2z1z2 2z3−2z2 1z2z3(cid:19) is not an Agler denominator. We now have the following result. Theorem 5.2. Let a polynomial p admit a representation (1.1) for some n ∈ Nd 0 and contractive K. Then zn ¯p(1/z) p(z) = det(−K∗ + √I − K∗KZn(I − KZn)−1√I − KK∗). (5.1) In particular, p is an eventual Agler denominator of order n. If deg p = n, then p is an Agler denominator. A lemma is needed; see, e.g., [23, Theorem 3.1.2]. Lemma 5.3. Let A, B, C, and D be square matrices of the same size, and suppose that AC = CA. Then det(cid:20)A B C D(cid:21) = det(AD − CB). Proof of Theorem 5.2. By Lemma 2.2, the right hand side of (5.1) equals det(cid:20) −K∗ √I − KK∗ −√I − K∗KZn I − KZn (cid:21) . (5.2) det(I − KZn) Let K = U ΣV ∗ be a singular value decomposition of K. Then the numerator of (5.2) equals det(cid:20)V U∗ 0 0 I(cid:21) det(cid:20) −U ΣU∗ U√I − Σ2U∗ −U√I − Σ2V ∗Zn I − KZn (cid:21) . get that (5.3) equals Applying Lemma 5.3, noting that −U ΣU∗ and U√I − Σ2U∗ commute, we det(V U∗) det(−U ΣU∗(I − U ΣV ∗Zn) + UpI − Σ2U∗UpI − Σ2V ∗Zn) (5.3) = det(Zn − K∗). (5.4) (5.5) To prove (5.1) it remains to observe that znp(1/z) = zn det(I − KZ−1 n ) = det(I − KZ−1 n ) det Zn where in the last step we used that Z⊤n = Zn. As the Julia operator = det(Zn − K) = det(Zn − K∗), (cid:21) (cid:20) −K∗ √I − KK∗ √I − K∗K K is unitary, the multivariable rational inner matrix function −K∗ + √I − K∗KZn(I − KZn)−1√I − KK∗ 14 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman is in the Schur–Agler class. By Corollary 4.2, so is its determinant, and thus zn ¯p(1/z)/p(z) is in the Schur–Agler class. (cid:3) Corollary 5.4. For every p ∈ C[z1, . . . , zd] with p(0) = 1, there exists r > 0 such that the polynomial pr(z) := p(rz) is an eventual Agler denominator. In fact, if p is given by (1.1), then one can choose any 0 < r ≤ 1/kKk. Proof. Since, by Theorem 2.1, every polynomial p with p(0) = 1 admits a representation (1.1), the assertion follows from the identity and Theorem 5.2. pr(z) = det(In − rKZn), (cid:3) Remark 5.5. The first statement of Corollary 5.4 can also be deduced from Corollary 3.3: since k1 − p(0)kA = 0, the inequality k1 − prkA ≤ 1 holds, by continuity, for a sufficiently small r > 0. For multi-affine symmetric polynomi- als a stronger statement is true [18, Theorem 1.5]: pr is an Agler denominator for sufficiently small r > 0. Following [3], we call a semi-stable polynomial p scattering Schur if p and zdeg p ¯p(1/z) have no factor in common. In [19, Theorem 1] it was proven that every two-variable scattering Schur polynomial p of degree n = (n1, n2) is of the form (1.1) with K an (n1 + n2) × (n1 + n2) contraction. Thus every two-variable scattering Schur polynomial p is an Agler denominator. The following result provides a partial converse to Theorem 5.2. Theorem 5.6. Let p be a d-variable scattering Schur polynomial with p(0) = 0, the rational inner function zm ¯p(1/z)/p(z) has a 1. If, for some m ∈ Nd transfer-function realization (1.5) of order m, then p admits a representation (1.1) with n = m and K a contraction. Proof. Taking the determinant of both sides of the equality zm ¯p(1/z) p(z) = A + BZm(I − DZm)−1C and using Lemma 2.2, we obtain zm ¯p(1/z) p(z) = C I − DZm(cid:21) det(cid:20)A −BZm det(I − DZm) =: r(z) s(z) . Note that both r(z) and s(z) are of degree at most m. We now obtain that (zm ¯p(1/z))s(z) = r(z)p(z). As p is scattering Schur we must have that p(z) divides s(z), say s(z) = q(z)p(z). Dividing out p(z) in the above equation, we obtain that (zm ¯p(1/z))q(z) = r(z). As the left hand side has degree m + deg q and the right hand side degree at most m, we obtain that q must be a constant. But then, using p(0) = 1 and s(0) = det(I − DZm)z=0 = 1, we obtain that q = 1, and thus p(z) = det(I − KZm) with K = D. (cid:3) Norm-constrained determinantal representations of polynomials 15 1, are Agler denominators. Corollary 5.7. The polynomials p(z1, . . . , zd) = 1−Pd Proof. Let K be a d × d rank 1 contraction with diagonal entries a1, . . . , ad. One such choice is given by i=1 aizi with Pd i=1 ai ≤ K = (cid:20)qajakei argak(cid:21)d j,k=1 . Then det(Id−KZ(1,...,1)) = p(z) and the result follows directly from Theorem 5.2. (cid:3) Remark 5.8. The matrix K in the proof of Corollary 5.7 is clearly of minimal size. It is also of minimal norm, kKk = a1 + ··· + ad, for otherwise p(z) = det(Id − KZ(1,...,1)) would be stable. It was shown in [18, Theorem 3.3] that a multi-affine symmetric poly- nomial is an Agler denominator if and only if a certain matrix B constructed from the Christoffel–Darboux equation is positive semidefinite. Subsequently, for p(z) = 1 − 1 i=1 zi, the positivity of the matrix B was computationally checked up to d = 11. Using our Corollary 5.7, we deduce this fact for all d. Corollary 5.9. Let p(z) = t − 1 i=1 zi, where t ≥ 1, and let d Pd d Pd B := (Bα∩β α,β )α,β⊆{1,...,d−1}, via: be the 2(d−1)×(d−1) matrix indexed by subsets α, β of {1, . . . , d − 1}, defined j(cid:19)−1(cid:18)d (cid:18)d where 0 ≤ i ≤ j, k ≤ d − 1, p0 = t, p1 = 1 i, j, k not satisfying 0 ≤ i ≤ j, k ≤ d − 1. Then B is positive semidefinite. j,k − iBi−1 d , pj = 0, j ≥ 2, and Bi (pj ¯pk − ¯pd−jpd−k) = (d − j − k + i)Bi k(cid:19)−1 j−1,k−1, j,k = 0 for (5.6) 1 18 B = −3t 3t2 + 1 2 −3t To illustrate, we choose t real and d = 3:  −3t 0 −3t 2  3t2 + 1 −3t 6t2 −3t  6t2 −3t  −3t 0 √6 √6t  √6 0  0 p3(t2 − 1) p3(t2 − 1) where A = 1 √18 − 1 − 1 −1 − 1 − 1 1 √6 √6 2 2 2 2 0 = A∗A, 0 √6t 0 0   . Remark 5.10. In the context of (1.1), the question of whether a given polyno- mial is an (eventual) Agler denominator is reduced to analyzing the matrix of its determinantal representation. This provides a possible alternative to the 16 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman transfer-function realization method. For example, p(z) = 1− 1 admits a representation (1.1) with 3 (z1 + z2 + z3) K = 1 3 1 1 1   1 1 1 1 1 1   , which is minimal both in size and in norm; see Remark 5.8. At the same time, the minimal order of a transfer-function realization (1.5) of 3z1z2z3 − z2z3 − z1z3 − z1z2 ¯p(1/z) z1z2z3 = p(z) 3 − z1 − z2 − z3 is m = (2, 2, 2) [17, 5]. 6. Variations on the Kaijser–Varopoulous–Holbrook example For s real, consider the multivariable polynomial p(z1, . . . , zd) = (1 + s) d Xm=1 z2 m −(cid:18) d Xm=1 zm(cid:19)2 . The case of d = 3 and s = 1 corresponds to the polynomial from [26, 14]. A = Proposition 6.1. Let d > 1, s > d/2 − 1, and p be defined as above. Then kpk∞ = (1 + s)d, if d is even, and kpk∞ < (1 + s)d, if d is odd, while kpkA = (1 + s)d for all d. In particular, p/kpk∞ is not in SAd for odd d > 1. Proof. Write p(z1, . . . , zd) = z⊤Az, where . . . −1 . . . −1 . . . . . . z1 ... zd and z =    s −1 s −1  kAk = max(cid:8)1 + s, s − d + 1(cid:9) = 1 + s. −1 −1 Observe that A is symmetric with eigenvalues 1 + s and s − d + 1, so Hence kpk∞ ≤ (1 + s)d. If d is even, one immediately has equality since p(1,−1, . . . , 1,−1) = (1 + s)d. If d is odd, the inequality is strict. Indeed, otherwise p would be maximized for some unimodular z1, . . . , zd with zero sum, as z would then lie in the eigenspace of A corresponding to 1 + s. But then the equality     s . d p(z1, . . . , zd) = (1 + s)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xi=1 z2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (1 + s)d would force zi = ±eiα, i = 1, . . . , d, in conflict with the zero sum condition. Norm-constrained determinantal representations of polynomials 17 Next, for a tuple of commuting contractions T = (T1, . . . , Td), we have p(T1, . . . , Td) = (cid:2)T1 . . . Td(cid:3) (A ⊗ I)  kp(T1, . . . , Td)k ≤ kAkd = (1 + s)d. T1 ... Td ,   Ti =   0 v⊤i 0 0 0 0 0 vi 0   ∈ R4×4, i = 1, . . . , d, so that Choose v1, . . . , vd to be any unit vectors in R2 with zero sum. Then the matrices i=1 Ti = 0, and p(T1, . . . , Td)e4 = (1 + s)de1, are such that kTik = 1, TiTj = TjTi = hvi, vjie1e⊤4 , Pd where ej is the jth standard unit vector in R4. Hence kpkA = (1 + s)d. (cid:3) Remark 6.2. In the case of d = 3, maximizing kpkA/kpk∞, the von Neu- mann constant of p, over s, we find that the maximum possible ratio is 3q 35+13√13 1 ≈ 1.23 (occuring for s = 6 bound for the von Neumann constant was 6 ). The previously known lower √13+1 6 5 [14]. Acknowledgment We thank Victor Vinnikov and Bernd Sturmfels for energizing discussions, Greg Knese and the anonymous referee for valuable suggestions, and David Scheinker for bringing [6, 7] to our attention. References [1] J. Agler. On the representation of certain holomorphic functions defined on a polydisc. In Topics in operator theory: Ernst D. Hellinger memorial volume, volume 48 of Oper. Theory Adv. Appl., pages 47–66. Birkhauser, Basel, 1990. [2] T. Ando. On a pair of commutative contractions. Acta Sci. Math. (Szeged), 24:88–90, 1963. [3] S. Basu and A. Fettweis. New results on stable multidimensional polynomials. II. Discrete case. IEEE Trans. Circuits and Systems 34:1264–1274, 1987. [4] R. Bhatia. Matrix analysis. Graduate texts in Mathematics, 169. Springer- Verlag, New York, 1997. [5] K. Bickel and G. Knese. Inner functions on the bidisk and associated Hilbert spaces. arXiv: 1207.2486. [6] J. Borcea, P. Brand´en, and T. M. Liggett. Negative dependence and the ge- ometry of polynomials. J. Amer. Math. Soc. 22 (2009), no. 2, 521–567. [7] P. Brand´en, J. Haglund, M. Visontai, and D. G. Wagner. Proof of the monotone column permanent conjecture. arXiv:1010.2565v2 [8] M. J. Crabb and A. M. Davie. Von Neumann's inequality for Hilbert space operators. Bull. London Math. Soc., 7:49–50, 1975. 18 Grinshpan, Kaliuzhnyi-Verbovetskyi and Woerdeman [9] J. C. Doyle. Analysis of feedback systems with structured uncertainties. Proc. IEE-D 129 (1982), no. 6, 242–250 [10] H. Flanders. Tensor and exterior powers. J. Algebra, 7:1–24, 1967. [11] W. H. Greub. Multilinear algebra. Die Grundlehren der mathematischen Wis- senschaften, Band 136 Springer-Verlag New York, Inc., New York 1967 x+225 pp. [12] K. Griffin and M. J. Tsatsomeros. Principal minors. II. The principal minor assignment problem. Linear Algebra Appl. 419 (2006), no. 1, 125–171. [13] A. Grinshpan, D. S. Kaliuzhnyi-Verbovetskyi, V. Vinnikov, and H. J. Woerde- man. Classes of tuples of commuting contractions satisfying the multivariable von Neumann inequality. J. Funct. Anal. 256 (2009), no. 9, 3035-3054. [14] J. A. Holbrook. Schur norms and the multivariate von Neumann inequality. In Recent advances in operator theory and related topics (Szeged, 1999), volume 127 of Oper. Theory Adv. Appl., pages 375–386. Birkhauser, Basel, 2001. [15] O. Holtz and H. Schneider. Open problems on GKK τ -matrices. Linear Algebra Appl. 345 (2002), 263-267. [16] O. Holtz and B. Sturmfels. Hyperdeterminantal relations among symmetric principal minors. J. Algebra 316 (2007), no. 2, 634-648 [17] G. Knese. Rational inner functions in the Schur–Agler class of the polydisk. Publ. Mat. 55:343–357, 2011. [18] G. Knese. Stable symmetric polynomials and the Schur-Agler class. Preprint. [19] A. Kummert. 2-D stable polynomials with parameter-dependent coefficients: generalizations and new results. IEEE Trans. Circuits Systems I: Fund. Theory Appl. 49:725–731, 2002. [20] S. Lin and B. Sturmfels. Polynomial relations among principal minors of a 4 × 4-matrix. J. Algebra 322 (2009), no. 11, 4121–4131. [21] T. Netzer and A. Thom. Polynomials with and without determinantal repre- sentations. arXiv:1008.1931. [22] V. Paulsen. Completely bounded maps and operator algebras. Cambridge Stud- ies in Advanced Mathematics, 78. Cambridge University Press, Cambridge, 2002. [23] V. V. Prasolov. Problems and theorems in linear algebra. Translated from the Russian manuscript by D. A. Leites. Translations of Mathematical Mono- graphs, 134. American Mathematical Society, Providence, RI, 1994. [24] W. Rudin. Function theory in the polydisk. W. A. Benjamin, New York, 1969. [25] E. B. Stouffer. On the independence of principal minors of determinants. Trans. Amer. Math. Soc. 26 (1924), no. 3, 356–368 [26] N. Th. Varopoulos. On an inequality of von Neumann and an application of the metric theory of tensor products to operators theory. J. Functional Analysis, 16:83–100, 1974. [27] V. Vinnikov. LMI representations of convex semialgebraic sets and determi- nantal representations of algebraic hypersurfaces: past, present, and future. in "Mathematical Methods in Systems, Optimization, and Control: Festschrift in Honor of J. William Helton" (Eds.Harry Dym, Mauricio C. de Oliveira, Mihai Putinar), Operator Theory: Advances and Applications, Birkhauser, to appear. Norm-constrained determinantal representations of polynomials 19 [28] J. von Neumann. Eine Spektraltheorie fur allgemeine Operatoren eines unitaren Raumes. Math. Nachr., 4:258–281, 1951. [29] K. Zhou and J. C. Doyle. Essentials of robust control, Prentice Hall, 1997, 411 pp. Anatolii Grinshpan Dmitry S. Kaliuzhnyi-Verbovetskyi Hugo J. Woerdeman Department of Mathematics Drexel University 3141 Chestnut St. Philadelphia, PA, 19104 e-mail: {tolya,dmitryk,hugo}@math.drexel.edu
1802.03616
1
1802
2018-02-10T16:33:05
Disjointness of continuous g-frames and Riesz-type continuous g-frames
[ "math.FA" ]
In this paper we introduce concepts of disjoint, strongly disjoint and weakly disjoint continuous $g$-frames in Hilbert spaces and we get some equivalent conditions to these notions. We also construct a continuous g-frame by disjoint continuous g-frames. Furthermore, we provide some results related to the Riesz-type continuous $g$-frames.
math.FA
math
DISJOINTNESS OF CONTINUOUS G-FRAMES AND RIESZ-TYPE CONTINUOUS G-FRAMES Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Abstract. In this paper we introduce concepts of disjoint, strongly disjoint and weakly disjoint continuous g-frames in Hilbert spaces and we get some equivalent conditions to these notions. We also construct a continuous g-frame by disjoint continuous g-frames. Furthermore, we provide some results related to the Riesz-type continuous g-frames. 1. Introduction In 1952, the concept of frames for Hilbert spaces was defined by Duffin and Schaeffer [6]. Frames are important tools in the signal pro- cessing, image processing, data compression, etc. Let H be a separable Hilbert space. We call a sequence F = {fi}i∈I ⊆ H a frame for H if there exist two constant AF , BF > 0 such that hf, fii2 ≤ BFkfk2, f ∈ H. (1.1) AFkfk2 ≤Xi∈I If in (1.1), AF = BF = 1 we say that F = {fi}i∈I is a Parseval frame for H. Let F = {fi}i∈I be a frame for H, then the operator TF : l2(I) → H, TF ({ci}i∈I) =Xi∈I cifi, is well define and onto, also its adjoint is T ∗ F : H → l2(I), T ∗ F f = {hf, fii}i∈I. The operators TF and T ∗ F are called the synthesis and analysis operators of frame F. The concepts of disjoint frames and strongly disjoint frames introduced by Han and Larson [10]. Definition 1.1. Let F = {fi}i∈I and G = {gi}i∈I be frames for Hilbert spaces H and K, respectively. We say that F and G are (i) Disjoint, if {fi ⊕ gi}i∈I is a frame for H ⊕ K. MSC(2010): Primary 41A58, 42C15. Keywords: continuous frame, continuous g-frame, Riesz-type continuous g-frame. ∗Corresponding author . 1 2 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ (ii) Strongly disjoint, if there are invertible operator L1 ∈ B(H) and L2 ∈ B(K) such that {L1fi}i∈I ,{L2gi}i∈I and {L1fi ⊕ L2gi}i∈I are respective Parseval frames for H,K and H ⊕ K. Proposition 1.2. [10] Let F = {fi}i∈I and G = {gi}i∈I be frames for Hilbert spaces H and K, respectively. Then G = {0} F and RangeT ∗ G (ii) F and G are strongly disjoint if and only if RangeT ∗ (i) F and G are disjoint if and only if RangeT ∗ G is closed subspace of l2(I). F ∩ RangeT ∗ and RangeT ∗ F + RangeT ∗ are orthogonal. In 1993, Ali, Antoine and Gazeau developed the notion of ordinary frame to a family indexed by a measurable space which are known as continuous frames [4]. Definition 1.3. Let H be a complex Hilbert space and (Ω, µ) be a measure space. The mapping F : Ω → H is called a continuous frame if (i) F is weakly-measurable, i.e., for all f ∈ H, ω → hf, F (ω)i is a (ii) there exist constants AF , BF > 0 such that measurable function on Ω, AFkfk2 ≤ZΩ hf, F (ω)i2 dµ(ω) ≤ BFkfk2, hSF f, gi =ZΩhf, F (ω)ihF (ω), gidµ(ω), f ∈ H. f, g ∈ H, If F : Ω → H is a continuous frame then the operator SF : H → H defined by is positive and invertible. SF is called the continuous frame operator of F. In 2006, g-frames or generalized frames introduced by Sun [11]. Ab- dollahpour and Faroughi introduced and investigated continuous g- frames and Riesz-type continuous g-frames [2]. Disjointness notions were developed to continuous frames by Gabardo and Han [7] and to g-frames by Abdollahpour [1]. In the rest of this paper we assume that H and K are complex Hilbert spaces and (Ω, µ) is a measure space with positive measure µ and {Kω : ω ∈ Ω} is a family of Hilbert spaces. Now, we summarize some facts about continuous g-frames from [2]. We say that F ∈Qω∈Ω Kω is strongly measurable if F as a mapping of Ω toLω∈Ω Kω is measurable, where Yω∈Ω Kω =(f : Ω → [ω∈Ω Kω : f (ω) ∈ Kω) . DISJOINTNESS OF CONTINUOUS G-FRAMES 3 Definition 1.4. We say that Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame for H with respect to {Kω : ω ∈ Ω} if (i) for each f ∈ H, {Λωf : ω ∈ Ω} is strongly measurable, (ii) there are two constants 0 < AΛ ≤ BΛ < ∞ such that AΛkfk2 ≤ZΩ kΛωfk2dµ(ω) ≤ BΛkfk2, f ∈ H. (1.2) We call AΛ, BΛ the lower and upper continuous g-frame bounds, re- spectively. Λ is called a tight continuous g-frame if AΛ = BΛ, and a Parseval continuous g-frame if AΛ = BΛ = 1. If for each ω ∈ Ω, K = Kω, then Λ is called a continuous g-frame with respect to K. Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} is called a continuous g-Bessel family if the right hand inequality in (1.2) holds for all f ∈ H. In this case, BΛ is called the Bessel constant. If there is no confusion, we use continuous g-frame (continuous g- Bessel family) instead of continuous g-frame for H with respect to {Kω : ω ∈ Ω} (continuous g-Bessel family for H with respect to {Kω : ω ∈ Ω}). Proposition 1.5. [2] Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a continu- ous g-frame. Then there exists a unique positive and invertible operator SΛ : H → H such that hSΛf, gi =ZΩhf, Λ∗ ωΛωgidµ(ω), f, g ∈ H, and AΛI ≤ SΛ ≤ BΛI. The operator SΛ in Proposition 1.5 is called the continuous g-frame operator of Λ. Also, we have (1.3) hf, gi =ZΩhS−1 Λ f, Λ∗ ωΛωgi dµ(ω) =ZΩhf, Λ∗ ωΛωS−1 Λ gi dµ(ω), for all f, g ∈ H. We consider the space bK =(F ∈Yω∈Ω Kω : F is strongly measurable, ZΩ kF (ω)k2dµ(ω) < ∞) . It is clear that bK is a Hilbert space with point wise operations and with the inner product given by hF, Gi =ZΩhF (ω), G(ω)idµ(ω), F, G ∈ bK. 4 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ (1.4) Proposition 1.6. [2] Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a contin- hTΛF, gi =ZΩhΛ∗ uous g-Bessel family. Then the mapping TΛ : bK → H defined by is linear and bounded with kTΛk ≤ √BΛ. Also, for each g ∈ H and ω ∈ Ω, ωF (ω), gidµ(ω), F ∈ bK, g ∈ H, (T ∗ Λg)(ω) = Λωg. Theorem 1.7. [2] Let (Ω, µ) be a measure space, where µ is σ-finite. Suppose that Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} is a family of operators such {Λωf : ω ∈ Ω} is strongly measurable, for each f ∈ H. Then Λ is by (1.4) is bounded and onto. a continuous g-frame if and only if the operator TΛ : bK → H defined Λ in Theorem 1.7 are called the synthesis and The operators TΛ and T ∗ analysis operators of Λ, respectively. Definition 1.8. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two continuous g-frames such that hf, gi =ZΩhf, Θ∗ ωΛωgidµ(ω), f, g ∈ H, then Θ is called a dual continuous g-frame of Λ. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a continuous g-frame. Then eΛ = {ΛωS−1 Λ ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame and by (1.3), eΛ is a dual of Λ and we call eΛ the canonical dual of Λ. One can always get a tight continuous g-frame from any continuous g-frame, in fact, if Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame then {ΛωS−1/2 ∈ B(H,Kω) : ω ∈ Ω} is a Parseval continuous g-frame. Two continuous g-Bessel families Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} are weakly equal, if for all f ∈ H, Λ Λωf = Θωf, a.e. ω ∈ Ω. If the continuous g-frame Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} have only one dual (weakly), i.e., every dual of Λ is weakly equal to the canonical dual of Λ, then Λ is called a Riesz-type continuous g-frame. Theorem 1.9. [2] Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a continuous g-frame. Then Λ is a Riesz-type continuous g-frame if and only if RangeT ∗ We mention that the authors of this paper studied some properties of continuous g-frames and Riesz-type continuous g-frames in [3]. Λ = bK. DISJOINTNESS OF CONTINUOUS G-FRAMES 5 2. Disjointness of continuous g-frames In this section we study disjointness, strongly disjointness, weakly disjointness for continuous g-frames. We prove some results concern with these concepts and we construct a continuous g-frame by disjoint and strongly disjoint continuous g-frames. Definition 2.1. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(K,Kω) : ω ∈ Ω} be two continuous g-frames. Then Λ and Θ are called: is a closed subspace of bK. (i) Strongly disjoint, if RangeT ∗ (ii) Disjoint, if RangeT ∗ Λ∩RangeT ∗ (iii) Complementary pair, if RangeT ∗ Λ⊥RangeT ∗ Θ. Θ = {0} and RangeT ∗ Λ∩RangeT ∗ (iv) Strongly complementary pair, if RangeT ∗ (v) Weakly disjoint, if RangeT ∗ Λ ∩ RangeT ∗ Θ = bK. RangeT ∗ Λ+RangeT ∗ Θ Θ = {0} and RangeT ∗ Λ+ Λ ⊕ RangeT ∗ Θ = {0}. Θ = bK. Theorem 2.2. Let (Ω, µ) be a mesure space. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(K,Kω) : ω ∈ Ω} be two continuous g- frames. Consider Γ = {Γω ∈ B(H ⊕ K,Kω) : ω ∈ Ω} where Γω(h ⊕ k) = Λωh + Θωk, ω ∈ Ω, h ∈ H, k ∈ K. Then Λ and Θ are (i) Strongly disjoint if and only if there exist invertible operators L1 ∈ B(H) and L2 ∈ B(K) such that {ΛωL1 ∈ B(H,Kω) : ω ∈ Ω} , {ΘωL2 ∈ B(K,Kω) : ω ∈ Ω} and {∆ω ∈ B(H ⊕ K,Kω) : ω ∈ Ω} are Parseval continuous g-frames, where ∆ω(h ⊕ k) = ΛωL1h + ΘωL2k, ω ∈ Ω, h ∈ H, k ∈ K. (ii) Disjoint if and only if Γ is a continuous g-frame. (iii) Complementary pair if and only if Γ is a Riesz-type continuous g-frame. (iv) Strongly complementary pair if and only if they are strongly disjoint and Γ is a Riesz-type continuous g-frame. (v) Weakly disjoint if and only if {f ⊕ g : Γω(f ⊕ g) = 0, ω ∈ Ω} = {0}. 6 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Proof. (i) For any h ∈ H and k ∈ K we have ZΩ kΛωS−1/2 Λ h + ΘωS−1/2 Θ kk2dµ(ω) =ZΩ kΛωS−1/2 + 2ReZΩ(cid:10)ΛωS−1/2 =ZΩ kΛωS−1/2 hk2dµ(ω) +ZΩ kΘωS−1/2 k(cid:11)dµ(ω) hk2dµ(ω) +ZΩ kΘωS−1/2 h, ΘωS−1/2 Θ Θ Θ Λ = khk2 + kkk2 = kh ⊕ kk2. Λ Λ kk2dµ(ω) kk2dµ(ω) It is sufficient to take L1 = S−1/2 Conversly, for every h ∈ H and k ∈ K we have and L2 = S−1/2 Θ Λ . kh ⊕ kk2 =ZΩ k∆ω(h ⊕ k)k2dµ(ω) =ZΩ kΛωL1hk2dµ(ω) +ZΩ kΘωL2kk2dµ(ω) + 2ReZΩ(cid:10)ΛωL1h, ΘωL2k(cid:11)dµ(ω) = khk2 + kkk2 + 2ReZΩ(cid:10)ΛωL1h, ΘωL2k(cid:11)dµ(ω). ReZΩhΛωL1h, ΘωL2kidµ(ω) = 0, h ∈ H, k ∈ K. Therefore On the other hand ImZΩhΛωL1h, ΘωL2kidµ(ω) = −ReZΩhΛωL1(ih), ΘωL2kidµ(ω) = 0. Thus ZΩhΛωL1h, ΘωL2kidµ(ω) = 0, h ∈ H, k ∈ K. Now, since L1 and L2 are invertible operators, RangeT ∗ (ii) Let Γ be a continuous g-frame for H⊕K. Then for any h ∈ H, k ∈ K we have Λ⊥RangeT ∗ Λ. Λh + T ∗ AΓ(khk2 + kkk2) ≤ kT ∗ (2.1) Let there exist h1 ∈ H and k1 ∈ K such that T ∗ left hand inequality (2.1), h1 = k1 = 0 and so, T ∗ sequently, RangeT ∗ Θkk2 ≤ BΓ(khk2 + kkk2). Λh1 = T ∗ Λh1 = T ∗ Θ = {0}. Also, RangeT ∗ Λ ∩ RangeT ∗ Θk1. By the Θk1 = 0. Con- Θ = Λ + RangeT ∗ DISJOINTNESS OF CONTINUOUS G-FRAMES 7 RangeT ∗ Conversely, the operator Γ is a closed subspace of bK. L : RangeT ∗ Λ ⊕ RangeT ∗ Θ → RangeT ∗ L(F ⊕ G) = F + G Λ + RangeT ∗ Θ, is a bijective bounded operator, In fact kL(F ⊕ G)k2 = kF + Gk2 ≤ (kFk + kGk)2 ≤ 2(kFk2 + kGk2) = 2kF ⊕ Gk2. Then for any h ∈ H, k ∈ K we have kL−1k−2. min{AΛ, AΘ}.kh ⊕ kk2 ≤ZΩ kΓω(h ⊕ k)k2dµ(ω) Λh ⊕ T ∗ = kL(T ∗ ≤ kLk2. max{BΛ, BΘ}.kh ⊕ kk2. Θk)k2 (iii) Let Γ be a Riesz-type continuous g-frame for H ⊕ K. Then by Theorem 1.9, we have RangeT ∗ Λ + RangeT ∗ Θ = RangeT ∗ On the other hand, let φ ∈ RangeT ∗ Λh = T ∗ h ∈ H and k ∈ K such that T ∗ T ∗ Γ(h ⊕ 0) = φ = T ∗ Λ ∩ RangeT ∗ Θk = φ. Therefore Γ(0 ⊕ k), Γ is one-to-one, h = k = 0 and so φ = 0. since T ∗ Conversely, by the part (ii), Γ is a continuous g-frame and Θ. Then there exist Γ = bK. Θ = bK. RangeT ∗ Γ = RangeT ∗ Λ + RangeT ∗ So, by Theorem 1.9, Γ is a Riesz-type continuous g-frame. (iv) By applying the definition of strongly disjoint and (iii), the poof is completed. (v) Let Λ and Θ be weakly disjoint and h ∈ H, k ∈ K such that Γω(h ⊕ k) = 0 for all ω ∈ Ω, then T ∗ Λh = T ∗ Θ(−k) ∈ RangeT ∗ Λ ∩ RangeT ∗ Θ = {0}, and so, h = k = 0. Conversely, let φ ∈ RangeT ∗ there exist h ∈ H, k ∈ K such that T ∗ Γ(h ⊕ (−k)) = T ∗ T ∗ Λh = T ∗ Λh − T ∗ Θk = 0. Λ ∩ RangeT ∗ Θk = φ, therefore Θ. Then Hence h = k = 0, consequently, φ = 0. (cid:3) 8 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Lemma 2.3. [5] Suppose that T : K → H is a linear bounded, surjec- tive operator. Then there exists a linear bounded operator (called the pseudo-inverse of T ) T † : H → K for which T T †f = f , for any f ∈ H. By generalizing a result from [8] we get a following proposition to construct a continuous g-frame from disjoint continuous g-frames. Proposition 2.4. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two disjoint continuous g-frames and L1, L2 ∈ B(H). If L1 or L2 is surjective, then ΛL∗ 2 ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame. Proof. Let L1 is surjective, then by Lemma 2.3, there exist L† such that L1L† 2 = {ΛωL∗ 1 ∈ B(H) 1 = I, so (L† 1 + ΘωL∗ 1 + ΘL∗ 1)∗L∗ khk = k(L† 1 = I. Hence for any h ∈ H we have 1)∗L∗ 1hk ≤ kL† 1kkL∗ 1hk, thus 1hk ≥ khk kL∗ kL† 1k . Suppose Γω(h⊕ k) = Λωh + Θωk, for all h, k ∈ H and for all ω ∈ Ω. By part (ii) of Theorem 2.2, Γ = {Γω ∈ B(H⊕H) : ω ∈ Ω} is a continuous g-frame. Therefore, for any h ∈ H we have AΓ 1k2khk2 ≤ AΓkL∗ kL† 2hk2(cid:1) = AΓkL∗ 1hk2 ≤ AΓ(cid:0)kL∗ ≤ZΩ kΓω(L∗ ≤ BΓ(cid:0)kL∗ 1hk2 + kL∗ 2hk2 1h ⊕ L∗ 1h ⊕ L∗ 2hk2 1h ⊕ L∗ 1hk2 + kL∗ ≤ BΓkL∗ 2h)k2dµ(ω) Therefore we have 2hk2(cid:1) 1k2khk2 ≤ZΩ k(ΛωL1 + ΘωL2)hk2dµ(ω) ≤ 2BΓ. max{kL1k2,kL2k2}khk2. ≤ 2BΓ. max{kL1k2,kL2k2}khk2, h ∈ H. AΓ kL† The proof is similar whenever L2 is surjective. Corollary 2.5. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two disjoint continuous g-frames. Then Λ + Θ = {Λω + Θω ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame. (cid:3) DISJOINTNESS OF CONTINUOUS G-FRAMES 9 Proof. By considering L1 = L2 = I in Proposition 2.4 the proof is completed. (cid:3) In the following results, we construct a continuous g-frame by strongly disjoint continuous g-frames by generalizing a result from [10]. Proposition 2.6. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be strongly disjoint continuous g-frames and let L1, L2 ∈ B(H) such that L∗ 2L2 = AI for some A > 0. Then ΛL1 + ΘL2 = {ΛωL1 + ΘωL2 ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame. In particular, αΛ + βΘ = {αΛω + βΘω ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame for α, β ∈ C with α2 + β2 > 0. Proof. For any h ∈ H we have 1L1 + L∗ ZΩ k(ΛωL1 + ΘωL2)hk2dµ(ω) = kT ∗ ΛL1h + T ∗ ΘL2hk2 = kT ∗ ΛL1hk2 + kT ∗ ΘL2hk2, and (cid:0)BΛkL1k2 + BΘkL2k2(cid:1)khk2 ≥ kT ∗ ΛL1hk2 + kT ∗ ΘL2hk2 ≥ min{AΛ, AΘ}.(cid:0)kL1hk2 + kL2hk2(cid:1) 2L2)h, h(cid:11) = min{AΛ, AΘ}.(cid:10)(L∗ = A. min{AΛ, AΘ}khk2. 1L1 + L∗ (cid:3) By taking L1 = αI and L2 = βI the particular case is obvious. Proposition 2.7. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be strongly disjoint Parseval continuous g-frames and let L1, L2 ∈ B(H). Then L∗ 2L2 = AI for some A > 0 if and only if ΛL1 + ΘL2 = {ΛωL1 + ΘωL2 ∈ B(H,Kω) : ω ∈ Ω} is a tight continuous g-frame with bound A. In particular, αΛ + βΘ = {αΛω + βΘω ∈ B(H,Kω) : ω ∈ Ω} is a tight continuous g-frame if and only if α2 + β2 > 0 for α, β ∈ C. Proof. For any h ∈ H we have 1L1 + L∗ ZΩ k(ΛωL1 + ΘωL2)hk2dµ(ω) = kT ∗ ΛL1h + T ∗ ΘL2hk2 ΛL1hk2 + kT ∗ = kT ∗ = kL1hk2 + kL2hk2 1L1 + L∗ ΘL2hk2 2L2)h, h(cid:11) = Akhk2. =(cid:10)(L∗ In particular, it is sufficient to take L1 = αI and L2 = βI. (cid:3) 10 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Corollary 2.8. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be strongly disjoint Parseval continuous g-frames and let L1, L2 ∈ B(H). Then L∗ 2L2 = I if and only if ΛL1 + ΘL2 = {ΛωL1 + ΘωL2 ∈ B(H,Kω) : ω ∈ Ω} is a Parseval continuous g-frame. In particular, αΛ + βΘ = {αΛω + βΘω ∈ B(H,Kω) : ω ∈ Ω} is a Parseval continuous g-frame if and only if α2 + β2 = 1 for α, β ∈ C. 1L1 + L∗ Now, to get a dual continuous g-frames by strongly disjoint contin- uous g-frames we generalize results of [1] and [9]. Proposition 2.9. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Ψ = {Ψω ∈ B(K,Kω) : ω ∈ Ω} be duals of continuous g-frames Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} and Φ = {Φω ∈ B(K,Kω) : ω ∈ Ω}, respectively. If Λ, Φ and Θ, Ψ are strongly disjoint. Then Γ = {Γω ∈ B(H ⊕ K,Kω) : ω ∈ Ω} and ∆ = {∆ω ∈ B(H ⊕ K,Kω) : ω ∈ Ω} are dual continuous g-frames, where Γω(h ⊕ k) = Λωh + Ψωk, ∆ω(h ⊕ k) = Θωh + Φωk, for all ω ∈ Ω and for all h ∈ H, k ∈ K. Proof. It is clear that Γ and ∆ are continuous g-Bessel families for H ⊕ K. For any h1, h2 ∈ H and k1, k2 ∈ K, we have ZΩDΓω(h1 ⊕ k1), ∆ω(h2 ⊕ k2)Edµ(ω) =ZΩhΛωh1, Θωh2idµ(ω) +ZΩhΛωh1, Φωk2idµ(ω) +ZΩhΨωk1, Θωh2idµ(ω) +ZΩhΨωk1, Φωk2idµ(ω) = hh1, h2i + hk1, k2i = hh1 ⊕ k1, h2 ⊕ k2i. Thus by Proposition 3.2 of [2], Γ and ∆ are dual continuous g-frames for H ⊕ K. (cid:3) Example 2.10. Let F : Ω → H and G : Ω → K be two continu- ous frames. We define two families of bounded operators Λ = {Λω ∈ B(H, C2) : ω ∈ Ω} and Θ = {Θω ∈ B(H, C2) : ω ∈ Ω} where for all f ∈ H, ω ∈ Ω. It is obvious that Λ and Θ are continuous g-frames. Also for any f, g ∈ H Λωf =(cid:0)hf, F (ω)iH, 0(cid:1) ZΩhΘωf, Λωgidµ(ω) =ZΩ(cid:10)f, S−1 F f, F (ω)iH, 0(cid:1), , Θωf =(cid:0)hS−1 F F (ω)(cid:11)hF (ω), gidµ(ω) = hf, gi. DISJOINTNESS OF CONTINUOUS G-FRAMES 11 So, Λ and Θ are dual continuous g-frames. We also define two families of bounded operators Φ = {Φω ∈ B(K, C2) : ω ∈ Ω} and Ψ = {Ψω ∈ B(K, C2) : ω ∈ Ω} where Φωg =(cid:0)0,hS−1 G g, G(ω)iK(cid:1), Ψωg =(cid:0)0,hg, G(ω)iK(cid:1), for all g ∈ K, ω ∈ Ω. Similarly, Ψ and Φ are dual continuous g-frames. On the other hand, for any f ∈ H and g ∈ K Λf, T ∗ hT ∗ Φgi =ZΩhΛωf, Φωgidµ(ω) = 0, so Λ and Φ are strongly disjoint. Also, Θ and Ψ are strongly disjoint. Let us consider and for all f ∈ H and g ∈ K. Then F f, F (ω)iH,hS−1 Γω : H ⊕ K → C2, Γω(f ⊕ g) =(cid:0)hf, F (ω)iH,hg, G(ω)iK(cid:1), ∆ω : H ⊕ K → C2, ∆ω(f ⊕ g) =(cid:0)hS−1 G g, G(ω)iK(cid:1), ZΩ(cid:10)Γω(h1 ⊕ k1), ∆ω(h2 ⊕ k2)(cid:11)dµ(ω) =ZΩhh1, F (ω)iHhS−1 +ZΩhk1, G(ω)iKhS−1 F F (ω), h2iHdµ(ω) G G(ω), k2iKdµ(ω) = hh1, h2i + hk1, k2i = hh1 ⊕ k1, h2 ⊕ k2i. Which Proposition 2.9 confirm this result. Proposition 2.11. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two strongly disjoint continuous g-frames and 1 = {ΛωS−1 1 ∈ B(H,Kω) : ω ∈ Ω} is a dual for both ΛL∗ 1 ∈ B(H,Kω) : ω ∈ Ω} and ΛL∗ L1, L2 ∈ B(H). If L1 is surjective, theneΛL† Proof. It is obvious that eΛL† 2 are continuous g- Bessel families. Since L1 is surjective, then by Lemma 2.3 there exist 2 ∈ B(H,Kω) : ω ∈ Ω}. 1 + ΘL∗ 1, ΛL∗ 1 = {ΛωL∗ 2 = {ΛωL∗ 1 and ΛL∗ 1 + ΘωL∗ 1 + ΘL∗ Λ L† 12 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Lt 1 ∈ B(H,Kω), such that L1Lt 1 + ΘωL∗ 2)h, ΛωS−1 ZΩ(cid:10)(ΛωL∗ 1 = I. For any h, k ∈ H we have Λ L† 1k(cid:11)dµ(ω) 1h, ΛωS−1 Λ L† 1kidµ(ω) 2h, ΛωS−1 Λ L† 1kidµ(ω) ΛS−1 2h, T ∗ ΘL∗ 1ki + hT ∗ 1ki + 0 = hh, ki + 0 = hh, ki. 1ki Λ L† =ZΩhΛωL∗ +ZΩhΘωL∗ 1h, L† = hL∗ = hh, L1L† And also we have ZΩhΛωL∗ 1f, ΛωS−1 Λ L† 1gidµ(ω) = hL∗ 1f, L† 1gi = hf, L1L† 1gi = hf, gi. Therefore the proof is completed. (cid:3) Corollary 2.12. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two strongly disjoint continuous g-frames. Then Λ ∈ B(H,Kω) : ω ∈ Ω} is a dual for both Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Λ + Θ = {Λω + Θω ∈ B(H,Kω) : ω ∈ Ω}. Proof. By considering L1 = L2 = I in Proposition 2.11 the proof is completed. (cid:3) eΛ = {ΛωS−1 3. Some results related to Riesz-type continuous g-frames In this section by generalizing some results of [12], we get some equiv- alet conditions for Riesz-type continuous g-frames. Theorem 3.1. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a continuous g-frame. Then the following are equivalent: (i) Λ is a Riesz-type continuous g-frame. (ii) There exist constants A, B > 0 such that (3.1) (iii) If Akφk2 ≤ kTΛφk2 ≤ Bkφk2, ZΩhΛ∗ ωφ(ω), fidµ(ω) = 0 φ ∈ bK. for some φ ∈ bK and for any f ∈ H, then φ = 0. DISJOINTNESS OF CONTINUOUS G-FRAMES 13 Λf = φ. Then Proof. (i) ⇒ (ii) By Proposition 1.6, it remains to prove the left-hand such that T ∗ inequality in (3.1). By Theorem 1.9, for any φ ∈ bK, there exist f ∈ H kφk4 =(cid:16)ZΩ kΛω(f )k2dµ(ω)(cid:17)2 = hSΛf, fi2 ≤ kSΛfk2kfk2 ≤ 1 AΛkSΛfk2ZΩ kΛωfk2dµ(ω), and hence AΛkφk2 ≤ kSΛfk2 = kTΛT ∗ Λfk2 = kTΛφk2. (ii) ⇒ (iii) Let for some φ ∈ bK and any f ∈ H, we have ωφ(ω), fidµ(ω) = 0. hTΛφ, fi =ZωhΛ∗ Then TΛφ = 0 and by inequality (3.1), φ = 0. (iii) ⇒ (i) Since Λ is a continuous g-frame, TΛ is onto and by (iii) TΛ is one to one, so TΛ is invertible. Consequently, T ∗ Λ is invertible. Therefore, by Theorem 1.9, the proof is completed. (cid:3) Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(K,Kω) : ω ∈ Ω} be two continuous g-Bessel families. Consider the well defined operator SΘΛ : H → K, SΘΛ = TΘT ∗ Λ. Then hSΘΛf, gi =ZΩhΛωf, Θωgidµ(ω), f ∈ H, g ∈ K, and S∗ ΘΛ = SΛΘ. Theorem 3.2. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} and Θ = {Θω ∈ B(H,Kω) : ω ∈ Ω} be two continuous g-Bessel family such that SΛΘ = IH. Assume that L1, L2 : H → H are bounded linear operators so that L∗ 1L2 = I. Then the following statements are equivalent: (i) Γ = {ΛωL1 + ΘωL2 ∈ B(H,Kω) : ω ∈ Ω} is a Riesz-type continuous g-frame. (ii) The operator T ∗ ΘL2 is surjective. (iii) There exists a constant M > 0 such that ΛL1 + T ∗ Mkφk2 ≤ k(L∗ 1TΛ + L∗ 2TΘ)φk2, φ ∈ bK. =ZΩ kΛωL1fk2dµ(ω) +Z (cid:10)ΛωL1f , ΘωL2f(cid:11)dµ(ω) +Z (cid:10)ΘωL2f , ΛωL1f(cid:11)dµ(ω) +ZΩ kΘωL2fk2dµ(ω) =ZΩ kΛωL1fk2dµ(ω) + 2kfk2 +ZΩ kΘωL2fk2dµ(ω). 2kfk2 ≤ZΩ k(ΛωL1 + ΘωL2)fk2dµ(ω) ≤(cid:0)BΛkL1k2 + 2 + BΘkL2k2(cid:1)kfk2. So 14 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Proof. For any f ∈ H we have ZΩ k(ΛωL1 + ΘωL2)fk2dµ(ω) Hence Γ is a continuous g-frame. On the other hand T ∗ Γ = T ∗ ΛL1 + T ∗ ΘL2. By Theorem 1.9, (i) and (ii) are equivalent. (i) ⇔ (iii) It is concluded by Theorem 3.1 and TΓ = L∗ Proposition 3.3. Let (Ω, µ) be a measure space, that µ is σ-finite and Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} is a continuous g-frame. Suppose that Θ = {Θω ∈ B(K,Kω) : ω ∈ Ω} is a continuous g-Bessel family. If SΘΛ is surjective, then Θ is a continuous g-frame. If Θ is a continuous g- frame and Λ is a Riesz-type continuous g-frame then SΘΛ is surjective. 1TΛ + L∗ 2TΘ. (cid:3) Proof. Since SΘΛ is surjective, it follows that TΘ is surjective. On the other hand, by Proposition 1.6, TΘ is bounded. Hence by Theorem 1.7, Θ is a continuous g-frame. If Λ is a Riesz-type continuous g-frame and Θ is a continuous g-frame then by Theorems 1.9 and 1.7, T ∗ Λ and TΘ are surjective. So, SΘΛ is surjective. (cid:3) Theorem 3.4. Let Λ = {Λω ∈ B(H,Kω) : ω ∈ Ω} be a continuous g-frame and Θ = {Θω ∈ B(K,Kω) : ω ∈ Ω} be a continuous g-Bessel family. Suppose that there exists a number λ with 0 < λ < AΛ such that (3.2) kSΘΛf − SΛfk ≤ λkfk, f ∈ H. Then Λ is a Riesz-type continuous g-frame if and only if Θ is a Riesz- type continuous g-frame. DISJOINTNESS OF CONTINUOUS G-FRAMES 15 Proof. For all f ∈ H we have kSΘΛfk = kSΘΛf − SΛf + SΛfk ≥ kSΛfk − kSΘΛf − SΛfk ≥ (AΛ − λ)kfk. Then SΘΛ is injective with closed range. On the other hand, kSΛΘf − SΛfk ≤ k(SΘΛ − SΛ)∗kkfk ≤ λkfk. So SΛΘ is also injective with closed range.Therefore RangeSΘΛ = ker(SΛΘ)⊥ = H and RangeSΛΘ = H. Thus, SΘΛ and SΛΘ are invertible. Hence, T ∗ is invertible if and only if T ∗ Θ is invertible. Then by Theorem 1.9 the proof is completed. (cid:3) Λ Acknowledgment: References [1] M. R. Abdollahpour, Dilation of dual g-frames to dual g-Riesz bases, Banach Journal of Mathematical Analysis 9(1) (2015), 54-66. [2] M. R. Abdollahpour and M. H. Faroughi, Continuous g-frames in Hilbert spaces, Southeast Asian Bulletin of Mathematics 32(1) (2008), 1-19. [3] M. R. Abdollahpour and Y. Khedmati, On some properties of continuous g- frames and Riesz-type continuous g-frames, Indian Journal of Pure and Applied Mathematics 48(1) (2017), 59-74. [4] S. T. Ali, J. P. Antoine and J. P. Gazeau, Continuous frames in Hilbert space, Annals of Physics 222(1) (1993), 1-37. [5] O. Christensen and K. Jensen Torben, An introduction to the theory of bases, frames, and wavelets, (1999). [6] R. J. Duffin and A. C. Schaeffer, A class of nonharmonic Fourier series, Trans- actions of the American Mathematical Society 72(2) (1952), 341-366. [7] J. P. Gabardo and D. Han, Frames associated with measurable space, Advances in Computational Mathematics 18(2) (2003), 127-147. [8] X. Guo, Constructions of frames by disjoint frames, Numerical Functional Analysis and Optimization 35(5) (2014), 576-587. [9] X. Guo, Characterizations of disjointness of g-frames and constructions of g- frames in Hilbert spaces, Complex Analysis and Operator Theory 8(7) (2014), 1547-1563. [10] D. Han and D. Larson, Frames, bases and group representations, Memoirs of the American Mathematical Society 697 (2000), 149-182. [11] W. Sun, G-frames and g-Riesz bases, Journal of Mathematical Analysis and Applications 322(1) (2006), 437-452. [12] Z. Q. Xiang, New characterizations of Riesz-type frames and stability of alter- nate duals of continuous frames, Advances in Mathematical Physics (2013). 16 Y. KHEDMATI AND M. R. ABDOLLAHPOUR∗ Yavar Khedmati and Mohammad Reza Abdollahpour Department of Mathematics Faculty of Sciences University of Mohaghegh Ardabili Ardabil 56199-11367 Iran E-mail address: [email protected], [email protected] E-mail address: [email protected], [email protected]
1103.3906
2
1103
2011-05-20T20:41:57
An index formula in connection with meromorphic approximation
[ "math.FA", "math.CA", "math.CV" ]
Let $\Phi$ be a continuous $n\times n$ matrix-valued function on the unit circle $\T$ such that the $(k-1)$th singular value of the Hankel operator with symbol $\Phi$ is greater than the $k$th singular value. In this case, it is well-known that $\Phi$ has a unique superoptimal meromorphic approximant $Q$ in $H^{\infty}_{(k)}$; that is, $Q$ has at most $k$ poles in the unit disc $\mathbb{D}$ (i.e. the McMillan degree of $Q$ in $\mathbb{D}$ is at most $k$) and $Q$ minimizes the essential suprema of singular values $s_{j}((\Phi-Q)(\zeta))$, $j\geq0$, with respect to the lexicographic ordering. For each $j\geq 0$, the essential supremum of $s_{j}((\Phi-Q)(\zeta))$ is called the $j$th superoptimal singular value of $\Phi$ of degree $k$. We prove that if $\Phi$ has $n$ non-zero superoptimal singular values of degree $k$, then the Toeplitz operator $T_{\Phi-Q}$ with symbol $\Phi-Q$ is Fredholm and has index \[ \ind T_{\Phi-Q}=\dim\ker T_{\Phi-Q}=2k+\dim\mathcal{E}, \] where $\mathcal{E}=\{\xi\in\ker H_{Q}: \|H_{\Phi}\xi\|_{2}=\|(\Phi-Q)\xi\|_{2}\}$ and $H_{\Phi}$ denotes the Hankel operator with symbol $\Phi$. In fact, this result can be extended from continuous matrix-valued functions to the wider class of $k$-\emph{admissible} matrix-valued functions, i.e. essentially bounded $n\times n$ matrix-valued functions $\Phi$ on $\T$ for which the essential norm of the Hankel operator $H_{\Phi}$ is strictly less than the smallest non-zero superoptimal singular value of $\Phi$ of degree $k$.
math.FA
math
AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION ALBERTO A. CONDORI Abstract. Let Φ be a continuous n × n matrix-valued function on the unit circle T such that the (k − 1)th singular value of the Hankel operator with symbol Φ is greater than the kth singular value. In this case, it is well-known that Φ has a unique superoptimal meromorphic approximant Q in H∞ (k); that is, Q has at most k poles in the unit disc D (in the sense that the McMillan degree of Q in D is at most k) and Q minimizes the essential suprema of singular values sj ((Φ − Q)(ζ)), j ≥ 0, with respect to the lexicographic ordering. For each j ≥ 0, the essential supremum of sj ((Φ − Q)(ζ)) is called the jth superoptimal singular value of degree k of Φ. We prove that if Φ has n non-zero superoptimal singular values of degree k, then the Toeplitz operator TΦ−Q with symbol Φ−Q is Fredholm and has index ind TΦ−Q = dim ker TΦ−Q = 2k + dim E, where E = {ξ ∈ ker HQ : kHΦξk2 = k(Φ − Q)ξk2} and HΦ denotes the Hankel operator with symbol Φ. This result can in fact be extended from continuous matrix-valued functions to the wider class of k-admissible matrix- valued functions, i.e. essentially bounded n × n matrix-valued functions Φ on T for which the essential norm of the Hankel operator HΦ is strictly less than the smallest non-zero superoptimal singular value of degree k of Φ. 1. Introduction Let ϕ be a bounded measurable function defined on the unit circle T. For k ≥ 0, let H ∞ (k) denote the collection of meromorphic functions in the unit disc D which are bounded near T and have at most k poles in D (counting multiplicities). The Nehari-Takagi problem is to find a q ∈ H ∞ (k) which is closest to ϕ with respect to the L∞-norm, i.e. to find q ∈ H ∞ (k) such that kϕ − qk∞ = distL∞ (ϕ, H ∞ (k)) = inf f ∈H∞ (k) kϕ − fk∞. (k) to ϕ. Any such function q is called a best approximant in H ∞ Although a best approximant in H ∞ (k) need not be unique in general, if ϕ is a continuous function on T, then uniqueness holds. Moreover, under this assumption, it can be shown that the function defined by ϕ − q has constant modulus (equal to sk(Hϕ)) a.e. on T, the Toeplitz operator Tϕ−q is Fredholm, and where µ denotes the multiplicity of the singular value sk(Hϕ) of the Hankel opera- tor Hϕ with symbol ϕ (e.g. see Chapter 4 in [Pe1]). In fact, the best meromorphic ind Tϕ−q = 2k + µ, (1.1) 1991 Mathematics Subject Classification. Primary 47A57; Secondary 47B35, 46E40. Key words and phrases. Nehari-Takagi problem, Hankel and Toeplitz operators, best approx- imation, badly approximable matrix-valued functions, superoptimal approximation. 1 2 ALBERTO A. CONDORI approximant to ϕ in H ∞ erties. (k) is the unique function in H ∞ (k) that has these three prop- The index formula in (1.1) not only provides a uniqueness criterion for the best meromorphic approximant in H ∞ (k), it also appears in applications such as the study of singular values of Hankel operators with perturbed symbols (see [Pe2] or Chapter 7 in [Pe1]). This index formula can also be used to obtain a sharp estimate on the degree of the best meromorphic approximant to a given (scalar) rational function [Pe1]. Note that in the case of 2 × 2 matrix-valued rational functions, sharp estimates have also been obtained but only for their analytic approximants [PV]. These estimates were made without use of the index formula argument used for scalar functions. Thus, the main focus of this paper is to obtain an analogous index formula for matrix-valued functions on T. Unlike the scalar-case, if Φ is a continuous matrix- valued function on T, then Φ may not have a unique best meromorphic approximant Q with at most k poles in D (i.e. the McMillan degree of Q in D is at most k). However, the uniqueness of a superoptimal meromorphic approximant Q having at most k poles in D does hold under the additional assumption that sk(HΦ) < sk−1(HΦ). Therefore, the natural question arises whether the index formula in (1.1) holds for such matrix-valued functions Φ with superoptimal meromorphic approximant Q; that is, does ind TΦ−Q = 2k + µ (1.2) hold? Recall that for continuous n×n matrix-valued functions Ψ, the Toeplitz operator TΨ is Fredholm if and only if det Ψ does not vanish a.e. on T. It follows from known results regarding the error term Φ − Q that a necessary and sufficient condition for the Toeplitz operator TΦ−Q to be Fredholm (when Φ is continuous) is that all superoptimal singular values of degree k of Φ are non-zero. Unfortunately, as shown in Example 1.1 below, the index formula in (1.2) fails to hold under this additional assumption. Example 1.1. Consider the matrix-valued function 3 ¯z2 1 Φ = √2(cid:18) ¯z5 + 1 ¯z4 3 ¯z − 1 1 3 ¯z (cid:19) . It is not difficult to verify that the non-zero singular values of HΦ are s0(HΦ) = √10 3 , s1(HΦ) = s2(HΦ) = s3(HΦ) = 1, s4(HΦ) = 1 √2 , and s5(HΦ) = 1 3 . In particular, if µ denotes the multiplicity of the singular value s1(HΦ) = 1 of the Hankel operator HΦ, then 2k + µ = 5. Using an algorithm due to Peller and Young ([PY2] or section 17 of Chapter 14 (1)(M2) to Φ is in [Pe1]), it can be shown that the superoptimal approximant in H ∞ Q = 1 √2(cid:18) 1 3 ¯z O O O (cid:19) . However, ind TΦ−Q = dim ker TΦ−Q = 6 (by Theorem 7.4 of Chapter 14 in [Pe1] or Theorem 2.2 in [PY3]), because Φ − Q admits a ("thematic") factorization of the AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 3 form Φ − Q = 1 √2(cid:18) ¯z −1 z (cid:19)(cid:18) ¯z4 O 3 ¯z2 (cid:19) . O 1 1 Hence the index formula in (1.2) fails to hold for this choice of Φ. In this paper, we establish the correct analog to the index formula in (1.1), namely ind TΦ−Q = dim ker TΦ−Q = 2k + dimE for continuous n × n matrix-valued functions Φ such that sk(HΦ) < sk−1(HΦ) and whose superoptimal singular values of degree k are all non-zero, where E = {ξ ∈ ker HQ : kHΦξk2 = k(Φ − Q)ξk2}. In fact, we prove that our analog holds in the more general case of "k-admissible" bounded matrix-valued functions. This is accomplished using a result involving (k-admissible) matrix-valued weights for Hankel operators, the proof of which was inspired by Treil's approach to superoptimal approximation (actually the main ideas go back to [T2]). We also show that if all superoptimal singular values of degree k of Φ are equal, then our index formula agrees with the formula in (1.2) (see Corollary 6.4). This result is obtained using a characterization of the space of Schmidt vectors E(k)(Φ) that correspond to the singular value sk(HΦ) of the Hankel operator HΦ with symbol Φ. Note that this characterization involves any best approximant in H ∞ (k)(Mn) to Φ. The organization of the paper is as follows. All necessary background on su- peroptimal approximation appears in section 2. The characterization of the space of Schmidt vectors is given in section 3. We prove in section 4 that the Toeplitz operator induced by the error term Φ − Q is Fredholm and establish in section 5 a result concerning matrix-valued weights for Hankel operators. Section 6 contains proof that our analog to the index formula holds for k-admissible matrix-valued functions. 1.1. Notation and terminology. Throughout the paper, we use the following notation and terminology: m denotes normalized Lebesgue measure on the unit circle T so that m(T) = 1; O denotes the matrix-valued function which equals the zero matrix on T (its size will be clear in the context); Mm,n denotes the space of m × n matrices equipped with the operator norm k · kMm,n and Mn def = Mn,n; At denotes the transpose of a matrix A ∈ Mm,n; X(Mm,n) denotes the space of m× n matrix-valued functions on T whose entries belong to a space X of scalar functions on T and X(Cn) ζ∈T kΨ(ζ)kMm,n for Ψ ∈ L∞(Mm,n); def = ess sup kΨkL∞(Mm,n) Φt denotes the function Φt(ζ) B(X, Y ) denotes the collection of bounded linear operators T : X → Y between if T ∈ B(X, Y ), we say that a non-zero vector x ∈ X is a maximizing vector of T whenever kT xkY = kTk · kxkX; H and K denote Hilbert spaces; = (Φ(ζ))t, ζ ∈ T, when Φ ∈ L∞(Mm,n); normed spaces X and Y ; def = X(Mn,1); def 4 ALBERTO A. CONDORI if T ∈ B(H,K), the singular values sn(T ), n ≥ 0, of T are defined by sn(T ) = inf{kT − Rk : R ∈ B(H,K), rank R ≤ n} and the essential norm of T is defined by kTke = inf{kT − Kk : K ∈ B(H,K), K is a compact operator }; and if T ∈ B(H,K) and s is a singular value of T , a non-zero vector x ∈ H is called a Schmidt vector corresponding to s whenever T ∗T x = s2x. 2. Background 2.1. Best and superoptimal approximation in H ∞ class H ∞ product. (k)(Mm,n). To introduce the (k)(Mm,n), we must first define the notion of a finite Blaschke-Potapov A matrix-valued function B ∈ H ∞(Mn) is called a finite Blaschke-Potapov prod- uct if it admits a factorization of the form B = U B1B2 . . . Bm, where U is a unitary matrix and, for each 1 ≤ j ≤ m, Pj + (I − Pj) Bj = z − λj 1 − ¯λj z for some λj ∈ D and orthogonal projection Pj on Cn. The degree of the Blaschke- Potapov product B is defined to be deg B def = rank Pj. m Xj=1 Alternatively, B is a finite Blaschke-Potapov product of degree k if and only if B admits a factorization of the form B(z) = U0(cid:18) z−a1 1−¯a1z O O In−1 (cid:19) U1 . . . Uk−1(cid:18) z−ak 1−¯akz O O In−1 (cid:19) Uk, (2.1) where a1, . . . , ak ∈ D; U0, U1, . . . , Uk are constant n× n unitary matrices; and In−1 denotes the (n − 1) × (n − 1) identity matrix. It turns out that every invariant subspace L of multiplication by z on H 2(Cn) of finite codimension is of the form BH 2(Cn) for some Blaschke-Potapov product B of finite degree. Moreover, the degree of B equals codimL (e.g. see Lemma 5.1 in Chapter 2 of [Pe1]). A matrix-valued function Q ∈ L∞(Mm,n) is said to have at most k poles in D if there is a finite Blaschke-Potapov product B of degree k such that QB ∈ H ∞(Mm,n). We denote the collection of m × n matrix-valued functions Q that have at most k poles in D by H ∞ For Q ∈ L∞(Mm,n) with at most k poles in D, the McMillan degree of Q in D is the smallest number j ≥ 0 such that Q has at most j poles in D. In particu- lar, H ∞ (k)(Mm,n) consists of matrix-valued functions Q ∈ L∞(Mm,n) which can be written in the form Q = R + F for some F ∈ H ∞(Mm,n) and some rational m × n matrix-valued function R with poles in D such that the McMillan degree of R in D is at most k. In this paper, we do not need (explicitly) the general definition of McMillan degree (which omits the restriction to the disc D) and thus refer the (k)(Mm,n). AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 5 interested reader to Chapter 2 in [Pe1] for further information regarding McMillan degree. Definition 2.1. Let k ≥ 0. Given an m×n matrix-valued function Φ ∈ L∞(Mm,n), we say that Q is a best approximant in H ∞ (k)(Mm,n) and kΦ − QkL∞(Mm,n) = distL∞(Mm,n)(Φ, H ∞ (k)(Mm,n) to Φ if Q ∈ H ∞ (k)(Mm,n)). Note that by a compactness argument, a matrix-valued function Φ ∈ L∞(Mm,n) always has a best approximant in H ∞ (k)(Mm,n). That is, the set Ω(k) 0 (Φ) def = (Q ∈ H ∞ (k)(Mm,n) : Q minimizes ess sup ζ∈T kΦ(ζ) − Q(ζ)kMm,n) is always non-empty (e.g. see section 3 in Chapter 4 of [Pe1]). As in the case of scalar-valued bounded functions, Hankel operators on Hardy spaces are very useful tools in the study of best approximation by matrix-valued (k)(Mm,n). For a matrix-valued function Φ ∈ L∞(Mm,n), we define functions in H ∞ the Hankel operator HΦ by HΦf = P−Φf, for f ∈ H 2(Cn), where P− denotes the orthogonal projection of L2(Cm) onto H 2 H 2(Cm). It is well-known ([T1] or section 3 of Chapter 4 in [Pe1]) that −(Cm) = L2(Cm) ⊖ distL∞(Mm,n)(Φ, H ∞ (k)(Mm,n)) = sk(HΦ). (2.2) However in contrast to the case of scalar-valued functions, it is known that the condition kHΦke < sk(HΦ) does not guarantee uniqueness of a best approximant in H ∞ (k)(Mm,n) to Φ. Since the set of best approximants Ω(k) 0 (Φ) to Φ may contain distinct elements, it is natural to refine the notion of optimality if possible to obtain the "very best" matrix-valued function in Ω(k) Definition 2.2. Let k ≥ 0 and Φ ∈ L∞(Mm,n). For j > 0, define the sets sj(Φ(ζ) − Q(ζ))) . j−1(Φ) : Q minimizes ess sup ζ∈T = (Q ∈ Ω(k) 0 (Φ). Ω(k) (Φ) def j We say that Q is a superoptimal approximant in H ∞ (k)(Mm,n) to Φ if Q belongs to min{m,n}−1(Φ) and in this case we define the superoptimal singular (Φ) = Ω(k) Ω(k) j \j≥0 values of degree k of Φ by t(k) j (Φ) = ess sup ζ∈T sj((Φ − Q)(ζ)) for j ≥ 0. In the case k = 0, we also use the notations Ωj(Φ) and tj(Φ) to denote Ω(0) t(0) j (Φ), respectively, for j ≥ 0. In [T2], Treil proved that a unique superoptimal approximant Q in H ∞ (k)(Mm,n) to Φ exists whenever Φ ∈ (H ∞ + C)(Mm,n) and sk(HΦ) < sk−1(HΦ). (Recall that H ∞ + C denotes the closed subalgebra of L∞ that consists of functions of the form f + g with f ∈ H ∞ and g ∈ C(T).) Shorty after, Peller and Young also j (Φ) and 6 ALBERTO A. CONDORI proved this result in [PY2] using a diagonalization argument which also constructs (in principle) the superoptimal approximant. j A matrix-valued function Φ ∈ L∞(Mm,n) is called k-admissible if sk(HΦ) < sk−1(HΦ) and kHΦke is strictly less than the smallest non-zero number in the set {t(k) (Φ)}j≥0. (Note that the statement regarding the singular values of the Hankel operator is vacuous when k = 0.) For notational simplicity, we refer to 0-admissible matrix-valued functions as admissible. In particular, any matrix-valued function Φ that belongs to (H ∞ + C)(Mm,n) is admissible because the Hankel operator HΦ has essential norm equal to zero. It is now known (see section 17 of Chapter 14 in [Pe1]) that if Φ ∈ L∞(Mm,n) (k)(Mm,n) is k-admissible, then Φ has a unique superoptimal approximant Q in H ∞ and sj((Φ − Q)(ζ)) = t(k) j (Φ) a.e. ζ ∈ T, j ≥ 0. (2.3) 2.2. Very badly approximable functions. Let G ∈ L∞(Mm,n). We say that G is very badly approximable if the matrix-valued function O is a superoptimal approximant in H ∞(Mm,n) to G. It is well-known that if G is an admissible very badly approximable m×n matrix- valued function such that m ≤ n and tm−1(G) > 0, then the Toeplitz operator TzG : H 2(Cn) → H 2(Cm) has dense range. A proof can be found in Chapter 14 of [Pe1]. This result was originally proved in the case of matrix-valued functions G ∈ (H ∞ + C)(Mm,n) in [PY1]. Recall that the Toeplitz operator TΨ : H 2(Cm) → H 2(Cn) with symbol Ψ ∈ L∞(Mm,n) is defined by TΨf = P+Ψf, for f ∈ H 2(Cn), and P+ denotes the orthogonal projection of L2(Cm) onto H 2(Cm). Let k ≥ 0, Φ ∈ L∞(Mm,n) and ℓ ≥ 0 be fixed. It follows from Definition 2.2 (Φ), and (Φ) whenever F ∈ Ωj(Φ − Q) for 0 ≤ j ≤ ℓ. In particular, if Φ has a (k)(Mm,n), then Φ−Q is very badly approximable (Φ), then O belongs to Ωℓ(Φ − Q), tj(Φ − Q) = t(k) j j ℓ that if Q ∈ Ω(k) Q + F ∈ Ω(k) superoptimal approximant Q in H ∞ and tj(Φ − Q) = t(k) (Φ) for all j ≥ 0. j In [PT2], Peller and Treil characterized admissible very badly approximable func- tions in terms of certain families of subspaces. To state their result, let Ψ be a matrix-valued function in L∞(Mm,n) and σ > 0. For ζ ∈ T, we denote by Sσ Ψ(ζ) the linear span of all Schmidt vectors of Ψ(ζ) that correspond to the singular values of Ψ(ζ) that are greater than or equal to σ. Note that the subspaces Sσ Ψ(ζ) are defined for almost all ζ ∈ T. Theorem 2.3 ([PT2]). Suppose Ψ is an admissible matrix-valued function in L∞(Mm,n). Then Ψ is very badly approximable if and only if for each σ > 0, there are functions ξ1, . . . , ξℓ ∈ ker TΨ such that Sσ Ψ(ζ) = span{ ξj(ζ) : 1 ≤ j ≤ ℓ } for a.e. ζ ∈ T. For proofs of many of the previously mentioned results, we refer the reader to [Pe1] and the references therein. AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 7 3. Schmidt vectors of Hankel operators Henceforth, let k be an integer such that k ≥ 1 and Φ ∈ L∞(Mm,n). In this section, we study the collection of Schmidt vectors E(k)(Φ) def = (cid:8)ξ ∈ H 2(Cn) : H ∗ ΦHΦξ = s2 k(HΦ)ξ(cid:9) which correspond to the singular value sk(HΦ) of the Hankel operator HΦ. To improve the transparency of some computations, we make use of the flip operator J : L2(Cm) → L2(Cm), defined by Jf = ¯z ¯f for f ∈ L2(Cm). It is easy to see that J is an involution and satisfies JHΦ = H ∗ Φt J and HΦJ = JH ∗ Φt , because J intertwines with the Riesz projections, i.e. J P+ = P−J. It follows that sj(HΦ) = sj(HΦt ) holds for all j ≥ 0 and JHΦE(k)(Φ) = E(k)(Φt). We make use of the following well-known lemma. Lemma 3.1. Let m ≥ 0. If T ∈ B(H,K) satisfies sm(T ) > kTke, then sm(T ) is an eigenvalue of (T ∗T )1/2. Proof of this lemma can be based on the fact that any point λ in the spectrum of (T ∗T )1/2 that does not belong to the essential spectrum of (T ∗T )1/2 must be an isolated eigenvalue of finite multiplicity of (T ∗T )1/2 (see Chapter XI in [Co]). This fact is a consequence of the Spectral Theorem for normal operators on a Hilbert space. For the remainder of this section, we assume that kHΦke < sk(HΦ) < sk−1(HΦ). (3.1) Then Lemma 3.1 implies that s = sk(HΦ) is the kth largest eigenvalue of (H ∗ and has finite multiplicity µ = dim ker(H ∗ ΦHΦ)1/2 ΦHΦ − s2I). Therefore, sk−1(HΦ) > sk(HΦ) = . . . = sk+µ−1(HΦ) > sk+µ(HΦ). def = rank HQ ≤ k and B be a def = QB Let Q be a best approximant in H ∞ (k)(Mm,n) to Φ, r finite Blaschke-Potapov product such that ker HQ = BH 2(Cn). Clearly, F belongs to H ∞(Mm,n) and sk(HΦ) ≤ kHΦBH 2(Cn)k = kHΦBk ≤ kΦB − FkL∞(Mm,n) = kΦ − QkL∞(Mm,n), because BH 2(Cn) has co-dimension r ≤ k and B takes unitary values on T. For- mula (2.2) allows us to conclude that sk(HΦ) = kHΦBk = kΦ − QkL∞(Mm,n) (3.2) and so codim BH 2(Cn) = k; otherwise, sk−1(HΦ) ≤ kHΦBH 2(Cn)k = sk(HΦ) holds, contradicting the assumption sk(HΦ) < sk−1(HΦ). 8 ALBERTO A. CONDORI It is known that E(k)(Φ) consists of maximizing vectors of HΦ−Q and E(k)(Φ) ⊆ ker HQ (Lemma 17.2 in Chapter 14 of [Pe1]). Therefore, if ξ ∈ E(k)(Φ) satisfies kξk2 = 1, then sk(HΦ) = kHΦ − HQk = kP−(Φ − Q)ξk2 ≤ k(Φ − Q)ξk2 ≤ kΦ − Qk∞ = sk(HΦ). Thus, for any ξ ∈ E(k)(Φ), it follows that HΦξ = HΦ−Qξ =(Φ − Q)ξ, k(Φ − Q)(ζ) ξ(ζ)kCn =sk(HΦ)kξ(ζ)kCn for a.e. ζ ∈ T, kHΦ−Qk = kΦ − Qk∞ =sk(HΦ) and kHΦ−Qke = kHΦke. (3.3) (3.4) (3.5) The latter result in (3.5) is an immediate consequence of the formula (Theorem 3.8 in Chapter 4 of [Pe1]) kHΨke = distL∞(Mm,n)(Ψ, (H ∞ + C)(Mm,n)), It is now easy to see that the following result holds. for Ψ ∈ L∞(Mm,n). Theorem 3.2. Suppose Φ ∈ L∞(Mm,n) satisfies (3.1). If Q is a best approximant in H ∞ (k)(Mm,n) to Φ, then E(k)(Φ) = { ξ ∈ ker HQ : kHΦ−Qξk2 = sk(HΦ)kξk2, JHΦξ ∈ ker HQt } . Proof. Suppose Q is any best approximant in H ∞ (k)(Mm,n) to Φ. Recalling that transposition does not change the norm of a matrix-valued function, we see that Qt is a best approximant in H ∞ (k)(Mm,n) to Φt by (2.2). Therefore, E(k)(Φ) ⊆ ker HQ and so JHΦE(k)(Φ) = E(k)(Φt) ⊆ ker HQt . Thus, ξ ∈ ker HQ, JHΦξ ∈ ker HQt , and kHΦ−Qξk2 = sk(HΦ)kξk2 by (3.3) and (3.5). On the other hand, if ξ ∈ H 2(Cn) satisfies kHΦ−Qξk2 = sk(HΦ)kξk2, then ξ is a k(HΦ)ξ, because kHΦ−Qk = Q, it (cid:3) maximizing vector of HΦ−Q and so H ∗ sk(HΦ). Therefore if, in addition, ξ ∈ ker HQ and HΦξ ∈ J ker HQt = ker H ∗ must be that H ∗ Φ−QHΦ−Qξ = s2 ΦHΦξ = s2 k(HΦ)ξ and so ξ ∈ E(k)(Φ), as desired. 4. Fredholm Toeplitz operators and matrix weights In this section, we prove the following result. Theorem 4.1. Let Ψ ∈ L∞(Mn) be an admissible very badly approximable func- tion. If Ψ has n non-zero superoptimal singular values of degree 0, then the Toeplitz operator TΨ is Fredholm and ind TΨ = dim ker TΨ > 0. It is well-known and contained in the literature that admissible very badly ap- proximable functions induce Toeplitz operators with non-trivial finite dimensional kernels (e.g. see Chapter 14 in [Pe1]); however, we provide a simple proof of this fact using matrix-valued weights (defined below). It is the hope of the author that this will clarify some of the (similar) ideas used in the next section. AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 9 Let W be an n × n matrix-valued weight ; that is, a bounded matrix-valued function whose values are non-negative n × n matrices. We define the weighted inner product def (f, g)W (W (ζ)f (ζ), g(ζ))dm(ζ) for f, g ∈ L2(Cn). = ZT As usual, k·kW denotes the norm induced by the inner product (·,·)W , i.e. kfk2 (f, f )W for f ∈ L2(Cn). Recall that an operator T ∈ B(H,K) is said to be Fredholm if Range T is closed in K, dim ker T < ∞, and dim ker T ∗ < ∞. In this case, the index of T is defined by W = ind T def = dim ker T − dim ker T ∗. More information concerning Fredholm operators and index can be found in Chap- ter XI of [Co]. Proof of Theorem 4.1. By Lemma 3.1, the admissibility of Ψ guarantees that kHΨk2 is an eigenvalue of H ∗ ΨHΨ of finite multiplicity and thus a maximizing vector of HΨ. It follows that ker TΨ = { f ∈ H 2(Cn) : kHΨfk2 = kΨfk2 } is non-empty; after all, since Ψ is (very) badly approximable, then kHΨk = kΨk∞ and the inequalities kHΨfk2 ≤ kΨfk2 ≤ kΨk∞kfk2 = kHΨk · kfk2 are all equalities for any maximizing vector f of the Hankel operator HΨ. Let us show that dim ker TΨ < ∞. To this end, consider the matrix-valued weight W = Ψ∗Ψ. Since Ψ is an admissible very badly approximable function, it follows from (2.3) that sj(Ψ(ζ)) = tj is non-zero (by the admissibility of Ψ) for a.e. ζ ∈ T, 0 ≤ j ≤ n − 1. In particular, W is invertible a.e. on T and kW (ζ)−1k = s−1 n−1 for a.e. ζ ∈ T. n−1(W (ζ)) = t−2 It is easy to see that { ξ ∈ W 1/2H 2(Cn) : kHΨW −1/2ξk2 = kξk2 } = { f ∈ H 2(Cn) : kHΨfk2 = kW 1/2fk2 } = { f ∈ H 2(Cn) : kHΨfk2 = kΨfk2 } (4.1) = ker TΨ and so the operator HΨW −1/2 defined on W 1/2H 2(Cn) and equipped with the L2-norm (on its domain and range) has operator norm equal to 1. Furthermore, the corresponding space of maximizing vectors of HΨW −1/2 equals ker TΨ because kHΨfk2 ≤ kΨfk2 holds for all f ∈ H 2(Cn). The essential norm of this operator also admits the estimate kHΨW −1/2W 1/2H 2(Cn)ke ≤ kHΨkekW −1/2k∞ = kHΨket−1 n−1(Ψ) < 1 due to the admissibility of Ψ. Therefore, the space of maximizing vectors of the operator HΨW −1/2W 1/2H 2(Cn) is finite dimensional. In view of (4.1), we deduce now that ker TΨ is non-empty and finite dimensional. zΨ has trivial Ψ is trivial as well. Therefore it suffices to show that TΨ has By Theorem 5.4 in Chapter 14 of [Pe1], the Toeplitz operator T ∗ kernel and so ker T ∗ closed range. To this end, let τΨ def = sup(cid:8)kHΨfk2 : kfkW = 1, f ∈ (ker TΨ)⊥(cid:9) . (4.2) 10 ALBERTO A. CONDORI Clearly, τΨ ≤ 1 because kHΨfk2 ≤ kfkW for all f ∈ H 2(Cn). Moreover, the trivial 2 is valid for all f ∈ H 2(Cn) and implies that identity kΨfk2 2 + kHΨfk2 2 = kTΨfk2 kfk2 W ≤ kTΨfk2 2 + τ 2 Ψkfk2 W for f ∈ (ker TΨ)⊥, or equivalently, (1 − τ 2 Ψ)kfk2 W ≤ kTΨfk2 2 for f ∈ (ker TΨ)⊥. Since the matrix-valued weight W satisfies the inequality kfkW ≥ tn−1kfk2 for f ∈ H 2(Cn), then Ψ)t2 (1 − τ 2 n−1kfk2 2 ≤ kTΨfk2 (4.3) Thus, the Toeplitz operator TΨ is bounded from below on (ker TΨ)⊥ if τΨ < 1. In particular, this implies that the Toeplitz operator TΨ has closed range because the restriction TΨ to (ker TΨ)⊥ does. To complete the proof, it remains to show that τΨ < 1. The following lemma is needed, the proof of which can be deduced from the ideas used above and Lemma 3.1. 2 for f ∈ (ker TΨ)⊥. Lemma 4.2 ([PT2]). Let W be an invertible admissible weight for a Hankel op- erator HΨ such that W (ζ) ≥ a2I, a > kHΨke, and let K be a closed subspace of H 2(Cn). If q = sup{kHΨfk : f ∈ K,kfkW = 1} equals 1, then there exists a (non-zero) vector f0 ∈ K such that kHΨf0k2 = kf0kW . By Lemma 4.2, if K = (ker TΨ)⊥ = {f ∈ H 2(Cn) : kHΨfk2 = kfkW}⊥, W = Ψ∗Ψ (as before), a = tn−1, and τΨ = 1, then there is an f0 ∈ K such that kHΨf0k2 = kf0kW and so f0 ∈ ker TΨ, a contradiction to our choice of K. This completes the proof of Theorem 4.1. (cid:3) The following corollary is a well-known consequence of results concerning "the- matic" factorizations of very badly approximable functions (see [PY1] and [PY3], or Chapter 14 in [Pe1]). Additionally, it is now a consequence of Theorem 4.1. Corollary 4.3. If Ψ satisfies the hypotheses of Theorem 4.1, then dim ker TΨ ≥ n. Proof. By Theorem 4.1, TΨ is Fredholm and so TzΨ is Fredholm as well. Moreover, the assumptions on Ψ imply that the Toeplitz operator TzΨ has dense range (see section 2.2) and thus Range TzΨ = H 2(Cn). It follows that, for each c ∈ Cn, there is an f ∈ H 2(Cn) such that TzΨf = c and so TΨf = P+¯zP+(zΨ)f = 0, i.e. f ∈ ker TΨ. Hence dim ker TΨ ≥ n. Corollary 4.4. Suppose Φ is k-admissible and Q is the superoptimal approximant in H ∞ (k)(Mn) of Φ. If the number of non-zero superoptimal singular values of degree k of Φ equals n, then the Toeplitz operator TΦ−Q is Fredholm and ind TΦ−Q = dim ker TΦ−Q > 0. Proof. As observed in section 2.2, the matrix-valued function Ψ = Φ − Q is very badly approximable and tj(Ψ) = t(k) (Φ) for all j ≥ 0. In particular, Ψ is admissible and has n non-zero superoptimal singular values. Hence, the conclusion follows from Theorem 4.1. (cid:3) (cid:3) j AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 11 5. k-Admissible weights for Hankel operators −(Cm) and a matrix- Definition 5.1. Given a Hankel operator HΨ : H 2(Cn) → H 2 valued weight W , we say that W is a k-admissible weight for HΨ if the inequality holds for all f which belong to an invariant subspace of H 2(Cn) under multiplication by z whose codimension is at most k. kHΨfk2 ≤ kfkW Equivalently, W is a k-admissible weight if the operator TW − H ∗ ΨHΨ is non- negative on an invariant subspace of H 2(Cn) under multiplication by z whose codi- mension is at most k. (Note that our definition of k-admissibility for weights is stated differently than in [T2] yet it is equivalent.) The reason for studying k-admissible weights for Hankel operators arises nat- urally from the problem of best approximation in H ∞ Indeed for Φ ∈ L∞(Mm,n), it is not difficult to see (by Nehari's theorem, i.e. equation (2.2) with k = 0) that finding a best approximant Q in H ∞ (k)(Mm,n) to Φ is equivalent to finding a non-trivial (closed) invariant subspace M of H 2(Cn) under multiplication by z with codimension at most k such that kHΦfk2 ≤ sk(HΦ)kfk2 for all f ∈ M. That is, the best approximation problem in H ∞ (k)(Mm,n) is equivalent to verifying that the matrix-valued weight W = s2 k(HΦ)I is k-admissible for the Hankel operator HΦ, where I denotes the identity on H 2(Cn). (k)(Mm,n). The collection of k-admissible weights for Hankel operators was characterized by Treil. Theorem 5.2 ([T2]). Suppose Φ ∈ L∞(Mm,n). A matrix-valued weight W is k-admissible for HΦ if and only if there is a Q ∈ H ∞ (Φ − Q)∗(Φ − Q) ≤ W. (k)(Mm,n) such that The proof is contained in that of Theorem 16.1 in [T2]. Actually, one does not need to use spectral measures nor the Iokhvidov-Ky Fan theorem; the version above can be deduced from the case (of the Weighted Nehari Problem) k = 0 (see Theorem 6.1 in [Pe1]) due to the characterization of the invariant subspaces of multiplication by z on H 2(Cn) with finite codimension stated in section 2.1. Given Φ ∈ L∞(Mm,n), an n × n matrix-valued weight W , and a subspace L of H 2(Cn), we define (5.1) A few observations are in order. Let Φ ∈ L∞(Mm,n) satisfy (3.1) and Q be a = {ξ ∈ L : kHΦξk2 = kξkW}. EW (Φ;L) def best approximant in H ∞ (k)(Mm,n) to Φ. If W def = (Φ − Q)∗(Φ − Q), then W is a k-admissible weight for HΦ, (1) the operator TW − H ∗ (2) the kernel of (TW − H ∗ (3) the space of maximizing vectors of the operator ΦHΦ) ker HQ coincides with EW (Φ; ker HQ), ΦHΦ is non-negative when restricted to ker HQ, i.e. HΦ : (ker HQ,k · kW ) → (H 2 −(Cn),k · k2) (5.2) equals EW (Φ; ker HQ), and (4) HΦξ = (Φ − Q)ξ and HΦξ ∈ J ker HQt hold for ξ ∈ EW (Φ; ker HQ). 12 ALBERTO A. CONDORI Remark 5.3. It is easy to see that EW (Φ; ker HQ) contains the non-empty finite dimensional subspace E(k)(Φ) (e.g. see equation (3.3)) however these sets need not be equal in general. On the other hand, if W equals the identity a.e. on T, then E(k)(Φ) = EW (Φ; ker HQ) holds (by Lemma 5.6 below). We now provide a suitable extension of Theorem 2.3 to the class of k-admissible functions. Theorem 5.4. Suppose Φ is k-admissible and Q is the superoptimal approximant in H ∞ (k)(Mn) of Φ. If the number of non-zero superoptimal singular values of degree k of Φ equals n, then there are functions ξ1, . . . , ξℓ ∈ EW (Φ; ker HQ), where W = (Φ − Q)∗(Φ − Q), such that Sσ Φ−Q(ζ) = span{ ξj(ζ) : 1 ≤ j ≤ ℓ } for a.e. ζ ∈ T. (5.3) Remark 5.5. The conclusion of Theorem 5.4 is not an immediate consequence of Theorem 2.3. Indeed, the very bad approximability of Φ − Q only implies that there are finitely many functions ξ1, . . . , ξℓ ∈ ker TΦ−Q such that (5.3) holds; how- ever, the conclusion of Theorem 5.4 states that one can choose these functions in EW (Φ; ker HQ). In fact, EW (Φ; ker HQ) = ker HQ∩ker TΦ−Q is a subset of ker TΦ−Q, but these sets are not equal when k > 0. The proof of Theorem 5.4 follows from ideas in the "matrix-weight proof" of Theorem 4.1 in [PT2]. Some of these arguments were used earlier in [T2] in a different context. def Proof of Theorem 5.4. Let L = ker HQ. The matrix-valued weight W is a k- admissible weight for the Hankel operator HΦ on L, as observed above. Moreover, in view of the choice of the superoptimal approximant Q of Φ, Φ − Q is very badly approximable and (2.3) holds. Let σ0, . . . , σr denote all of the distinct superoptimal singular values of degree k of Φ arranged in decreasing order. weights Wj (ζ) = Λj((Φ − Q)∗(ζ)(Φ − Q)(ζ)) for ζ ∈ T. For 0 ≤ j ≤ r, define the functions Λj(x) = max{x, σ2 j} for x ≥ 0 and the Notice that W (ζ) ≤ Wj(ζ) for a.e. ζ ∈ T and so the weight Wj is k-admissible = EWj (Φ;L) for 0 ≤ j ≤ r. Then E(k)(Φ) ⊆ E0 for the Hankel operator HΦ. Let Ej and E0 ⊂ . . . ⊂ Er = EW (Φ;L). Moreover, for each 0 ≤ j ≤ r, Wj (ζ) is invertible a.e. on T and kWj(ζ)−1/2k = σj for a.e. ζ ∈ T because def kWj(ζ)−1k = s−1 n−1(Wj(ζ)) = σ−2 j Fix 0 ≤ j ≤ r. By considering the operator HΦW −1/2 j j L (equipped with for a.e. ζ ∈ T. on W 1/2 (5.4) the L2-norm), we see that W 1/2 kHΦW −1/2 j j Lk = kHΦ : (L,k · kWj ) → (H 2 −(Cn),k · k2)k = 1 and kHΦW −1/2 j W 1/2 j Lke ≤ kHΦkekW −1/2 j k∞ ≤ kHΦkeσ−1 j < 1. j Therefore the space of maximizing vectors Ej, as defined in (5.1), of the operator HΦW −1/2 W 1/2 j L is non-empty and finite dimensional. Let = kHΦL ⊖Wj Ejk = sup{ kHΦξk2 : ξ ∈ L ⊖Wj Ej,kξkWj = 1 }, qj def AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 13 where def L ⊖Wj Ej = { g ∈ L : (f, g)Wj = 0 for all f ∈ Ej }. Clearly, qj ≤ 1. If qj = 1, then the restriction to L ⊖Wj Ej of the operator in (5.2) must also have a maximizing vector by Lemma 4.2, a contradiction as Ej contains all maximizing vectors. Hence, it must be that qj < 1. def σj = span{ f (ζ) : f ∈ Ej }. It is easy to verify that dim Ej(ζ) For ζ ∈ T, let Ej (ζ) equals a constant for a.e. ζ ∈ T (e.g. see Chapter VII in [He]). Therefore, it suffices Φ−Q(ζ) for a.e. ζ ∈ T. to show that Ej(ζ) = S Assume that there is a function f ∈ Ej such that f (ζ) /∈ S σj Φ−Q(ζ) on a set σj Φ−Q(ζ) equals a constant a.e. on T, it follows Φ−Q(ζ) holds for a.e. ζ ∈ T. However, this implies that k(Φ − of positive measure. Since dim S that f (ζ) /∈ S Q)(ζ)f (ζ)kCn < σjkf (ζ)kCn and so σj kHΦfk2 ≤ k(Φ − Q)fk2 < σjkfk2 ≤ kfkWj , because f ∈ L, a contradiction to the assumption f ∈ Ej. Thus, Ej(ζ) ⊆ S for a.e. ζ ∈ T. Ej(ζ) is a proper subspace of S it must be that Ej(ζ) is a proper subspace of S dim S Assume now for the sake of contradiction that the subspace-valued function σj Φ−Q(ζ) on a set of positive measure. In this case, Φ−Q(ζ) for a.e. ζ ∈ T because Φ−Q(ζ) and dim Ej(ζ) are constant for a.e. ζ ∈ T. Consider now the one-parameter family of weights W [a], a > 0, defined by σj σj σj Φ−Q(ζ) W [a](ζ) def = PEj (ζ)WjPEj (ζ) + a2PEj (ζ)⊥, where PEj (ζ) and PEj (ζ)⊥ denote the orthogonal projections from Cn onto Ej(ζ) and Ej(ζ)⊥, respectively, for a.e. ζ ∈ T. If f ∈ EWj and g ∈ L ⊖Wj EWj , then HΦf⊥HΦg and so Wj + kHΦgk2 kHΦ(f + g)k2 2 = kHΦfk2 2 + kHΦgk2 2 = kfk2 2 ≤ kfk2 Wj + q2 jkgk2 Wj . Let a = σjqj > 0. It is not difficult to show that (c.f. the "matrix-weight proof" of Theorem 4.1 in [PT2]). Therefore, kfkWj = kfkW [a] and qjkgkWj ≤ kgkW [a] W [a] + kgk2 kHΦ(f + g)k2 2 ≤ kfk2 (5.5) in L with respect to the inner products as the orthogonal complements of EWj (·,·)Wj and (·,·)W [a] are equal. The inequality in (5.5) implies that the weight W [a] is k-admissible for the Hankel operator HΦ on L and so, by Theorem 5.2, there is a matrix-valued function Q# ∈ H ∞ (k)(Mn) such that W [a] = kf + gk2 W [a] Let N be the largest integer such that sN ((Φ− Q)(ζ)) = σj for a.e. ζ ∈ T. Then, for any j < N , (Φ − Q#)∗(Φ − Q#) ≤ W [a]. sj((Φ − Q#)(ζ)) ≤ sj(W [a](ζ))1/2 ≤ sj(W (ζ))1/2 = sj((Φ − Q)(ζ)) and sN −1((Φ − Q#)(ζ)) ≤ sN −1(W [a](ζ))1/2 = a < σj = sN −1((Φ − Q)(ζ)) 14 ALBERTO A. CONDORI hold for a.e. ζ ∈ T, contradicting the choice of the superoptimal approximant Q. Hence Ej (ζ) = S Φ−Q(ζ) for a.e. ζ ∈ T, as desired. σj (cid:3) Corollary 5.6. Let Φ be k-admissible, Q be the superoptimal approximant in H ∞ (k)(Mn) of Φ, and W = (Φ − Q)∗(Φ − Q). If all of the superoptimal singular values of degree k of Φ are equal, then and E∗(ζ) def = span{ f (ζ) : f ∈ E(k)(Φ)} equals Cn for a.e. ζ ∈ T. EW (Φ; ker HQ) = E(k)(Φ), (5.6) Proof. By assumption, every superoptimal singular value of degree k of Φ equals sk(HΦ) and so W (ζ) = sk(HΦ)I a.e. on T (see (2.3)), where I denotes the n × n identity matrix valued function. Thus if ξ ∈ EW (Φ; ker HQ), then ξ is a maximizing vector of HΦ−Q and so the equality in (5.6) holds by Theorem 3.2. Let σ = sk(HΦ). The statement concerning E∗(ζ) is now trivial in view of (cid:3) Theorem 5.4 and (5.6); after all, E∗(ζ) = Sσ Φ−Q(ζ) = Cn for a.e. ζ ∈ T. 6. The Index formula Let Φ be k-admissible and Q be the superoptimal approximant in H ∞ (k)(Mn) of Φ. Recall that from the remarks following Lemma 3.1, there are finite Blaschke- Potapov products B and Λ, of degree k, such that ker HQ = BH 2(Cn) and ker HQt = ΛH 2(Cn). In addition, the weight W because the operator TW −H ∗ def = (Φ − Q)∗(Φ − Q) is a k-admissible weight for HΦ ΦHΦ is non-negative on ker HQ and codim ker HQ = k. A key observation is made in the following theorem. Theorem 6.1. Suppose Φ is k-admissible and Q is the superoptimal approximant in H ∞ (k)(Mn) of Φ. If the number of non-zero superoptimal singular values of degree = Λt(Φ − Q)B is very badly (6.1) k of Φ equals n, then the matrix-valued function U approximable and def EW (Φ; ker HQ) = B ker TU . In particular, dimEW (Φ; ker HQ) ≥ n. Proof. We first establish the equality in (6.1) holds. Recall that if ξ ∈ EW (Φ; ker HQ), then (Φ − Q)ξ = HΦξ ∈ J ker HQt and ξ ∈ ker HQ. It follows that there is an f ∈ H 2(Cn) such that ξ = Bf and U f = Λt(Φ − Q)Bf = ΛtHΦξ ∈ ΛtJ ker HQt = H 2 −(Cn); that is, EW (Φ; ker HQ) ⊆ B ker TU . On the other hand, it is evident that ker TU ⊆ ker T(Φ−Q)B and EW (Φ; ker HQ) = { Bf : f ∈ H 2(Cn) and kHΦBfk2 = kBfkW } = B{ f ∈ H 2(Cn) : kH(Φ−Q)Bfk2 = k(Φ − Q)Bfk2 } = B ker T(Φ−Q)B. Therefore, B ker TU ⊆ EW (Φ; ker HQ) and thus the proof of (6.1) is complete. to see that BSσ Let σ > 0. In view of the equalities sj((Φ − Q)(ζ)) = sj(U (ζ)), j ≥ 0, it is easy U (ζ) = Sσ Φ−Q(ζ) for a.e. ζ ∈ T. AN INDEX FORMULA IN CONNECTION WITH MEROMORPHIC APPROXIMATION 15 By Theorem 5.4, there are ξ1, . . . , ξℓ ∈ EW (Φ; ker HQ) = B ker TU such that BSσ U (ζ) = Sσ Φ−Q(ζ) = span{ ξj(ζ) : 1 ≤ j ≤ ℓ } for a.e. ζ ∈ T. We deduce now from Theorem 2.3 that the matrix-valued function U is very badly approximable. By Corollary 4.3, we conclude dimEW (Φ; ker HQ) ≥ n. (cid:3) We now prove our main result. Theorem 6.2. Suppose Φ is k-admissible and Q is the superoptimal approximant in H ∞ (k)(Mn) of Φ. If the number of non-zero superoptimal singular values of degree k of Φ equals n, then the Toeplitz operator TΦ−Q is Fredholm and ind TΦ−Q = dim ker TΦ−Q = 2k + dimEW (Φ; ker HQ). (6.2) Proof. By Corollary 4.4, the Toeplitz operator TΦ−Q is Fredholm and the first equality in (6.2) holds. Therefore, the matrix-valued function U = Λt(Φ − Q)B induces a Fredholm Toeplitz operator TU with index ind TU = ind TΦ−Q − 2k. By virtue of Theorem 6.1, U is very badly approximable and so ind TU = dim ker TU = dimEW (Φ; ker HQ). The conclusion follows immediately from (6.3) and (6.4). Finally, we state two simple consequences of Theorem 6.2. (6.3) (6.4) (cid:3) Corollary 6.3. Suppose Φ is an n × n matrix-valued function, continuous on T, such that sk(HΦ) < sk−1(HΦ). Let Q be the superoptimal approximant in H ∞ (k)(Mn) of Φ. If det(Φ − Q) 6= 0 a.e. on T, then dim ker TΦ−Q ≥ 2k + n. Proof. In view of the equation in (2.3), the assumption det(Φ − Q) 6= 0 a.e. on T implies Φ has n non-zero superoptimal singular values of degree k. The conclusion now follows from Theorem 6.2. (cid:3) Corollary 6.4. Suppose Φ is k-admissible and all of the superoptimal singular values of degree k of Φ are equal. If Q is a best approximant in H ∞ (k)(Mn) to Φ, then where µ denotes the multiplicity of the singular value sk(HΦ). dim ker TΦ−Q = 2k + µ, Proof. The formula is an immediate consequence of Theorem 6.2 and Lemma 5.6. (cid:3) Acknowledgment. The author would like to thank Professor S.R. Treil for use- ful comments concerning superoptimal approximation by meromorphic functions. 16 ALBERTO A. CONDORI References [AP] R.B. Alexeev and V.V. Peller. Badly approximable matrix functions and canonical factor- [Co] izations. Indiana Univ. Math. J. 49 (2000), 1247 -- 1285. J. B. Conway. A Course in Functional Analysis. Second Edition. Graduate Texts in Math- ematics. Springer, New York, 1990. H. Helson. Lectures on Invariant Subspaces. Acad. Press, New York, 1964. [He] [Pe1] V.V. Peller. Hankel Operators and Their Applications. Springer Monographs in Mathe- matics. Springer, New York, 2003. [Pe2] V.V. Peller. Hankel operators and continuity properties of best approximation operators. Algebra i Analiz 2:1 (1990), 163-189. English transl.: Leningrad Math. J. 2:1 (1991), 139-160. [PY1] V.V. Peller and N.J. Young. Superoptimal analytic approximation of matrix functions. J. Funct. Anal. 120 (1994), 300 -- 343. [PY2] V.V. Peller and N.J. Young. Superoptimal approximation by meromorphic functions. Math. Proc. Camb. Phil. Soc. 119 (1996), 497 -- 511. [PY3] V.V. Peller and N.J. Young. Superoptimal singular values and indices of matrix functions. Int. Eq. Op. Theory. 20 (1994), 350 -- 363. [PT1] V.V. Peller and S.R. Treil. Approximation by analytic matrix functions. The four block problem. J. Funct. Anal. 148 (1997), 191 -- 228. [PT2] V.V. Peller and S.R. Treil. Very badly approximable matrix functions. Sel. math., New ser. 11 (2005), 127 -- 154. [PV] V.V. Peller and V.I. Vasyunin. Analytic approximation of rational matrix functions. Indi- [T1] [T2] ana Univ. Math. J. 56 (2007), 1913 -- 1937. S.R. Treil. The Adamyan-Arov-Krein theorem: a vector version. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 141 (1985), 56 -- 71 (Russian) S.R. Treil. On superoptimal approximation by analytic and meromorphic matrix-valued functions. J. Funct. Anal. 131 (1995), 386 -- 414. Department of Chemistry and Mathematics, Florida Gulf Coast University, 10501 FGCU Boulevard South, Fort Myers, FL 33965 E-mail address: [email protected]
1103.2874
1
1103
2011-03-15T10:14:00
Strong q-variation inequalities for analytic semigroups
[ "math.FA", "math.DS" ]
Let T : Lp --> Lp be a positive contraction, with p strictly between 1 and infinity. Assume that T is analytic, that is, there exists a constant K such that \norm{T^n-T^{n-1}} < K/n for any positive integer n. Let q strictly betweeen 2 and infinity and let v^q be the space of all complex sequences with a finite strong q-variation. We show that for any x in Lp, the sequence ([T^n(x)](\lambda))_{n\geq 0} belongs to v^q for almost every \lambda, with an estimate \norm{(T^n(x))_{n\geq 0}}_{Lp(v^q)}\leq C\norm{x}_p. If we remove the analyticity assumption, we obtain a similar estimate for the ergodic averages of T instead of the powers of T. We also obtain similar results for strongly continuous semigroups of positive contractions on Lp-spaces.
math.FA
math
STRONG q-VARIATION INEQUALITIES FOR ANALYTIC SEMIGROUPS CHRISTIAN LE MERDY, QUANHUA XU Abstract. Let T : Lp(Ω) → Lp(Ω) be a positive contraction, with 1 < p < ∞. Assume that T is analytic, that is, there exists a constant K ≥ 0 such that kT n − T n−1k ≤ K/n for any integer n ≥ 1. Let 2 < q < ∞ and let vq be the space of all complex sequences with a finite strong q-variation. We show that for any x ∈ Lp(Ω), the sequence (cid:0)[T n(x)](λ)(cid:1)n≥0 where Mn(T ) = (n+1)−1Pn belongs to vq for almost every λ ∈ Ω, with an estimate k(T n(x))n≥0kLp(vq) ≤ Ckxkp. If we remove the analyticity assumption, we obtain an estimate k(Mn(T )x)n≥0kLp(vq) ≤ Ckxkp, k=0 T k denotes the ergodic averages of T . We also obtain similar results for strongly continuous semigroups (Tt)t≥0 of positive contractions on Lp-spaces. 2000 Mathematics Subject Classification : 47A35, 37A99, 47B38 1. Introduction. Variational inequalities in probability, ergodic theory and harmonic analysis have been the subject of many recent research papers. One important character of these inequalities is the fact that they can be used to measure the speed of convergence for the family of operators in consideration. To be more precise, consider, for instance, a measure space (Ω, µ) and an operator T on L1(Ω) + L∞(Ω). Form the ergodic averages of T : Mn(T ) = 1 n + 1 nXk=0 T k , n ≥ 0. A fundamental theorem in ergodic theory states that if T is a contraction on Lp(Ω) for every 1 ≤ p ≤ ∞, then the limit limn→∞ Mn(T )x exists a.e. for every x ∈ Lp(Ω). One can naturally ask what is the speed of convergence of this limit. A classical tool for measuring that speed is the following square function 1 1 0 2 r a M 5 1 ] . A F h t a m [ 1 v 4 7 8 2 . 3 0 1 1 : v i X r a S(x) =(cid:16)Xn≥0 , n(cid:12)(cid:12)Mn+1(T )x − Mn(T )x(cid:12)(cid:12)2(cid:17) 1 2 the problem being to estimate its norm kS(x)kp. This issue goes back to Stein [31], who proved that if T as above is positive on L2(Ω) (in the Hilbertian sense), then a square function inequality kS(x)kp ≤ Cpkxkp, x ∈ Lp(Ω), Date: November 11, 2018. The authors are both supported by the research program ANR-06-BLAN-0015. 1 2 CHRISTIAN LE MERDY, QUANHUA XU holds for any 1 < p < ∞. Stein's inequality is closely related to Dunford-Schwartz's maximal ergodic inequality, (cid:13)(cid:13) sup n≥0 Mn(T )x(cid:13)(cid:13)p ≤ Cpkxkp, x ∈ Lp(Ω), 1 < p ≤ ∞. This maximal inequality and its weak type (1, 1) substitute for p = 1 are key ingredients in the proof of the previous pointwise ergodic theorem. The strong q-variation is another (better) tool to measure the speed of limn→∞ Mn(T )x. Bourgain was the first to consider variational inequalities in ergodic theory. To state his inequality we need to recall the definition of the strong q-variation. Given a sequence (an)n≥0 of complex numbers and a number 1 ≤ q < ∞, the strong q-variation norm is defined as k(an)n≥0kvq = supn(cid:0)a0q + Xk≥1 qo, ank − ank−1q(cid:1) 1 where the supremum runs over all increasing sequences (nk)k≥0 of integers such that k0 = 0. It is clear that the set vq of all sequences with a finite strong q-variation is a Banach space for the norm k kvq . Bourgain [4] proved that if T is induced by a measure preserving transformation on (Ω, µ), then for any 2 < q < ∞, (cid:13)(cid:13)(Mn(T )x)n≥0(cid:13)(cid:13)L2(vq ) ≤ Cq kxk2, x ∈ L2(Ω). This inequality was then extended to Lp(Ω) for any 1 < p < ∞ by Jones, Kaufman, Rosen- blatt and Wierdl [16]. The latter paper contains many other interesting results on the sub- ject. Note that the predecessor of Bourgain's inequality is L´epingle's variational inequality for martingales [21]. The latter says that if (En)n≥0 is an increasing sequence of conditional ex- pectations on a probability space Ω, then we have an estimate(cid:13)(cid:13)(En(x))n≥0(cid:13)(cid:13)Lp(vq) ≤ C kxkp (see also [30] for further results on that theme). Since [4], variational type inequalities have been extensively studied in ergodic theory and harmonic analysis. Many classical sequences of operators and semigroups have been proved to satisfy strong variational bounds, see in particular [9, 17, 18, 19, 26] and references therein. The main purpose of this paper is to exhibit a large class of operators T on Lp(Ω) for a fixed 1 < p < ∞ with the following property: for any 2 < q < ∞, there exists a constant C > 0 (which may depend on q and T ) such that for any x ∈ Lp(Ω), the sequence (T n(x))n≥0 belongs to Lp(Ω; vq), and We show that this holds true provided that T is a positive contraction (more generally, a contractively regular operator) and T is analytic, in the sense that (cid:13)(cid:13)(T n(x))n≥0(cid:13)(cid:13)Lp(vq) ≤ C kxkp. nkT n − T n−1k < ∞. sup n≥1 (1.1) (1.2) Inequality (1.1) implies, of course, a similar variational inequality for the ergodic averages Mn(T ). However, in this latter case, the analytic assumption above can be removed. Namely, for a positive contraction T on Lp(Ω) with 1 < p < ∞ we have an estimate (cid:13)(cid:13)(Mn(T )x)n≥0(cid:13)(cid:13)Lp(vq) ≤ C kxkp, x ∈ Lp(Ω), 3 for any 2 < q < ∞. This result extends those of [4] and [16] quoted previously. Note that inequality (1.1) for positive analytic contractions considerably improves our previous maximal ergodic inequality for such operators T proved in [20]. In this sense, this paper is a continuation of [20]. On the other hand, our proof of (1.1) heavily relies on the square function inequality of [20]. We also establish results similar to (1.1) and (1.2) for strongly continuous semigroups. This requires the following continuous analog of vq. Given a complex family (at)t>0, define k(at)t>0kV q = supn(cid:0)at0q + Xk≥1 qo, atk − atk−1q(cid:1) 1 where the supremum runs over all increasing sequences (tk)k≥0 of positive real numbers. Then we let V q be the resulting Banach space of all (at)t>0 such that k(at)t>0kV q < ∞ . Consider a bounded analytic semigroup (Tt)t≥0 on Lp(Ω), with 1 < p < ∞. We show that if Tt is a positive contraction (more generally, a contractively regular operator) for any t ≥ 0, then for any 2 < q < ∞ and any x ∈ Lp(Ω), the family (cid:0)[Tt(x)](λ)(cid:1)t>0 belongs to V q for almost every λ ∈ Ω, and we have an estimate (1.3) ≤ C kxkp, x ∈ Lp(Ω). (cid:13)(cid:13)(cid:13)λ 7→(cid:13)(cid:13)(cid:0)[Tt(x)](λ)(cid:1)t>0(cid:13)(cid:13)V q(cid:13)(cid:13)(cid:13)p We mention that Jones and Reinhold [17] proved (1.3) for positive, unital symmetric diffusion semigroups and (1.1) for a certain class of convolution operators (only in the case p = 2). Our results turn out to extend these contributions in various directions. As in the discrete case, we obtain similar results for the averages of the semigroup if we remove the analyticity assumption. Inequalities for ergodic averages, such as (1.2), will be established in Section 3. Then our main results leading to (1.1) and (1.3) will be established in Section 4, using the above mentioned results from Section 3, as well as square function estimates from [20]. Section 5 is devoted to various complements. On the one hand, we establish individual (= pointwise) ergodic theorems in our context. For example, if T : Lp(Ω) → Lp(Ω) is a positive analytic contraction, then limn→∞ T n(x) exists a.e. for every x ∈ Lp(Ω). On the other hand, it is well-known that (1.2) cannot be extended to the case q = 2. Then we show analogs of (1.2) and (1.1) when vq is replaced by the oscillation space o2. We give examples and applications in Section 6. We end this introduction with a few notation. If X is a Banach space, we let B(X) denote the algebra of all bounded operators on X and we let IX denote the identity operator on X (or simply I if there is no ambiguity on X). For any T ∈ B(X), we let σ(T ) denote the spectrum of T . Also we let D = {z ∈ C : z < 1} be the open unit disc of C. For a measurable function x : Ω → C acting on a measure space (Ω, µ), kxkp denotes the Lp-norm of x. We refer to [27] and [15] for background on strongly continuous and analytic semigroups on Banach space. 4 CHRISTIAN LE MERDY, QUANHUA XU 2. Preliminaries on q-variation. The aim of this section is to provide some elementary background on the spaces vq and V q defined above. We fix some 1 ≤ q < ∞. Let (an)n≥0 be an element of vq. Then the sequence (an)n≥0 is bounded, with k(an)n≥0k∞ ≤ 2k(an)n≥0kvq , and lim m→∞(cid:13)(cid:13)(0, . . . , 0, am+1 − am, . . . , an − am, . . .)(cid:13)(cid:13)vq = 0. Thus the space of eventually constant sequences is dense in vq. Consequently any element of vq is a converging sequence. For any integer m ≥ 1, let vq m be the space of (m + 1)-tuples (a0, a1, . . . , am) of complex numbers, equipped with the norm It follows from above that an infinite sequence (an)n≥0 belongs to vq if and only if there is a ≤ C for any m ≥ 1 and in this case, (cid:13)(cid:13)(a0, a1, . . . , am)(cid:13)(cid:13)vq m constant C ≥ 0 such that (cid:13)(cid:13)(a0, a1, . . . , am)(cid:13)(cid:13)vq k(an)n≥0kvq = lim (2.1) = (cid:13)(cid:13)(a0, a1, . . . , am, am, . . .)(cid:13)(cid:13)vq . m→∞(cid:13)(cid:13)(a0, a1, . . . , am)(cid:13)(cid:13)vq . m m Let (Ω, µ) be a measure space, let 1 < p < ∞ and let Lp(Ω; vq) denote the corresponding Bochner space. Any element of that space can be naturally regarded as a sequence of Lp(Ω). The following is a direct consequence of the above approximation properties. Lemma 2.1. Let (xn)n≥0 be a sequence of Lp(Ω), the following assertions are equivalent. (i) The sequence (xn)n≥0 belongs to Lp(Ω; vq). (ii) The sequence (xn(λ))n≥0 belongs to vq for almost every λ ∈ Ω and the function λ 7→ k(xn(λ))n≥0kvq belongs to Lp(Ω). (iii) There is a constant C ≥ 0 such that for any m ≥ 1. (cid:13)(cid:13)(x0, x1, . . . , xm)(cid:13)(cid:13)Lp(vq m) ≤ C In this case, (cid:13)(cid:13)λ 7→ k(xn(λ))n≥0kvq(cid:13)(cid:13)p = k(xn)n≥0kLp(vq) = lim m→∞(cid:13)(cid:13)(x0, x1, . . . , xm)(cid:13)(cid:13)Lp(vq . m) We now consider the continuous case. We note for further use that any element (at)t>0 of V q admits limits lim t→0+ at and lim t→∞ at. The following is a continuous analog of Lemma 2.1. Note however that families satisfying the next statement do not necessarily belong to the Bochner space Lp(Ω; V q). Lemma 2.2. Let (xt)t>0 be a family of Lp(Ω) and assume that: (1) For a.e. λ ∈ Ω, the function t 7→ xt(λ) is continuous on (0, ∞). (2) There exists a constant C ≥ 0 such that whenever t0 < t1 < · · · < tm is a finite 5 increasing sequence of positive real numbers, we have Then (xt(λ))t>0 belongs to V q for a.e. λ ∈ Ω, the function λ 7→ (cid:13)(cid:13)(xt(λ))t>0(cid:13)(cid:13)V q belongs to Lp(Ω) and m) ≤ C. (cid:13)(cid:13)(xt0 , xt1, . . . , xtm)(cid:13)(cid:13)Lp(Ω;vq (cid:13)(cid:13)λ 7→ k(xt(λ))t>0kV q(cid:13)(cid:13)p ≤ C. k≥1(cid:13)(cid:13)(cid:0)xn2−N (λ)(cid:1)n≥k(cid:13)(cid:13)vq , ϕN (λ) = sup λ ∈ Ω. Proof. For any integer N ≥ 1, define It follows from (2.1) that ϕN is measurable. Moreover the sequence (ϕN )N ≥1 is nondecreasing, and we may therefore define ϕ(λ) = lim N→∞ ϕN (λ), λ ∈ Ω. By construction, ϕ is measurable and by the monotone convergence theorem, its Lp-norm is equal to limN kϕN kp. According to the assumption (2) and the approximation property (2.1), the Lp-norm of ϕN is ≤ C for any N ≥ 1. Hence ZΩ ϕ(λ)p dµ(λ) ≤ C p. This implies that ϕ(λ) < ∞ for a.e. λ ∈ Ω. If λ ∈ Ω is such that t 7→ xt(λ) is continuous on (0, ∞), then ϕ(λ) =(cid:13)(cid:13)(xt(λ))t>0(cid:13)(cid:13)V q . According to the assumption (1), this holds true almost everywhere. Hence the quantity (cid:13)(cid:13)(xt(λ))t>0(cid:13)(cid:13)V q is finite for a.e. λ ∈ Ω. The lemma clearly follows from these properties. (cid:3) 3. Variation of ergodic averages. Throughout we let (Ω, µ) be a measure space and we let 1 < p < ∞. We first recall the notion of regular operators on Lp(Ω) and some of their basic properties which will be used in this paper. We refer e.g. to [23, Chap. 1] and to [28, 29] for more details and complements. An operator T : Lp(Ω) → Lp(Ω) is called regular if there is a constant C ≥ 0 such that (cid:13)(cid:13)sup k≥1 T (xk)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)sup k≥1 xk(cid:13)(cid:13)p for any finite sequence (xk)k≥1 in Lp(Ω). Then we let kT kr denote the smallest C for which this holds. The set of all regular operators on Lp(Ω) is a vector space on which k kr is a norm. We say that T is contractively regular if kT kr ≤ 1. Let E be a Banach space. If an operator T : Lp(Ω) → Lp(Ω) is regular, then the operator T ⊗ IE : Lp(Ω) ⊗ E → Lp(Ω) ⊗ E extends to a bounded operator on the Bochner space Lp(Ω; E), and (3.1) (cid:13)(cid:13)T ⊗ IE : Lp(Ω; E) −→ Lp(Ω; E)(cid:13)(cid:13) ≤ kT kr. 6 CHRISTIAN LE MERDY, QUANHUA XU Indeed by definition this holds true when E = ℓ∞ n for any n ≥ 1, and the general case follows from the fact that for any ε > 0, any finite dimensional Banach space is (1 + ε)-isomorphic to a subspace of ℓ∞ n for some large enough n ≥ 1. Any positive operator T (in the lattice sense) is regular and kT kr = kT k in this case. Thus all statements given for contractively regular operators apply to positive contractions. It is well-known that conversely, T is regular with kT kr ≤ C if and only if there is a positive operator S : Lp(Ω) → Lp(Ω) with kSk ≤ C, such that T (x) ≤ S(x) for any x ∈ Lp(Ω). Finally, following [28], we say that an operator T : L1(Ω) + L∞(Ω) → L1(Ω) + L∞(Ω) is an absolute contraction if it induces two contractions T : L1(Ω) −→ L1(Ω) and T : L∞(Ω) −→ L∞(Ω). Then the resulting operator T : Lp(Ω) → Lp(Ω) is contractively regular. The main result of this section is the following theorem, which might be known to experts. Its proof relies on the transference principle, already used in [4]. Theorem 3.1. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator, with 1 < p < ∞, and let 2 < q < ∞. Then we have for some constant Cp,q only depending on p and q. (cid:13)(cid:13)(cid:0)Mn(T )x(cid:1)n≥0(cid:13)(cid:13)Lp(vq ) ≤ Cp,q kxkp, x ∈ Lp(Ω), sp : ℓp Z −→ ℓp Z, denote the shift operator on ℓp Z. According to [16, Thm. B], sp satisfies Theorem 3.1 (the crucial case p = 2 going back to [4]). Thus for any 1 < p < ∞ and 2 < q < ∞ we have a constant sp(cid:0)(cn)n(cid:1) = (cn−1)n, Let (3.2) Cp,q = (cid:13)(cid:13)c 7→(cid:0)Mn(sp)c(cid:1)n≥0(cid:13)(cid:13)ℓp Z→ℓp Z(vq ) . The following lemma is a variant of the well-known Coifman-Weiss transference Theorem [7, Thm. 2.4] and is closely related to [10] and [2]. Lemma 3.2. Let U : Lp(Ω) → Lp(Ω) be an invertible operator such that U j is regular for any j ∈ Z and suppose that C = sup{kU jkr : j ∈ Z} < ∞ . Let (e1, . . . , eN ) be a basis of a finite dimensional Banach space E, and let a(1), . . . , a(N) be N elements of ℓ1 Z. Consider the two operators K : ℓp Z −→ ℓp Z(E), K(c) = and Then R : Lp(Ω) −→ Lp(Ω; E), a(k)j sj p(c)(cid:17) ⊗ ek, a(k)j U j(x)(cid:17) ⊗ ek. NXk=1(cid:16)Xj∈Z NXk=1(cid:16)Xj∈Z R(x) = kRk ≤ C 2kKk. Proof. The operator ILp(Ω) ⊗ K extends to a bounded operator Lp(cid:0)Ω; ℓp whose norm is equal to kKk. (Nothing special about K is required for this tensor extension property.) By Fubini's Theorem, Z(cid:1) → Lp(cid:0)Ω; ℓp Z(E)(cid:1), 7 Lp(cid:0)Ω; ℓp Z(cid:1) ≃ ℓp Z(cid:0)Lp(Ω)(cid:1) and Lp(cid:0)Ω; ℓp Z(E)(cid:1) ≃ ℓp Z(cid:0)Lp(Ω; E)(cid:1) isometrically. Further, under these identifications, the extension of ILp(Ω) ⊗ K corresponds to the operator Z(cid:0)Lp(Ω; E)(cid:1), NXk=1(cid:16)Xj a(k)j(cid:0)sj p⊗ILp(Ω)(cid:1)(z)(cid:17) ⊗ ek eK(z) = eK : ℓp Z(cid:0)Lp(Ω)(cid:1) −→ ℓp Thus we have keKk = kKk. By approximation, we may suppose that a(1), . . . , a(N) are finitely supported. Let m ≥ 1 be chosen such that a(k)j = 0 for any k and any j > m. Let x ∈ Lp(Ω) and let yk = mXj=−m a(k)j U j(x) for any k = 1, . . . , N. Our aim is to estimate the norm of Pk yk ⊗ ek in Lp(Ω; E). For any i ∈ Z, we have NXk=1 yk ⊗ ek = (U i ⊗ IE)(cid:16) NXk=1 U −iyk ⊗ ek(cid:17). Hence applying (3.1) to U i, we derive that (cid:13)(cid:13)(cid:13) NXk=1 yk ⊗ ek(cid:13)(cid:13)(cid:13)Lp(Ω;E) ≤ C(cid:13)(cid:13)(cid:13) NXk=1 U −iyk ⊗ ek(cid:13)(cid:13)(cid:13)Lp(Ω;E) Let n ≥ 1 be an arbitrary integer and let χ be the characteristic function of the interval [−(n + m), (n + m)]. We deduce from the above estimate that (2n + 1)(cid:13)(cid:13)(cid:13) NXk=1 Since Xi (cid:13)(cid:13)(cid:13) NXk=1Xj p p Lp(Ω;E) Lp(Ω;E) ≤ C p yk ⊗ ek(cid:13)(cid:13)(cid:13) nXi=−n(cid:13)(cid:13)(cid:13) a(k)j U j−i(x) ⊗ ek(cid:13)(cid:13)(cid:13) NXk=1 mXj=−m a(k)j χ(j − i) U j−i(x) ⊗ ek(cid:13)(cid:13)(cid:13) ≤ C p Xi (cid:13)(cid:13)(cid:13) NXk=1Xj a(k)j χ(j − i) U j−i(x) ⊗ ek(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)eKh(cid:0)χ(−j)U −j(x)(cid:1)ji(cid:13)(cid:13)(cid:13) Lp(Ω;E) p p p . Lp(Ω;E) ℓp Z(Lp(Ω;E)) 8 CHRISTIAN LE MERDY, QUANHUA XU and keKk = kKk, this yields (2n + 1)(cid:13)(cid:13)(cid:13) NXk=1 yk ⊗ ek(cid:13)(cid:13)(cid:13) p Lp(Ω;E) Letting n → ∞, we get the result. n+mXj=−(n+m) ≤ C pkKkp kU −j(x)kp p ≤ (cid:0)2(n + m) + 1(cid:1) C 2pkKkpkxkp p. (cid:3) Proof of Theorem 3.1. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator. There exists another measure space (bΩ,bµ), two positive contractions J : Lp(Ω) → Lp(bΩ) and Q : Lp(bΩ) → Lp(Ω) and an isometric invertible operator U : Lp(bΩ) → Lp(bΩ) such that T k = QU kJ, k ≥ 0. In the case when T is positive, this is Akcoglu's famous dilation Theorem (see [1]). The extension to regular operators stated here is from [28] or [6]. Moreover U can be chosen so that U and U −1 are both contractively regular. Thus ∀ j ∈ Z, kU jkr = 1. Note that when 1 < p 6= 2 < ∞, any isometry on Lp is contractively regular, so the latter information is relevant only when p = 2. We fix an integer m ≥ 1 and we consider vq m as defined in Section 2. For any n ≥ 0, we clearly have Mn(T ) = QMn(U)J. Since kQkr ≤ 1, it follows from (3.1) that (3.3) (cid:13)(cid:13)(cid:0)Mn(T )x(cid:1)0≤n≤m(cid:13)(cid:13)Lp(Ω;vq m) ≤ (cid:13)(cid:13)(cid:0)Mn(U)J(x)(cid:1)0≤n≤m(cid:13)(cid:13)Lp( bΩ;vq For any n = 0, 1, . . . , m, let a(n) ∈ ℓ1 Z be defined by letting a(n)j = (n + 1)−1 if 0 ≤ j ≤ n , x ∈ Lp(Ω). m) and a(n)j = 0 otherwise. Then Xj∈Z a(n)j U j = Mn(U) and Xj∈Z a(n)j sj p = Mn(sp). Applying Lemma 3.2 with E = vq m and recalling (3.2), we therefore deduce that m) ≤ Cp,q kzkp (cid:13)(cid:13)(cid:0)Mn(U)z(cid:1)0≤n≤m(cid:13)(cid:13)Lp( bΩ;vq (cid:13)(cid:13)(cid:0)Mn(T )x(cid:1)0≤n≤m(cid:13)(cid:13)Lp(Ω;vq for any z ∈ Lp(bΩ). Combining with the inequality (3.3) we obtain that m) ≤ Cp,q kxkp for any x ∈ Lp(Ω). Then the result follows from Lemma 2.1. (cid:3) Remark 3.3. The above Lemma 3.2 extends without any difficulty to amenable groups, as follows. Let G be a locally compact amenable group, with left Haar measure dt, let π : G → B(Lp(Ω)) be a strongly continuous representation valued in the space of regular operators on Lp(Ω), and assume that C = sup{kπ(t)kr : t ∈ G} < ∞ . Next let h1, . . . , hN be N elements of L1(G) and let K : Lp(G) → Lp(G; E) be defined by letting K(f ) =Pk(hk ∗ f ) ⊗ ek for any f ∈ Lp(G). Then for any x ∈ Lp(Ω), we have (cid:13)(cid:13)(cid:13)Xk (cid:16)ZG hk(t)π(t)x dt(cid:17) ⊗ ek(cid:13)(cid:13)(cid:13)Lp(Ω;E) ≤ C 2 kKk kxkp. 9 We conclude this section with a continuous version of Theorem 3.1. Given a strongly continuous semigroup T = (Tt)t≥0 on Lp(Ω), we let Mt(T ) = defined in the strong sense. 1 t Z t 0 Ts ds , t > 0, Corollary 3.4. Let T = (Tt)t≥0 be a strongly continuous semigroup on Lp(Ω) and assume that Tt : Lp(Ω) → Lp(Ω) is contractively regular for any t ≥ 0. Let 2 < q < ∞ and let x ∈ Lp(Ω). Then for a.e. λ ∈ Ω, the family (cid:0)[Mt(T )x](λ)(cid:1)t>0 belongs to V q and (cid:13)(cid:13)(cid:13)λ 7→(cid:13)(cid:13)(cid:0)[Mt(T )x](λ)(cid:1)t>0(cid:13)(cid:13)V q(cid:13)(cid:13)(cid:13)p ≤ Cp,q kxkp. Proof. Consider x ∈ Lp(Ω). According to [12, Section VIII.7], the function t 7→ [Mt(T )x](λ) is continuous for a.e. λ ∈ Ω. Let t0 < t1 < · · · < tm be positive real numbers and let ε > 0. It follows from the strong continuity of T = (Tt)t≥0 that there exist α > 0 and integers n0, n1, . . . , nm such that ∀ k = 0, . . . , m, Hence applying Theorem 3.1 and a limit argument, we deduce that (3.4) (cid:13)(cid:13)(cid:0)Mt0(T )x, Mt1(T )x, . . . , Mtm(T )x(cid:1)(cid:13)(cid:13)Lp(Ω;vq The result therefore follows from Lemma 2.2. (cid:13)(cid:13)Mtk (T )x − Mnk(Tα)x(cid:13)(cid:13)p < ε. m) ≤ Cpqkxkp. (cid:3) An alternative proof of (3.4) consists in using Fendler's dilation Theorem for semigroups (see [13]), and then arguing as in the proof Theorem 3.1. This only requires knowing that the result of Corollary 3.4 holds true for the translation group on Lp(R), which follows from [4, 16, 17], and using Remark 3.3 for G = R to transfer that result to strongly continuous groups of contractively regular isometries. 4. The analytic case. Let X be an arbitrary Banach space, and let (Tt)t≥0 be a strongly continuous semigroup on X. We call it a bounded analytic semigroup if there exists a positive angle ω ∈ (cid:0)0, π 2(cid:1) and a bounded analytic family z ∈ Σω 7→ Tz ∈ B(X) extending (Tt)t>0, where Σω = (cid:8)z ∈ C∗ : Arg(z) < ω(cid:9) is the open sector of angle 2ω around (0, ∞). We refer to [15, 27] for various characterizations and properties of bounded analytic semigroups. We simply recall that if we let A denote the infinitesimal generator of (Tt)t≥0, then the latter is a bounded analytic semigroup if and only 10 CHRISTIAN LE MERDY, QUANHUA XU if Tt maps X into the domain of A for any t > 0 and there exist two constants C0, C1 > 0 such that (4.1) ∀ t > 0, kTtk ≤ C0 and ktATtk ≤ C1. The definition of analyticity for discrete semigroups parallels (4.1). Let T ∈ B(X). We say that T is power bounded if and that it is analytic if moreover, kT nk < ∞ sup n≥0 nkT n − T n−1k < ∞. sup n≥1 This notion goes back to [8] and has been studied in various contexts so far. We gather here a few spectral properties of these operators and refer to [3, 22, 24, 25] for proofs and complements. The most important result is the following: an operator T is power bounded and analytic if and only if (4.2) σ(T ) ⊂ D and (cid:8)(z − 1)(zI − T )−1 : z > 1(cid:9) is bounded. This property is called the 'Ritt condition'. The key argument for this characterization is due to O. Nevanlinna [25]. For any angle γ ∈(cid:0)0, π 2(cid:1), let Bγ be the convex hull of 1 and the closed disc D(0, sin γ). γBγ 0 1 Then (4.2) implies that Figure 1. (4.3) (4.4) Furthermore, (4.3) is equivalent to π 2(cid:1) (cid:12)(cid:12) ∃ γ ∈(cid:0)0, ∃ K > 0 (cid:12)(cid:12) ∀ z ∈ σ(T ), σ(T ) ⊂ Bγ. 1 − z ≤ K(cid:0)1 − z(cid:1). The aim of this section is to show that under an analyticity assumption, the ergodic averages can be replaced by the semigroup itself in either Theorem 3.1 (discrete case) or Corollary 3.4 (continuous case). As in Section 3, we consider a measure space (Ω, µ) and a number 1 < p < ∞. We will consider an operator T : Lp(Ω) → Lp(Ω) and we let 11 for any integers n, m ≥ 0. Note that (∆m ∆m n = T n(T − I)m n )n≥0 is the m-difference sequence of (T n)n≥0. We will need the following Littlewood-Paley type inequalities which were estabished in [20]. Proposition 4.1. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator, with 1 < p < ∞, and assume that T is analytic. Then for any integer m ≥ 0, there is a constant Cm > 0 such that (cid:13)(cid:13)(cid:13)(cid:16) ∞Xn=0 (n + 1)2m+1(cid:12)(cid:12)∆m+1 n 2(cid:13)(cid:13)(cid:13)p (x)(cid:12)(cid:12)2(cid:17) 1 ≤ Cm kxkp, x ∈ Lp(Ω). We will also use the following elementary estimates, whose proofs are left to the reader. Lemma 4.2. For any integer m ≥ 0, there exists a constant Km such that for any n ≥ 1, (cid:16) 2nXj=n (j + 1)1−2m(cid:17) 1 2 ≤ Kmn−m+1. Lemma 4.3. For any sequences (δn)n≥0 ∈ v1 and (zn)n≥0 ∈ Lp(Ω; vq), we have (δnzn)n≥0 ∈ Lp(Ω; vq) and (cid:13)(cid:13)(δnzn)n≥0(cid:13)(cid:13)Lp(vq ) ≤ 3(cid:13)(cid:13)(δn)n≥0(cid:13)(cid:13)v1(cid:13)(cid:13)(zn)n≥0(cid:13)(cid:13)Lp(vq ) . In the next statements and their proofs, . will stand for an inequality up to a constant which may depend on T , q and m, but not on x. Theorem 4.4. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator, with 1 < p < ∞, and assume that T is analytic. Then for any 2 < q < ∞, we have an estimate More generally, for any integer m ≥ 0, we have an estimate (4.5) (4.6) Proof. It will be convenient to set (cid:13)(cid:13)(cid:0)T n(x)(cid:1)n≥0(cid:13)(cid:13)Lp(vq) . kxkp, n (x)(cid:1)n≥1(cid:13)(cid:13)Lp(vq ) . kxkp, (cid:13)(cid:13)(cid:0)nm∆m n−1Xj=0 n = nMn−1(T ) = ∆−1 T j , x ∈ Lp(Ω). x ∈ Lp(Ω). n ≥ 1. Then for any n < N and any m ≥ −1, we have (4.7) ∆m N − ∆m n = N −1Xj=n ∆m+1 j . With the above notation, (4.6) holds true for m = −1, by Theorem 3.1. 12 CHRISTIAN LE MERDY, QUANHUA XU We will proceed by induction. We fix an integer m ≥ 0 and assume that (4.6) holds true for (m − 1). Thus using Lemma 4.3, we both have (4.8) and (4.9) (cid:13)(cid:13)(cid:0)nm−1∆m−1 (cid:13)(cid:13)(cid:0)nm−1∆m−1 n+1 (x)(cid:1)n≥1(cid:13)(cid:13)Lp(vq ) . kxkp 2n+1(x)(cid:1)n≥1(cid:13)(cid:13)Lp(vq) . kxkp. Next for any n ≥ 1, we write 2nXj=n (j + 1)∆m+1 j = = 2nXj=n 2n+1Xj=n+1 2nXj=n+1 (j + 1)(∆m j+1 − ∆m j ) j∆m j − 2nXj=n (j + 1)∆m j = − ∆m j + (2n + 1)∆m 2n+1 − (n + 1)∆m n = n∆m 2n+1 + (n + 1)(∆m 2n+1 − ∆m n ) + ∆m−1 n+1 − ∆m−1 2n+1, using (4.7) in due places. Hence (4.10) nm∆m 2n+1 = nm−1 2nXj=n (j + 1)∆m+1 j − nm−1(n + 1)(∆m 2n+1 − ∆m n ) + nm−1∆m−1 2n+1 − nm−1∆m−1 n+1 . This identity suggests the introduction of the following two sequences of operators. For any n ≥ 1, we set An = nm−1 2nXj=n (j + 1)∆m+1 j and Also for any x ∈ Lp(Ω), we set Bn = nm(∆m 2n+1 − ∆m n ). Φm(x) = (cid:16) ∞Xj=1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 j 2 (x)2(cid:17) 1 . According to Proposition 4.1, this function is an element of Lp(Ω). Let (nk)k≥0 be an increasing sequence of integers, with n0 = 1. For any k ≥ 1, we set 13 (j + 1)∆m+1 j nm−1 k nm−1 k (j + 1)∆m+1 j 2nkXj=2nk−1+1 2nkXj=nk nk−1Xj=nk−1 2nk−1Xj=nk−1 k−1(cid:17) 2nk−1Xj=nk −nm−1 k−1 (j + 1)∆m+1 j −nm−1 k−1 (j + 1)∆m+1 j ak =     ck =  bk = if 2nk−1 ≥ nk if 2nk−1 < nk, if 2nk−1 ≥ nk if 2nk−1 < nk, (cid:16)nm−1 k − nm−1 (j + 1)∆m+1 j if 2nk−1 ≥ nk 0 if 2nk−1 < nk. This yields a decomposition (4.11) Ank − Ank−1 = ak + bk + ck. Let x ∈ Lp(Ω). If 2nk−1 ≥ nk, we have, using Cauchy-Schwarz, (cid:12)(cid:12)ak(x)(cid:12)(cid:12) ≤ nm−1 k ≤ nm−1 k ≤ nm−1 k 2 (x)(cid:12)(cid:12)2(cid:17) 1 2 . 2nkXj=2nk−1+1 (cid:16) 2nkXj=2nk−1+1 (cid:16) 2nkXj=nk j j (x)(cid:12)(cid:12) (j + 1)(cid:12)(cid:12)∆m+1 2(cid:16) (j + 1)1−2m(cid:17) 1 2nkXj=2nk−1+1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 (j + 1)1−2m(cid:17) 1 (x)(cid:12)(cid:12)2(cid:17) 1 2(cid:16) 2nkXj=2nk−1+1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 (x)(cid:12)(cid:12)2(cid:17) 1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 2 (cid:16) 2nkXj=nk (j + 1)1−2m(cid:17) 1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 (j + 1)1−2m(cid:17) 1 2 (cid:16) 2nkXj=2nk−1+1 j j j 2 2 (x)(cid:12)(cid:12)2(cid:17) 1 . Similarly if 2nk−1 < nk, we have (cid:12)(cid:12)ak(x)(cid:12)(cid:12) ≤ nm−1 k ≤ nm−1 k (cid:16) 2nkXj=nk (cid:16) 2nkXj=nk Hence in both cases, we have (cid:12)(cid:12)ak(x)(cid:12)(cid:12)2 ≤ K 2 m 2nkXj=2nk−1+1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 j , (x)(cid:12)(cid:12)2 14 CHRISTIAN LE MERDY, QUANHUA XU by Lemma 4.2. Summing up, we deduce that (4.12) Likewise we have (4.13) ≤ K 2 ∞Xk=1(cid:12)(cid:12)ak(x)(cid:12)(cid:12)2 ∞Xk=1(cid:12)(cid:12)bk(x)(cid:12)(cid:12)2 ≤ K 2 m Φm(x)2. m Φm(x)2. We now turn to ck(x). Assume that 2nk−1 ≥ nk. Then using again Cauchy-Schwarz and Lemma 4.2, we have (cid:12)(cid:12)ck(x)(cid:12)(cid:12) ≤ (cid:12)(cid:12)nm−1 (cid:12)(cid:12)ck(x)(cid:12)(cid:12)2 k−1(cid:12)(cid:12)(cid:16)2nk−1Xj=nk m(cid:18) nm−1 (j + 1)1−2m(cid:17) 1 (cid:19)2 2nk−1Xj=nk k−1 2(cid:16)2nk−1Xj=nk j (j + 1)2m+1(cid:12)(cid:12)∆m+1 (x)(cid:12)(cid:12)2 (j + 1)2m+1(cid:12)(cid:12)∆m+1 j . k − nm−1 nm−1 k ≤ K 2 2 (x)(cid:12)(cid:12)2(cid:17) 1 , k − nm−1 hence For any integer j ≥ 1, define Jj = {k ≥ 1 : nk ≤ j ≤ 2nk−1}, and set Λj = Xk∈Jj (cid:18) nm−1 nm−1 k k − nm−1 k−1 (cid:19)2 . Then it follows from the above calculation that ∞Xk=1(cid:12)(cid:12)ck(x)(cid:12)(cid:12)2 ≤ K 2 m ∞Xj=1 Λj (j + 1)2m+1(cid:12)(cid:12)∆m+1 j (x)(cid:12)(cid:12)2 . Let us now estimate the Λj's. Observe that if Jj is a non empty set, then it is a finite interval of integers. Thus it reads as Jj = {kj − N + 1, kj − N + 2, . . . , kj − 1, kj}, where kj is the biggest element of Jj and N is its cardinal. Suppose that m ≥ 2, so that the sequence (nm−1 )k is increasing. Then k Xk∈Jj nm−1 k − nm−1 k−1 = N −1Xr=0 nm−1 kj −r − nm−1 kj −r−1 = nm−1 kj Since kj ∈ Jj, we have nkj ≤ j, hence Xk∈Jj nm−1 k − nm−1 k−1 ≤ jm−1. − nm−1 kj −N ≤ nm−1 kj . On the other hand, we have j ≤ 2nk for any k ∈ Jj, hence Xk∈Jj nm−1 k − nm−1 k−1 nm−1 k ≤ (cid:16)2 j(cid:17)m−1 Xk∈Jj nm−1 k − nm−1 k−1 ≤ 2m−1. We immediatly deduce that Λj ≤ 4m−1. In the case when m = 0, we have similarly 15 Xk∈Jj n−1 k−1 − n−1 k n−1 k ≤ j Xk∈Jj n−1 k−1 − n−1 k ≤ j n−1 kj −N ≤ 2. Hence Λj ≤ 4 in this case. Lastly, it is plain that if m = 1, we have Λj = 0. This shows that in all cases, we have an estimate Now recall (4.11). Combining the above estimate with (4.12) and (4.13), we obtain that 2 Φm(x)2. ≤ K ′ m ∞Xk=1(cid:12)(cid:12)ck(x)(cid:12)(cid:12)2 ≤ 3(cid:16) ∞Xk=1(cid:12)(cid:12)ak(x)(cid:12)(cid:12)2 ≤(cid:0)6K 2 m + K ′ + ∞Xk=1(cid:12)(cid:12)bk(x)(cid:12)(cid:12)2 2(cid:1) Φm(x)2. m + ∞Xk=1(cid:12)(cid:12)ck(x)(cid:12)(cid:12)2(cid:17) ∞Xk=1(cid:12)(cid:12)Ank(x) − Ank−1(x)(cid:12)(cid:12)2 Since the upper bound does not depend on the sequence (nk)k≥0, this estimate and Propo- sition 4.1 imply that the sequence (An(x))n≥0 belongs to Lp(Ω; v2), and that we have an estimate (4.14) x ∈ Lp(Ω). We will now apply a similar treatment to the sequence (Bn)n. According to (4.7), we can (cid:13)(cid:13)(cid:0)An(x)(cid:1)n≥1(cid:13)(cid:13)Lp(v2) . kxkp, 2nXj=n Bn = nm j ∆m+1 . write For any k ≥ 1, we set αk = nm k βk = −nm k−1 γk =(cid:16)nm , j ∆m+1 2nkXj=2nk−1+1 nk−1Xj=nk−1 k−1(cid:17) 2nk−1Xj=nk ∆m+1 j k − nm , ∆m+1 j 16 CHRISTIAN LE MERDY, QUANHUA XU if 2nk−1 ≥ nk, and αk = nm k , j ∆m+1 2nkXj=nk 2nk−1Xj=nk−1 βk = −nm k−1 ∆m+1 j , if 2nk−1 < nk. Arguing as above, we obtain that for any x ∈ Lp(Ω), we have γk = 0 (cid:12)(cid:12)αk(x)(cid:12)(cid:12) ≤ nm k(cid:16) 2nkXj=nk and then Likewise we have as well as an estimate 2(cid:16) (j + 1)−1−2m(cid:17) 1 ∞Xk=1(cid:12)(cid:12)αk(x)(cid:12)(cid:12)2 ∞Xk=1(cid:12)(cid:12)βk(x)(cid:12)(cid:12)2 ∞Xk=1(cid:12)(cid:12)γk(x)(cid:12)(cid:12)2 ≤ K 2 m+1 Φm(x)2. ≤ K 2 m+1 Φm(x)2, ≤ K ′2 m+1 Φm(x)2. 2nkXj=2nk−1+1 (j + 1)2m+1(cid:12)(cid:12)∆m+1 j 2 (x)(cid:12)(cid:12)2(cid:17) 1 , (4.15) These three inequalities imply that the sequence (Bn(x))n≥0 belongs to Lp(Ω; v2), and that we have an estimate We can now conclude our proof. Recall that q > 2, so that v2 ⊂ vq. Then using (4.8), (4.9), (4.14), (4.15) and Lemma 4.3, it follows from the decomposition formula (4.10) that x ∈ Lp(Ω). (cid:13)(cid:13)(cid:0)Bn(x)(cid:1)n≥1(cid:13)(cid:13)Lp(v2) . kxkp, 2n+1(x)(cid:1)n≥1 belongs to Lp(Ω; vq) and that we have an estimate (cid:13)(cid:13)(cid:0)nm∆m 2n+1(x)(cid:1)n≥1(cid:13)(cid:13)Lp(vq) . kxkp. Since n∆m n = nm∆m 2n+1 − Bn, a second application of (4.15) yields (4.6). (cid:3) for any x ∈ Lp(Ω), (cid:0)nm∆m The following is an analog of Theorem 4.4 for continuous semigroups. Corollary 4.5. Let (Tt)t≥0 be a bounded analytic semigroup on Lp(Ω) and assume that Tt : Lp(Ω) → Lp(Ω) is contractively regular for any t ≥ 0. Let 2 < q < ∞ and let x ∈ Lp(Ω). Then for a.e. λ ∈ Ω, the family (cid:0)[Tt(x)](λ)(cid:1)t>0 belongs to V q and we have an estimate (4.16) (cid:13)(cid:13)(cid:13)λ 7→(cid:13)(cid:13)(cid:0)[Tt(x)](λ)(cid:1)t>0(cid:13)(cid:13)V q(cid:13)(cid:13)(cid:13)p . kxkp. More generally, for any integer m ≥ 0, the family (cid:16)tm ∂m a.e. λ ∈ Ω and we have an estimate ∂tm(cid:0)Tt(x)(cid:1)(λ)(cid:17)t>0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)λ 7→(cid:16)tm ∂m ∂tm(cid:0)Tt(x)(cid:1)(λ)(cid:17)t>0(cid:13)(cid:13)V q(cid:13)(cid:13)(cid:13)p . kxkp, x ∈ Lp(Ω). Proof. Let m ≥ 0 be an integer. It follows from [31, Lemma, p. 72] that for any x ∈ Lp(Ω), the function 17 belongs to V q for t 7→ tm ∂m ∂tm(cid:0)Tt(x)(cid:1)(λ) is continuous for a.e. λ ∈ Ω. Next it follows from the proof of [20, Cor. 4.2] that there exists a constant Cm > 0 such (cid:13)(cid:13)(cid:13)(cid:16) ∞Xn=0 (n + 1)2m+1(cid:12)(cid:12)T n 2(cid:13)(cid:13)(cid:13)p t (Tt − I)m(x)(cid:12)(cid:12)2(cid:17) 1 ≤ Cm kxkp for any t > 0 and any x ∈ Lp(Ω). That is, the operators Tt satisfy Proposition 4.1 uniformly. Since they also satisfy Theorem 3.1 uniformly, it follows from the proof of Theorem 4.4 that they satisfy the estimate (4.6) uniformly. Hence using an approximation argument as in the proof of [20, Cor. 4.2], we deduce that for any x ∈ Lp(Ω) and for any 0 < t0 < t1 < · · · < tm, we have (4.17) that The result therefore follows from Lemma 2.2. (cid:3) (cid:13)(cid:13)(cid:0)Tt0(x), Tt1(x), . . . , Ttm(x)(cid:1)(cid:13)(cid:13)Lp(Ω;vq m) ≤ C. 5. Additional properties. We give here further properties of contractively regular operators and contractively regular semigroups, in connection with variational inequalities. Let (Ω, µ) be a measure space and let 1 < p < ∞. 5.1. Individual ergodic theorems Let T : Lp(Ω) → Lp(Ω) be a contraction. According to the Mean Ergodic Theorem, we have a direct sum decomposition Lp(Ω) = N(I − T ) ⊕ R(I − T ), where N(· ) and R(· ) denote the kernel and the range, respectively. Moreover if we let PT : Lp(Ω) → Lp(Ω) denote the corresponding projection onto N(I − T ), then (5.1) for any x ∈ Lp(Ω). It is well-known that if T is an absolute contraction, then Mn(T )x → PT (x) almost everywhere (see e.g. [12, Section VIII.6]). We extend this classical result, as follows. Mn(T )x Lp −→ PT (x) Corollary 5.1. Assume that T : Lp(Ω) → Lp(Ω) is contractively regular. Then for any x ∈ Lp(Ω), [Mn(T )x](λ) −→ [PT (x)](λ) for a.e. λ ∈ Ω. 18 CHRISTIAN LE MERDY, QUANHUA XU Proof. Let x ∈ Lp(Ω) and let 2 < q < ∞. According to Theorem 3.1, the sequence (cid:0)[Mn(T )x](λ)(cid:1)n≥0 belongs to vq for almost every λ ∈ Ω. Hence (cid:0)[Mn(T )x](λ)(cid:1)n≥0 converges for almost every λ ∈ Ω. Combining with (5.1), we obtain the result. (cid:3) If a contraction T : Lp(Ω) → Lp(Ω) is analytic, then kT n − T n+1k → 0, hence T n(x) → 0 for any x ∈ R(I − T ). Consequently, T n(x) Lp −→ PT (x) for any x ∈ Lp(Ω). Using Theorem 4.4 and arguing as above, we obtain the following. Corollary 5.2. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator and assume that T is analytic. Then for any x ∈ Lp(Ω), [T n(x)](λ) −→ [PT (x)](λ) for a.e. λ ∈ Ω. We now consider the continuous case. The situation is essentially similar, except that we can also consider the behaviour when the parameter t tends to 0+. Let T = (Tt)t≥0 be a strongly continuous semigroup of contractions. By definition, for any x ∈ Lp(Ω), Tt(x) → x in the Lp-norm when t → 0+. This implies that Mt(T )x → x when t → 0+. Let A denote the infinitesimal generator of T = (Tt)t≥0. As in the discrete case, we have a direct sum decomposition Lp(Ω) = N(A) ⊕ R(A). Moreover if we let PA : Lp(Ω) → Lp(Ω) denote the corresponding projection onto N(A), then for any x ∈ Lp(Ω). If further (Tt)t≥0 is a bounded analytic semigroup, then Mt(T )x Lp −→ PA(x) when t → ∞ Tt(x) Lp −→ PA(x) for any x ∈ Lp(Ω). Now applying Corollaries 3.4 and 4.5, we deduce the following individual ergodic theorems. Corollary 5.3. Let T = (Tt)t≥0 be a strongly continuous semigroup of contractively regular operators on Lp(Ω), and let x ∈ Lp(Ω). Then for almost every λ ∈ Ω, and [Mt(T )x](λ) −→ [PA(x)](λ) when t → ∞ [Mt(T )x](λ) −→ x(λ) when t → 0+. Corollary 5.4. Let T = (Tt)t≥0 be a bounded analytic semigroup and assume that Tt is contractively regular for any t ≥ 0. Let x ∈ Lp(Ω). Then for almost every λ ∈ Ω, and [Tt(x)](λ) −→ [PA(x)](λ) when t → ∞ [Tt(x)](λ) −→ x(λ) when t → 0+. 5.2. The case q = 2 In this section, we fix a increasing sequence (nk)k≥0 of integers, with n0 = 0. Given any sequence (an)n≥0 of complex numbers, we define the so-called oscillation norm k(an)n≥0ko2 = (cid:0)a02 + Xk≥0 max nk≤n,m≤nk+1 an − am2(cid:1) 1 2 , 19 and we let o2 denote the Banach space of all sequences with a finite oscillation norm, equipped with k ko2. This space (whose definition depends on the sequence (nk)k≥0) was used in [4, 16, 17] as a substitute to vq in the case q = 2 (see also [14]). Indeed, neither Theorem 3.1 nor Theorem 4.4 holds true for q = 2, see [16] and [19, Section 8]. Recall the shift operator sp : ℓp Z → ℓp Z for any 1 < p < ∞ (see Section 3). According to [16, Thm. A], there is a constant Cp,2 such that (cid:13)(cid:13)(cid:0)Mn(sp)c(cid:1)n≥0(cid:13)(cid:13)Lp(o2) ≤ Cp,2 kckp for any c ∈ ℓp o2-version of the latter statement. Z. Hence arguing as in the proof of Theorem 3.1, we obtain the following Theorem 5.5. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator, with 1 < p < ∞. Then we have (cid:13)(cid:13)(cid:0)Mn(T )x(cid:1)n≥0(cid:13)(cid:13)Lp(o2) ≤ Cp,2 kxkp, We also have an o2-version of Theorem 4.4, as follows. x ∈ Lp(Ω). Theorem 5.6. Let T : Lp(Ω) → Lp(Ω) be a contractively regular operator, with 1 < p < ∞, and assume that T is analytic. Then we have an estimate (5.2) (5.3) More generally, for any integer m ≥ 0, we have an estimate (cid:13)(cid:13)(cid:0)T n(x)(cid:1)n≥0(cid:13)(cid:13)Lp(o2) . kxkp, n (x)(cid:1)n≥1(cid:13)(cid:13)Lp(o2) . kxkp, (cid:13)(cid:13)(cid:0)nm∆m x ∈ Lp(Ω). x ∈ Lp(Ω). Proof. The proof is a variant of the one written for Theorem 4.4, let us explain this briefly. We use the notation from Section 4. For any m ≥ −1, consider the following three properties: (i)m (cid:13)(cid:13)(cid:0)nm∆m (ii)m (cid:13)(cid:13)(cid:0)nm∆m (iii)m (cid:13)(cid:13)(cid:0)nm∆m n (x)(cid:1)n≥1(cid:13)(cid:13)Lp(o2) . kxkp; n+1(x)(cid:1)n≥1(cid:13)(cid:13)Lp(o2) . kxkp; 2n+1(x)(cid:1)n≥1(cid:13)(cid:13)Lp(o2) . kxkp. Property (i)m is the result we wish to prove. Our strategy is to show, by induction, that these three estimates hold true. First, it is easy to deduce from Proposition 4.1 that (i)m and (ii)m are equivalent. Second it follows from (4.14), (4.15) and (4.10) that (ii)m−1 and (iii)m−1 imply (iii)m and (i)m. Indeed v2 ⊂ o2 and nm∆m 2n+1 − Bn. Hence it suffices to show (i)−1 and (iii)−1. This is obtained by applying Theorem 5.5 twice, the first time for the o2-space associated with the sequence (nk)k≥1, the second time for o2-space associated with the sequence (2nk +1)k≥1. (cid:3) n = nm∆m 20 CHRISTIAN LE MERDY, QUANHUA XU There are also o2-versions of Corollary 3.4 and Corollary 4.5, whose statements are left to the reader. 5.3. Jump functions It is well known that variational inequalities for a sequence of operators have consequences in terms of jump functions. For any τ > 0 and any sequence a = (an)n≥0 of complex numbers, let N(a, τ ) denote the number of τ -jumps of a, defined as the supremum of all integers N ≥ 0 for which there exist integers such that amk − ank > τ for each k = 1, . . . , N. It is clear that for any 1 ≤ q < ∞, 0 ≤ n1 < m1 ≤ n2 < m2 ≤ · · · ≤ nN < mN , τ qN(a, τ ) ≤ kakq vq . Combining with Theorem 3.1 and Theorem 4.4, we immediately obtain (as in [17, Thm. 3.15]) the following jump estimates. Corollary 5.7. Consider 1 < p < ∞ and 2 < q < ∞. Let T : Lp(Ω, µ) → Lp(Ω, µ) be a contractively regular operator. (1) We have an estimate (2) Assume moreover that T is analytic. Then we have similar estimates kxkp , τ kxkp p p q τ pK , . (cid:13)(cid:13)(cid:13)λ 7→ N(cid:16)(cid:0)[Mn(T )x](λ)(cid:1)n≥0 q(cid:13)(cid:13)(cid:13)p , τ(cid:17) 1 µnλ ∈ Ω(cid:12)(cid:12)(cid:12) N(cid:16)(cid:0)[Mn(T )x](λ)(cid:1)n≥0 , τ(cid:17) > Ko . q(cid:13)(cid:13)(cid:13)p (cid:13)(cid:13)(cid:13)λ 7→ N(cid:16)(cid:0)[T n(x)](λ)(cid:1)n≥0 , τ(cid:17) 1 µnλ ∈ Ω(cid:12)(cid:12)(cid:12) N(cid:16)(cid:0)[T n(x)](λ)(cid:1)n≥0 , τ(cid:17) > Ko . . τ kxkp , kxkp p p q τ pK , and for any K > 0, we also have and Furthermore for any integer m ≥ 0, similar results hold with nm∆m n instead of T n. Similar results for the continuous case can be deduced as well from Corollary 3.4 and Corollary 4.5. 6. Examples and applications. We will now exhibit various classes of operators or semigroups to which our results from Section 4 and Section 5 apply. We focus on statements involving q-variation, although statements involving the oscillation norm from the subsection 5.2 are also possible. We start with the continuous case. Let (Tt)t≥0 be a strongly continuous semigroup on L2(Ω) and assume that each Tt is an absolute contraction, that is, (6.1) kTt(x)k1 ≤ kxk1 and kTt(x)k∞ ≤ kxk∞ 21 for any x ∈ L1(Ω) + L∞(Ω) and any t > 0 (see Section 3). Thus for any 1 < p < ∞, (Tt)t≥0 extends to a strongly continuous semigroup of contractively regular operators. It is well-known that (Tt)t≥0 is a bounded analytic semigroup on Lp(Ω) for every 1 < p < ∞ if (and only if) it is a bounded analytic semigroup on Lp(Ω) for one 1 < p < ∞. Applying Corollary 4.5 and Corollary 5.4, we derive the following. Corollary 6.1. Let (Tt)t≥0 be a bounded analytic semigroup on L2(Ω) satisfying (6.1). Then it satisfies estimates (4.16) and (4.17) for every 1 < p < ∞ and every 2 < q < ∞. Moreover for any x ∈ Lp(Ω), Tt(x) converges almost everywhere when t → 0+ and when t → ∞. Note that if each Tt : L2(Ω) → Lp(Ω) is selfadjoint, then (Tt)t≥0 is a bounded analytic semigroup on L2(Ω). Hence the above corollary applies to symmetric diffusion semigroups and extends [17, Thm. 3.3]. We now consider the so-called subordinated semigroups. Let 1 < p < ∞ and let (Tt)t≥0 be a strongly continuous bounded semigroup on Lp(Ω). Let A denote its infinitesimal generator and let 0 < α < 1. Then −(−A)α generates a bounded analytic semigroup (Tα,t)t≥0 on Lp(Ω), and for any t > 0, there exists a continuous function fα,t : (0, ∞) → R such that (6.2) and (6.3) ∀ s > 0, fα,t(s) ≥ 0; Z ∞ 0 fα,t(s) ds = 1; Tα,t(x) = Z ∞ 0 fα,t(s) Ts(x) ds , x ∈ Lp(Ω). See e.g. [32, IX.11] for details and complements. Corollary 6.2. Let (Tt)t≥0 be a strongly continuous semigroup on Lp(Ω), with 1 < p < ∞, and assume that Tt : Lp(Ω) → Lp(Ω) is contractively regular for any t ≥ 0. Then for any 0 < α < ∞ and any 2 < q < ∞, we have an estimate (cid:13)(cid:13)(cid:13)λ 7→(cid:13)(cid:13)(cid:0)[Tα,t(x)](λ)(cid:1)t>0(cid:13)(cid:13)V q(cid:13)(cid:13)(cid:13)p . kxkp for the subordinated semigroup Tα,t = e−t(−A)α. Moreover for any x ∈ Lp(Ω), Tα,t(x) con- verges almost everywhere when t → 0+ and when t → ∞. Proof. Let 0 < α < ∞. It follows from (6.2) and (6.3) that for any t > 0, kTα,tkr ≤Z ∞ 0 fα,t(s) kTskr ds ≤ 1. Hence the result follows from Corollary 4.5 and Corollary 5.4. (cid:3) We now turn to the discrete case and consider normal operators on L2. Lemma 6.3. Let H be a Hilbert space and let T ∈ B(H) be a normal operator. Then T is an analytic power bounded operator if and only if it satisfies (4.4). 22 CHRISTIAN LE MERDY, QUANHUA XU Proof. The 'only if' part holds for any operator, as discussed at the beginning of Section 4. Conversely, assume that a normal operator T : H → H satisfies (4.4). Applying the Spectral Theorem, we deduce that T is a contraction and that for some constant K > 0, we have (cid:3) nkT n − T n−1k = sup z∈σ(T ) nzn − zn−1 = sup z∈σ(T ) nzn−1 1 − z n(rn−1 − rn) = K(cid:16) n − 1 for any n ≥ 1. Hence the sequence (cid:0)n(T n − T n−1)(cid:1)n≥1 is bounded. ≤ K sup r∈[0,1] n (cid:17)n−1 The following is a straightforward consequence of the above lemma, Theorem 4.4 and Corollary 5.2. Corollary 6.4. Let T : L2(Ω) → L2(Ω) be a contractively regular normal operator satisfying (4.4). Then it satisfies an estimate (cid:13)(cid:13)(cid:0)T n(x)(cid:1)n≥0(cid:13)(cid:13)L2(vq) . kxk2, x ∈ L2(Ω) for any 2 < q < ∞. Moreover for any x ∈ L2(Ω), T n(x) converges almost everywhere when n → ∞. The following is a discrete analog of Corollary 6.1. Corollary 6.5. Let T : L1(Ω) + L∞(Ω) → L1(Ω) + L∞(Ω) be an absolute contraction and assume that T : L2(Ω) → L2(Ω) is analytic. Then T satisfies estimates (4.5) and (4.6) for every 1 < p < ∞ and every 2 < q < ∞. Moreover for any x ∈ Lp(Ω), T n(x) converges almost everywhere when n → ∞. Proof. If T : L2(Ω) → L2(Ω) is analytic, then T : Lp(Ω) → Lp(Ω) is analytic as well for any 1 < p < ∞, by [3, Thm. 1.1]. Hence the result follows from Theorem 4.4 and Corollary 5.2. (cid:3) Of course the above corollary applies if T : L2(Ω) → L2(Ω) is a positive selfadjoint opera- tor, more generally if T is normal and satisfies (4.4). As a consequence, we extend the main result of [17], as follows. The next statement solves a problem raised in the latter paper. Corollary 6.6. Let G be a locally compact abelian group and let Lp(G) denote the corre- sponding Lp-spaces with respect to a Haar measure. Let ν be a probability measure on G and let T : L1(G) + L∞(G) → L1(G) + L∞(G) be the associated convolution operator, T (x) = ν ∗ x. Assume that there exists a constant K > 0 such that 1 −bν(s) ≤ K(1 − bν(s)) for any s ∈ bG. Then we have an estimate x ∈ Lp(Ω), for any 1 < p < ∞ and any 2 < q < ∞. (cid:13)(cid:13)(cid:0)T n(x)(cid:1)n≥0(cid:13)(cid:13)Lp(vq) . kxkp, 23 Proof. Regard T as an L2-operator. By Fourier analysis, its spectrum is equal to the essential It therefore follows from the assumption and Lemma 6.3 that the operator T : L2(G) → L2(G) is analytic. Hence T satisfies the assumptions of Corollary 6.5, which yields the result. (cid:3) range of bν. Remark 6.7. For an operator T ∈ B(L2(Ω)), let W (T ) = {hT (x), xi : kxk2 = 1} denote the numerical range. We recall that W (T ) is a compact convex set and that σ(T ) ⊂ W (T ). Assume that there exists γ ∈ (0, π 2 ) such that W (T ) ⊂ Bγ (which is a stronger condition than (4.3) or (4.4)). Then according to [11], there exists a constant C > 0 such that kf (T )k ≤ C sup(cid:8)f (z) : z ∈ Bγ(cid:9) for any polynomial f . Arguing as in the proof of Lemma 6.3, we deduce that T is an analytic power bounded operator. Consequently if T : L2(Ω) → L2(Ω) is contractively regular and W (T ) ⊂ Bγ for some γ ∈ (0, π 2 ), then it satisfies (4.5). References [1] M. Akcoglu, and L. Sucheston, Dilations of positive contractions on Lp spaces, Canad. Math. Bull. 20 (1977), 285-292. [2] N. Asmar, E. Berkson, and T. A. Gillespie, Transference of strong type maximal inequalities by separation-preserving representations, Amer. J. Math. 113 (1991), 47-74. [3] S. Blunck, Analyticity and discrete maximal regularity on Lp-spaces, J. Funct. Anal. 183 (2001), 211-230. [4] J. Bourgain, Pointwise ergodic theorems for arithmetic sets, Publ. Math. IHES 69 (1989), 5-41. [5] J. T. Campbell, R. L. Jones, K. Reinhold, and M. Wierdl, Oscillation and variation for the Hilbert transform, Duke Math. J., 105 (2000), 59-83. [6] R. Coifman, R. Rochberg, and G. Weiss, Applications of transference: the Lp version of von Neu- mann's inequality and the Littlewood-Paley-Stein theory, pp. 53-67 in "Linear spaces and Approxima- tion", Birkhauser, Basel, 1978. [7] R. R. Coifman, and G. Weiss, Transference methods in analysis, CBMS 31, Amer. Math. Soc., 1977. [8] T. Coulhon, and L. Saloff-Coste, Puissances d'un op´erateur r´egularisant, Ann. Inst. H. Poincar´e Probab. Statist. 26 (1990), no. 3, 419-436. [9] R. Crescimbeni, R. A. Mac´ıas, T. Men´arguez, J. L. Torrea, and B. Viviani, The ρ-variation as an operator between maximal operators and singular integrals, J. Evol. Equ., 9 (2009), 81-102. [10] A. de la Torre, A simple proof of the maximal ergodic theorem, Canad. J. Math. 28 (1976), 1073-1075. [11] B. Delyon, and F. Delyon, Generalization of von Neumann's spectral sets and integral representation of operators, Bull. Soc. Math. France 127 (1999), 25-41. [12] N. Dunford, and J. T. Schwartz, Linear operators, Part 1, Pure and Applied Mathematics, Vol. 7 Interscience Publishers, Inc., New York; Interscience Publishers, Ltd., London 1958 xiv+858 pp. [13] G. Fendler, Dilations of one parameter semigroups of positive contractions on Lp-spaces, Canad. J. Math. 49 (1997), 736-748. [14] V.F. Gaposhkin, Ergodic theorem for functions of normal operators (Russian), Funktsional. Anal. i Prilozhen. 18 (1984), 1-6. [15] J. A. Goldstein, Semigroups of linear operators and applications, Oxford University Press, New-York, 1985. [16] R. L. Jones, R. Kaufman, J. M. Rosenblatt, and M. Wierdl, Oscillation in ergodic theory, Ergodic Theory Dynam. Systems 18 (1998), 889-935. [17] R. L. Jones, and K. Reinhold, Oscillation and variation inequalities for convolution powers, Ergodic Theory Dynam. Systems 21 (2001), no. 6, 1809-1829. 24 CHRISTIAN LE MERDY, QUANHUA XU [18] R. L. Jones, A. Seeger, and J. Wright, Strong variational and jump inequalities in harmonic analysis, Trans. Amer. Math. Soc., 360 (2008), 6711-6742. [19] R. L. Jones, and G. Wang, Variation inequalities for the Fej´er and Poisson kernels, Trans. Amer. Math. Soc., 356 (2004), 4493-4518. [20] C. Le Merdy, and Q. Xu, Maximal theorems and square functions for analytic operators on Lp-spaces, Preprint 2010, arXiv:1011.1360v1. [21] D. L´epingle, La variation d'ordre p des semi-martingales (French) Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 36 (1976), 295-316. [22] Yu. Lyubich, Spectral localization, power boundedness and invariant subspaces under Ritt's type condi- tion, Studia Math. 134 (1999), 153-167. [23] P. Meyer-Nieberg, Banach lattices, Springer, Berlin-Heidelberg-NewYork, 1991. [24] B. Nagy, and J. Zemanek, A resolvent condition implying power boundedness, Studia Math. 134 (1999), 143-151. [25] O. Nevanlinna, Convergence of iterations for linear equations, Birkhauser, Basel, 1993. [26] R. Oberlin, A. Seeger, T. Tao, C. Thiele, J. Wright, A variation norm Carleson theorem, Preprint 2010, arXiv:1011.1360v1. [27] A. Pazy, Semigroups of linear operators and applications to partial differential equations, Springer, 1983. [28] V. Peller, An analogue of J. von Neumann's inequality for the space Lp (Russian), Dokl. Akad. Nauk SSSR 231 (1976), no. 3, 539-542. [29] G. Pisier, Complex interpolation and regular operators between Banach lattices, Arch. Math. (Basel) 62 (1994), no. 3, 261-269. [30] G. Pisier, and Q. Xu, The strong p-variation of martingale and orthogonal series, Probab. Th. Rel. Fields 77 (1988), 497-514. [31] E. M. Stein, Topics in harmonic analysis related to the Littlewood-Paley theory, Ann. Math. Studies, Princeton, University Press, 1970. [32] K. Yosida, Functional Analysis, Springer Verlag, 1968. Laboratoire de Math´ematiques, Universit´e de Franche-Comt´e, 25030 Besanc¸on Cedex, France E-mail address: [email protected] School of Mathematics and Statistics, Wuhan University, Wuhan 430072, Hubei, China Laboratoire de Math´ematiques, Universit´e de Franche-Comt´e, 25030 Besanc¸on Cedex, France E-mail address: [email protected]
1803.01397
1
1803
2018-03-04T18:23:07
Universal bounds for the Hardy--Littlewood inequalities on multilinear forms
[ "math.FA" ]
The Hardy--Littlewood inequalities for multilinear forms on sequence spaces state that for all positive integers $m,n\geq2$ and all $m$-linear forms $T:\ell_{p_{1}}^{n}\times\cdots\times\ell_{p_{m}}^{n}\rightarrow\mathbb{K}$ ($\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$) there are constants $C_{m}\geq1$ (not depending on $n$) such that \[ \left( \sum_{j_{1},\ldots,j_{m}=1}^{n}\left\vert T(e_{j_{1}},\ldots,e_{j_{m}})\right\vert ^{\rho}\right) ^{\frac{1}{\rho}}\leq C_{m}\sup_{\left\Vert x_{1}\right\Vert ,\dots,\left\Vert x_{m}\right\Vert \leq 1}\left\vert T(x_{1},\dots,x_{m})\right\vert, \] where $\rho=\frac{2m}{m+1-2\left( \frac{1}{p_{1}}+\cdots+\frac{1}{p_{m}}\right) }$ if $0\leq\frac{1}{p_{1}}+\cdots+\frac{1}{p_{m}}\leq\frac{1}{2}$ or $\rho=\frac{1}{1-\left( \frac{1}{p_{1}}+\cdots+\frac{1}{p_{m}}\right)}$ if $\frac{1}{2}\leq\frac{1}{p_{1}}+\cdots+\frac{1}{p_{m}}<1$. Good estimates for the Hardy-Littlewood constants are, in general, associated to applications in Mathematics and even in Physics, but the exact behavior of these constants is still unknown. In this note we give some new contributions to the behavior of the constants in the case $\frac{1}{2}\leq\frac{1}{p_{1}}+\cdots+\frac{1}{p_{m}}<1$. As a consequence of our main result, we present a generalization and a simplified proof of a result due to Aron et al. on certain Hardy--Littlewood type inequalities.
math.FA
math
UNIVERSAL BOUNDS FOR THE HARDY -- LITTLEWOOD INEQUALITIES ON MULTILINEAR FORMS G. ARA ´UJO AND K. C AMARA Abstract. The Hardy -- Littlewood inequalities for multilinear forms on sequence spaces state that for all positive integers m, n ≥ 2 and all m-linear forms T : ℓn p1 × · · · × ℓn pm → K (K = R or C) there are constants Cm ≥ 1 (not depending on n) such that T (ej1 , . . . , ejm )ρ! 1 ρ ≤ Cm sup T (x1, . . . , xm) , kx1k,...,kxmk≤1 n Xj1 ,...,jm=1 where ρ = 2m if 0 ≤ 1 p1 + · · · + 1 pm ≤ 1 2 or ρ = pm (cid:17) m+1−2(cid:16) 1 p1 1 p1 + · · · + 1 pm +···+ 1 2 ≤ 1 < 1. Good estimates for the Hardy-Littlewood constants are, in general, associated to applications in Mathematics and even in Physics, but the exact behavior of these constants is still unknown. In this note we give some new contributions to the behavior of the constants in the case 1 < 1. As a consequence of our main result, we present a generalization and a simplified proof of a result due to Aron et al. on certain Hardy -- Littlewood type inequalities. + · · · + 1 pm +···+ 1 2 ≤ 1 p1 1−(cid:16) 1 p1 1 if pm (cid:17) 1. Introduction Let E, E1, ..., Em and F be Banach spaces over K = R or C and for all m-linear maps T : E1 × ··· × Em → F let us denote kTk := sup kx1k,...,kxmk≤1kT (x1, . . . , xm)k . Also, let c0 = {(xn)∞ that n=1 ⊂ K : lim xn = 0}. Littlewood's 4/3 inequality [13] (1930) asserts ∞ Xj,k=1   T (ej, ek) 3 4 4 3  √2kTk, ≤ for all continuous bilinear forms T : c0 × c0 → C, and the exponent 4/3 is sharp. Littlewood's 4/3 inequality was the starting point of several important inequalities, such as an inequality due to Bohnenblust and Hille (1931), which nowadays is known to be important for applications in physics (see [14]). The Bohnenblust -- Hille inequality 2010 Mathematics Subject Classification. 46G25, 47H60 (primary), 47A63, 41A44, 34C11 (secondary). Key words and phrases. Absolutely summing operators, Constants, Hardy -- Littlewood inequalities. 1 2 ARA ´UJO AND C AMARA [9] assures the existence of a constant Bm ≥ 1 such that ∞ Xj1,...,jm=1   T (ej1, . . . , ejm) 2m m+1  m+1 2m ≤ Bm kTk , Of course, if m = 2 we recover Littlewood's 4/3 inequality. for all continuous m -- linear forms T : c0 × ··· × c0 → C. In 1934 Hardy and Littlewood [12] extended Littlewood's 4/3 inequality to bilinear maps defined on ℓp× ℓq, where by ℓs, s ≥ 1, we mean the Banach space of all absolutely s -- summable sequences in K (of course, if s = ∞ by ℓ∞ we mean the space of all bounded sequences in K). In 1981, Praciano-Pereira [17] extended the Hardy -- Littlewood inequalities to m-linear pm ≤ 1 forms on ℓp1 × ··· × ℓpm for 0 ≤ 1 2 and very recently Dimant and 2 ≤ 1 Sevilla-Peris [11] generalized the estimates for the case 1 < 1 (all these inequalities are nowadays called Hardy -- Littlewood inequalities). + ··· + 1 + ··· + 1 pm p1 p1 From now on, for any function f , whenever it makes sense we formally define f (∞) = limp→∞ f (p). Moreover, for p = (p1, . . . , pm) ∈ [1,∞]m and 1 ≤ k ≤ m, let us denote (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≤k := 1 p1 + ··· + 1 pk , := 1 pk + ··· + 1 pm and (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12) := (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≤m = (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥1 and, as usual, for s ∈ [1,∞] and a positive integer n we define ℓn s = Kn equipped with the ℓs-norm (sup norm if s = ∞); also, ej represents the canonical vector of c0 with 1 in the j-th coordinate and 0 elsewhere. The classical Hardy -- Littlewood inequalities can be stated as follows: Theorem 1.1 (Bohnenblust, Dimant, Hardy, Hille, Littlewood, Praciano-Pereira, Sevil- la-Perez). Let m ≥ 2 be a positive integer and p = (p1, . . . , pm) ∈ (1,∞]m with 0 ≤ (cid:12)(cid:12)(cid:12) < 1. Then there are constants C K p(cid:12)(cid:12)(cid:12) 1 (1.1) (1.2) n Xj1,...,jm=1 n Xj1,...,jm=1     T (ej1, . . . , ejm) T (ej1, . . . , ejm) 2m m+1−2 1 p m,p ≥ 1 such that p  1−(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12) ≤ C K 1 2m m+1−2 1 1 p 1− 1  ≤ C K 1 2 , ≤ 1 p(cid:12)(cid:12)(cid:12)(cid:12) m,p kTk if 0 ≤ (cid:12)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12)(cid:12) 1 1 < 1, m,p kTk if 2 ≤ (cid:12)(cid:12)(cid:12)(cid:12) m,p. When (cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12) 1 m+1−2(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) for all m-linear forms T : ℓn p1 × ··· × ℓn pm → K and all positive integers n. If p1 = ··· = pm = p we denote C K p1 = ··· = pm = ∞), since Hille inequality. Using the generalized Kahane -- Salem -- Zygmund inequality in (1.1) and and Holder's inequality in (1.2) it is possible to conclude that the exponents m+1−2(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12) 2m 1 m,p by C K = 0 (equivalently m+1 , we recover the classical Bohnenblust -- = 2m 2m are optimal: if replaced by smaller exponents the constants appearing on the 1 1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) right-hand-size will depend on n. UNIVERSAL BOUNDS FOR THE HARDY -- LITTLEWOOD INEQUALITIES ON MULTILINEAR FORMS3 The precise growth of the constants C K < 1, is important for many applications and remains an open problem in Mathematical Analysis. The first estimates for C K m,p had exponential growth; more precisely, 1 p(cid:12)(cid:12)(cid:12) m,p, 0 ≤ (cid:12)(cid:12)(cid:12) m,p ≤ (cid:16)√2(cid:17)m−1 . C K 1 p(cid:12)(cid:12)(cid:12) ≤ 1 The case 0 ≤ (cid:12)(cid:12)(cid:12) 2 was more explored since it appearance. Several studies have made significant progress in the context 0 ≤ (cid:12)(cid:12)(cid:12) 2 (see for instance [2, 3, 5, 6, 8]). For example, among other results, it was proved in [5, 8] that for 2m(m − 1)2 < p ≤ ∞ we have 1 p(cid:12)(cid:12)(cid:12) ≤ 1 ≈ κ1 · m0.36482, 2−log 2−γ 2 C R m,p < κ1 · m C C m,p < κ2 · m 1−γ 2 ≈ κ2 · m0.21139, for certain constants κ1, κ2 > 0, where γ is the Euler-Mascheroni constant. On the other hand, the case 1 < 1 was virtually unexplored and only recently in [1, 7] is that the original estimate was improved. Our main result generalizes some of the main results of [1, 7]. 2 ≤ (cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) One of the main results of [1] is the following result: Theorem 1.2. Let m ≥ 2 be a positive integer and p = (p1, . . . , pm) ∈ (1,∞] with pm → K and all positive < 1. Then, for all m-linear forms T : ℓn 1 2 ≤ (cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) integers n, n  Xj1,...,jm=1  T (ej1, . . . , ejm ) 1−(cid:12)(cid:12)(cid:12) 1 p 1− 1  1 p1 × ··· × ℓn p(cid:12)(cid:12)(cid:12) ≤ 2 (m−1)(cid:16)1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:17) kTk . As a consequence, when m < p1 = ··· = pm = p ≤ m+1, the optimal constants of the Hardy -- Littlewood inequalities are uniformly bounded by 2. In fact, for m < p ≤ m + 1 we have   n Xj1,...,jm=1 T (ej1, . . . , ejm) p−m p p p−m  ≤ 2 m−1 m+1 kTk < 2kTk , for all m-linear forms T : ℓn p × ··· × ℓn p → K and all positive integers n. Another important contribution in this setting ( 1 < 1) is the following result of Aron, N´unez-Alarc´on, Pellegrino and Serrano-Rodr´ıguez (see [7, Corollary 3.3]): Theorem 1.3. Let m ≥ 2 be a positive integer and p = (p1, . . . , pm) ∈ (1,∞]m be such that 1 < pm ≤ 2 < p1, . . . , pm−1 and 2 ≤ (cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) < 1. 1 2 ≤ (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12) 4 Then ARA ´UJO AND C AMARA n Xj1,...,j=m T (ej1, . . . , ejm)   1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) ≤ kTk, 1 p 1− 1  for all m-linear forms T : ℓn p1 × ··· × ℓn pm → K and all positive integers n. Our main result generalizes Theorem 1.2 and has as a consequence a more general result than Theorem 1.3. It is important to mention that the proof of our main result is not just an adaptation of the original proof of 1.2 and that the proof given in [7] for Theorem 1.3 is, in some sense, very extensive and complicated. Our approach is simpler and more self-contained. We begin this section by recalling some important auxiliary results that will be es- sential to our purpose. 2. Main results An important auxiliary result that will be used along this note is the Khinchine inequality for real and complex scalars. More precisely, the Khinchine inequality assures that for any 0 < q < ∞, there are positive constants AK q such that regardless of the positive integer n and of the scalar sequence (aj)n 1 2 n Aq  Xj=1  aj2  ≤  Z 1  0 where rj are the Rademacher functions. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) q j=1 we have dt  ajrj(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n Xj=1 1 q , The next result concerns the multilinear theory of absolutely summing operators initiated by Pietsch [16]. It was proved very recently by Albuquerque and Rezende in [4, Theorem 3] and also will be essential for us. First, let us present some required definitions. Let BE∗ be the closed unit ball of the topological dual of E. If 1 ≤ q ≤ ∞, the symbol q∗ represents the conjugate of q. It will be convenient to adopt that c ∞ = 0 s (E) the linear space of the sequences (xj)∞ for any c > 0; for s ≥ 1 we represent by ℓw j=1 in E such that (ϕ (xj))∞ j=1 ∈ ℓs for every continuous linear functional ϕ : E → K. For (xj)∞ s (E) the expression k(xj )∞ j=1 ∈ ℓw j=1 ks defines a norm on ℓw s (E). The space of all continuous m-linear operators T : E1 ×···× Em → F , with the sup norm, is denoted by L (E1, ..., Em; F ). For p, q ∈ [1, +∞)m, a multilinear operator T : E1 × ··· × Em → F is multiple (q; p)-summing if there exist a constant C > 0 such that j=1kw,s := supϕ∈BE∗ k (ϕ (xj))∞ ∞  Xj1=1  qm−1 qm ∞ qm T (x(1) j1 , . . . , x(m) jm  ··· Xjm=1(cid:13)(cid:13)(cid:13)  j )∞ j=1 ∈ ℓw (q;p) (E1, . . . , Em; F ). When q1 = ··· = qm = q, we denote Πm     )(cid:13)(cid:13)(cid:13) ≤ C ··· pk m F 1 q1 q1 q2  for all (x(k) tors by Πm (Ek). We represent the class of all multiple (q; p)-summing opera- Yk=1(cid:13)(cid:13)(cid:13) (x(k) j )∞ j=1(cid:13)(cid:13)(cid:13)w,pk (q;p) (E1, . . . , Em; F ) UNIVERSAL BOUNDS FOR THE HARDY -- LITTLEWOOD INEQUALITIES ON MULTILINEAR FORMS5 (q;p) (E1, . . . , Em; F ). For recent results on the theory of multiple (q; p)-summing by Πm operators we refer to [15]. Theorem 2.1 (Albuquerque and Rezende [4]). Let m be a positive integer and r ≥ 1, s, p, q ∈ [1,∞)m be such that and, for each k = 1, . . . , m, qk ≥ pk and 1 1 > 0 1 1 p(cid:12)(cid:12)(cid:12)(cid:12) r −(cid:12)(cid:12)(cid:12)(cid:12) q(cid:12)(cid:12)(cid:12)(cid:12)≥k sk −(cid:12)(cid:12)(cid:12)(cid:12) 1 q(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) r −(cid:12)(cid:12)(cid:12)(cid:12) 1 = . 1 p(cid:12)(cid:12)(cid:12)(cid:12)≥k Then Πm (r;p)(E1, . . . , Em; F ) ⊂ Πm (s,q)(E1, . . . , Em; F ) for any Banach spaces E1, . . . , Em, F and the inclusion operator has norm 1. Now we are able to present our main result. Theorem 2.2. Let m ≥ 2 be a positive integer and p = (p1, . . . , pm) ∈ (1,∞]m be such that Then   n Xj1,...,j=m T (ej1, . . . , ejm ) for all m-linear forms T : ℓn p1 × ··· × ℓn 1 1 < 1. p(cid:12)(cid:12)(cid:12)(cid:12) 2 ≤ (cid:12)(cid:12)(cid:12)(cid:12) 1−(cid:12)(cid:12)(cid:12) p pks(cid:19)(cid:21)kTk, 1− 1  pm → K and all positive integers n, where (s−1)(cid:20)1−(cid:18) 1 p(cid:12)(cid:12)(cid:12) ≤ 2 +···+ 1 pk1 1 1 s = min r : there exists pk1, . . . , pkr ∈ {p1, . . . , pm} with pki 6= pkj , i 6= j, and 1 < 1 2 ≤ 1 pk1 + ··· + 1 pkr Proof. For the sake of simplicity let us suppose that pk1 = p1, . . . , pks = ps. Since     it follows from the Theorem 1.2 that n Xj1,...,js=1   Ts(ej1, . . . , ejs) < 1, 1 1 p(cid:12)(cid:12)(cid:12)(cid:12)≤s 2 ≤ (cid:12)(cid:12)(cid:12)(cid:12) p≤s 1− 1  1 1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≤s 1 (s−1)(cid:20)1−(cid:12)(cid:12)(cid:12) ≤ 2 p(cid:12)(cid:12)(cid:12)≤s(cid:21)kTsk ps → K and all positive integers n. for all s-linear forms Ts : ℓn In view of the Kinchine's inequality we have, for every n and all (s + 1)-linear forms Ts+1 : p1 × ··· × ℓn 6 ARA ´UJO AND C AMARA ℓn p1 × ··· × ℓn ps × ℓn ∞ → K, n  Xj1,...,js=1  n  2 Xjs+1=1(cid:12)(cid:12)Ts+1(cid:0)ej1, . . . , ejs+1(cid:1)(cid:12)(cid:12)   1 2 · 1 p(cid:12)(cid:12)(cid:12)≤s 1− 1 1 1−(cid:12)(cid:12)(cid:12) p≤s  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n Xjs+1=1 Ts+1(cid:0)ej1, . . . , ejs+1(cid:1) rjs+1(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ej1, . . . , ejs, Ts+1 Xjs+1=1 n n Ts+1 ej1, . . . , ejs, Xjs+1=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ej1, . . . , ejs, Ts+1 ·, . . . ,·, n Xjs+1=1 n Xjs+1=1 ≤  Xj1,...,js=1 n  A−1 K, 1 1− 1 p≤s  Z 1  0 n Xj1,...,js=1 0  Z 1    sup t∈[0,1] Ts+1 n n Xj1,...,js=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj1,...,js=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p(cid:12)(cid:12)(cid:12)≤s(cid:21) sup p(cid:12)(cid:12)(cid:12)≤s(cid:21)kTs+1k, t∈[0,1] 1 1 sup t∈[0,1]   (s−1)(cid:20)1−(cid:12)(cid:12)(cid:12) (s−1)(cid:20)1−(cid:12)(cid:12)(cid:12) 2 2 = A−1 K, 1 1− 1 p≤s ≤ A−1 K, 1 1− 1 p≤s = A−1 K, 1 1− 1 p≤s ≤ A−1 K, = A−1 K, 1− 1 p≤s 1 1 1− 1 p≤s 1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≤s p≤s p≤s 1− 1 1− 1  p(cid:12)(cid:12)(cid:12)≤s 1−(cid:12)(cid:12)(cid:12) 1 dt  1−(cid:12)(cid:12)(cid:12) dt  1−(cid:12)(cid:12)(cid:12) dt  1 p(cid:12)(cid:12)(cid:12)≤s 1 p(cid:12)(cid:12)(cid:12)≤s 1 1− 1 p≤s dt  1 1− 1 p≤s 1 1− 1 p≤s 1 1− 1 p≤s (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ejs+1rjs+1(t)  ejs+1rjs+1(t)  ejs+1rjs+1(t)  ejs+1rjs+1(t)  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) where AK, Since 1 1− 1 p≤s is the constant of the Khinchine inequality. ≥ 2, 1 1 −(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≤s UNIVERSAL BOUNDS FOR THE HARDY -- LITTLEWOOD INEQUALITIES ON MULTILINEAR FORMS7 we have AK, 1 1− 1 p≤s inclusion of ℓp spaces) = 1 and thus (from the previous inequality together with canonical 1 p(cid:12)(cid:12)(cid:12)≤s 1 1−(cid:12)(cid:12)(cid:12) p≤s  n   1− 1 n n =   Xj1,...,js=1 Xj1,...,js+1=1(cid:12)(cid:12)Ts+1(cid:0)ej1, . . . , ejs+1(cid:1)(cid:12)(cid:12)    Xj1,...,js=1  (s−1)(cid:20)1−(cid:12)(cid:12)(cid:12) ≤ 2 Xjs+1=1(cid:12)(cid:12)Ts+1(cid:0)ej1, . . . , ejs+1(cid:1)(cid:12)(cid:12)  2 Xjs+1=1(cid:12)(cid:12)Ts+1(cid:0)ej1, . . . , ejs+1(cid:1)(cid:12)(cid:12)   p(cid:12)(cid:12)(cid:12)≤s(cid:21)kTs+1k, ≤ n n 1 1− 1 1 (cid:18)1−(cid:12)(cid:12)(cid:12) p≤s  p≤s  1− 1 1 2 · 1 1 p(cid:12)(cid:12)(cid:12)≤s(cid:19)· 1− 1 1 p(cid:12)(cid:12)(cid:12)≤s 1 1−(cid:12)(cid:12)(cid:12) p≤s  1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≤s for every n and all (s + 1)-linear forms Ts+1 : ℓn ∞ → K. Using the canonical isometric isomorphisms for the spaces of weakly summable sequences (see [10, Proposition 2.2]) we know that this is equivalent to assert that (see [11, p. 308]), p1 × ··· × ℓn ps × ℓn Πs+1 1 1− 1 p≤s (E1, . . . , Es+1; K) = L(E1, . . . , Es+1; K) ;p∗ 1,...,p∗ s,1! for all Banach spaces E1, . . . , Es+1. From Theorem 2.1 it is possible to prove that Πs+1 (E1, . . . , Es+1; K) ⊆ Πs+1 s,1! 1,...,p∗ ;p∗ 1 1− 1 p≤s Consequently, (E1, . . . , Es+1; K). ;p∗ 1,...,p∗ s+1! 1 1− 1 p≤s+1 Πs+1 1 1− 1 p≤s+1 (E1, . . . , Es+1; K) = L(E1, . . . , Es+1; K) ;p∗ 1,...,p∗ s+1! for all Banach spaces E1, . . . , Es+1. Again (see [11, p. 308]), this is equivalent to say that n  Xj1,...,js+1=1(cid:12)(cid:12)Ts+1(cid:0)ej1, . . . , ejs+1(cid:1)(cid:12)(cid:12)  (s−1)(cid:20)1−(cid:12)(cid:12)(cid:12) ≤ 2 p(cid:12)(cid:12)(cid:12)≤s(cid:21)kTs+1k, 1 1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)≤s+1 1 p≤s+1 1− 1  for all (s + 1)-linear forms Ts+1 : ℓn The proof is completed by a standard induction argument. p1 × ··· × ℓn ps × ℓn ∞ → K and all positive integers n. (cid:3) 8 ARA ´UJO AND C AMARA Just making s = 1 in the previous result, we get the following Hardy -- Littlewood type inequalities with constant 1: Corollary 2.3. Let m ≥ 2 be a positive integer and p = (p1, . . . , pm) ∈ (1,∞]m be such that 1 < pi ≤ 2 < p1, . . . , pi−1, pi+1, . . . , pm for some 1 ≤ i ≤ m and Then < 1. 1 2 ≤ (cid:12)(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12)(cid:12) n Xj1,...,j=m T (ej1, . . . , ejm)   1−(cid:12)(cid:12)(cid:12) 1 p(cid:12)(cid:12)(cid:12) ≤ kTk, 1 p 1− 1  for all m-linear forms T : ℓn p1 × ··· × ℓn pm → K and all positive integers n. Corollary 2.3 generalizes a recently result proved independently in [7, Corollary 3.3]. Our approach is different and we believe it is more self-contained. References [1] N. Albuquerque, G. Ara´ujo, M. Maia, T. Nogueira, D. Pellegrino, and J. Santos, Optimal Hardy -- Littlewood inequalities niformly bounded by a universal constant, Annales Mathmatiques Blaise Pascal, to appear. [2] N. Albuquerque, F. Bayart, D. Pellegrino, and J. B. Seoane -- Sep´ulveda, Sharp generalizations of the multilinear Bohnenblust -- Hille inequality, J. Funct. Anal. 266 (2014), 3726-3740. [3] , Optimal Hardy -- Littlewood type inequalities for polynomials and multilinear operators, Israel J. Math. 211 (2016), no. 1, 197-220. [4] N. Albuquerque and L. Rezende, Anisotropic regularity principle in Contemporary Mathematics, in sequence spaces appear, DOI to applications, Communications and 10.1142/S0219199717500870. [5] G. Ara´ujo and D. Pellegrino, On the constants of the Bohnenblust-Hille and Hardy-Littlewood inequalities, Bull. Braz. Math. Soc. (N.S.) 48 (2017), no. 1, 141-169. [6] G. Ara´ujo, D. Pellegrino, and D. D. Silva e Silva, On the upper bounds for the constants of the Hardy -- Littlewood inequality, J. Funct. Anal. 267 (2014), 1878-1888. [7] R. Aron, D. N´unez-Alarc´on, D. Pellegrino, and D. Serrano-Rodr´ıguez, Optimal exponents for Hardy -- Littlewood inequalities for m-linear operators, Linear Algebra Appl. 531 (2017), 399 -- 422. [8] F. Bayart, D. Pellegrino, and J. B. Seoane -- Sep´ulveda, The Bohr radius of the n -- dimensional polydisc is equivalent to p(log n)/n, Advances in Math. 264 (2014), 726-746. [9] H. F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. (2) 32 (1931), no. 3, 600 -- 622. [10] J. Diestel, H. Jarchow, and A. Tonge, Absolutely summing operators, Cambridge Studies in Ad- vanced Mathematics, vol. 43, Cambridge University Press, Cambridge, 1995. [11] A. Dimant and P. Sevilla-Peris, Summation of Coefficients of Polynomials on ℓp Spaces, Publ. Mat. 60 (2016), no. 2, 289 -- 310. [12] G. Hardy and J. E. Littlewood, Bilinear forms bounded in space [p, q], Quart. J. Math. 5 (1934), 241 -- 254. [13] J. E. Littlewood, On bounded bilinear forms in an infinite number of variables, Quart. J. (Oxford Ser.) 1 (1930), 164 -- 174. [14] A. Montanaro, Some applications of hypercontractive inequalities in quantum information theory, J. Math. Physics 53 (2012), no. 12, 122206. [15] D. Pellegrino, J. Santos, D. Serrano-Rodr´ıguez, and E. Teixeira, A regularity principle in sequence spaces and applications, Bull. Sci. Math. 141 (2017), no. 8, 802 -- 837. UNIVERSAL BOUNDS FOR THE HARDY -- LITTLEWOOD INEQUALITIES ON MULTILINEAR FORMS9 [16] A. Pietsch, Ideals of multilinear functionals, Proceedings of the Second International Conference on Operator Algebras, Ideals and Their Applications in Theoretical Physics, Teubner -- texte Math. 67 (1983), 185 -- 199. [17] T. Praciano-Pereira, On bounded multilinear forms on a class of lp spaces, J. Math. Anal. Appl. 81 (1981), no. 2, 561 -- 568, DOI 10.1016/0022-247X(81)90082-2. (G. Ara´ujo) Universidade Estadual da Para´ıba, Campina Grande - PB (58.429-500), Brazil E-mail address: [email protected] (K. Camara) Universidade Federal da Para´ıba and Universidade Federal Rural do Semi- ´Arido, Joao Pessoa - PB (58.051-900) and Mossor´o - RN (59.625-900), Brazil E-mail address: [email protected]
1102.2136
1
1102
2011-02-10T14:56:58
Some critical point theorems and applications
[ "math.FA", "math.AP" ]
This paper is a continuation of \cite{Lu1}. In Part I, applying the new splitting theorems developed therein we generalize previous some results on computations of critical groups and some critical point theorems to weaker versions. In Part II (in progress), they are used to study multiple solutions for nonlinear higher order elliptic equations described in the introduction of \cite{Lu1}.
math.FA
math
Some critical point theorems and applications Guangcun Lu ∗ School of Mathematical Sciences, Beijing Normal University, Laboratory of Mathematics and Complex Systems, Ministry of Education, Beijing 100875, The People's Republic of China ([email protected]) February 9, 2011 Abstract 1 1 0 2 b e F 0 1 ] . A F h t a m [ 1 v 6 3 1 2 . 2 0 1 1 : v i X r a This paper is a continuation of [36]. In Part I, applying the new splitting theorems developed therein we generalize previous some results on computations of critical groups and some critical point theorems to weaker versions. In Part II (in progress), they are used to study multiple solutions for nonlinear higher order elliptic equations described in the introduction of [36]. Contents I Some critical point theorems 1 Introduction 2 Compactness conditions and deformation lemmas 3 Computations of critical groups 3.1 Critical groups at infinity and computations . . . . . . . . . . . . . . . 3.2 Computations of critical groups at degenerate critical points . . . . . . 4 Morse inequalities and some critical point theorems 4.1 Morse inequalities 4.2 Some critical point theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Critical groups of sign-changing critical points 2 2 2 8 8 11 13 13 15 17 ∗Partially supported by the NNSF 10671017 and 10971014 of China, and PCSIRT and Research Fund for the Doctoral Program Higher Education of China (Grant No. 200800270003). 1 II Applications 28 Part I Some critical point theorems 1 Introduction In previous many critical point theorems involving computations of critical groups the functionals are often assumed to be at least C 2 smooth so that the usual splitting lemma can be used. Doubtlessly, it is possible to obtain some new critical point theo- rems or to generalize previous ones by combing our splitting lemmas for continuously directional differentiable functionals with and techniques and results in nonsmooth and continuous critical point theories. Firstly, we shall generalize the some results in [5, 24] to weaker versions. Though they are not the weakest versions, our theo- rems are more convenient in applications because we do not need to compute such as subdifferentials and weak slopes, which are not easy actually. These are the main context in Section 5. Next, in Section 6 we shall present the corresponding version of the results on critical groups of sign-changing critical points in [4] and [34] in our framework and sketch how to prove them with our results in the previous sections. 2 Compactness conditions and deformation lem- mas Let X be a normed vector space with dual space X ∗, U ⊂ X nonempty and open, and f : U → R be a continuous functional. Recall in [22, 26, 27] that the weak slope of f at u ∈ U is the nonnegative extended real number df (u), which is the supremum of the σ's in [0, +∞[ such that there exist δ > 0 and H : BX(u, δ) × [0, δ] → X continuous with kH(v, t) − vk ≤ t and f (H(v, t)) ≤ f (v) − σt. Clearly, u → df (u) is lower semicontinuous ([22, Prop.2.6]). A point u ∈ X satisfying df (u) = 0 is called a lower critical point of f , and call c = f (u) a lower critical value of f . By [20, Def.1.1], the strong slope of a continuous function f : X → R at u ∈ U is defined by ∇f (u) =( 0 limv→u f (v)−f (u) kv−uk if u is a local minimum of f, otherwise. Then df (u) ≤ ∇f (u) for any u ∈ U (see below Definition 2.8 in [22]), and ∇f (u) ≤ kf ′(u)k (2.1) 2 provided that f has F-derivative f ′(u) at u ∈ U . For a locally Lipschitz continuous function f : U → R, by [14, page 27] it has the (Clarke) generalized gradient at every u ∈ U , ∂f (u) = {g ∈ X ∗ g(h) ≤ f ◦(u, h) ∀h ∈ X}, which is the subdifferential at θ of the convex function X → R, h 7→ f ◦(u, h), where f ◦(u, h) = limw→θ,s↓0 f (u + w + sh) − f (u + w) s is the generalized directional derivative of f at u in the direction h. If X is a Banach space it was proved in [22] that df (u) ≥ ∂f (u) := min{kx∗k : x∗ ∈ ∂f (u)} ∀u ∈ U. (2.2) This inequality may be strict by Example 1.1 in Krist´aly's thesis [29]. By [14, Prop.2.2.1] or [45, Prop.3.2.4(iii)], the function f : U → R is strictly H-differentiable at u0 ∈ U if and only if f is locally Lipschitz continuous around x0 and strictly G-differentiable at u0 ∈ U . In this case we have ∂f (u0) = {f ′(u0)} by [14, Prop.2.2.4]. Moreover, by [10, Prop.(6)] the set-valued mapping u → ∂f (u) is weak* upper semi-continuous at u0 in the sense that for any ǫ > 0, v ∈ X there exists a δ > 0 such that hw − f ′(u0), vi < ǫ for each w ∈ ∂f (u) with ku − u0k < δ. By [10, Prop.(7)] the function u → k∂f k(u) is lower semi-continuous at u0, i.e. limu→u0kf ′(u)k ≥ kf ′(u0)k. These are summarized into the following proposition. Proposition 2.1 Let U be a nonempty open set of a normed vector space X with dual space X ∗, and let f : U → R be strictly H-differentiable at every u ∈ U . Then (i) ∂f (u) = {f ′(u)} at any u ∈ U , the map U → X ∗, x 7→ f ′(x) is weak* continuous and the function U ∋ u → kf ′(u)k is lower semi-continuous. (ii) ∇f (u) = df (u) = kf ′(u)k ∀u ∈ U provided that f is F-differentiable at u ∈ U . (ii) comes from (2.1) and (2.2). Clearly, Proposition 2.1 holds if f is C 1. Let X be a Banach space and let f : X → R be a G-differential functional. Denote by K(f ) = {x ∈ X f ′(x) = 0}. For c ∈ R let K(f )c = {x ∈ X f (x) = c, f ′(x) = 0} and f c = {x ∈ X f (x) ≤ c}. If f is only continuous we write LK(f )c = {x ∈ X f (x) = c, df (x) = 0}. Definition 2.2 Let X be a Banach space and let f : X → R be a strictly H- differentiable functional. For c ∈ R the usual Palais-Smale compactness condition at the level c, or (P S)c for short, means that every (P S)c sequence {xn} ⊂ X, i.e. satisfying f (xn) → c and f ′(xn) → 0, has a convergent subsequence; moreover ac- cording to Cerami [8] we say that f satisfies the condition (C)c if every (C)c sequence {xn} ⊂ X, i.e. such that f (xn) → c and (1 + kxnk)kf ′(xn)k → 0, has a convergent subsequence. 3 Clearly, the second condition is weaker than the first one. Note that the semi- continuity of the map u 7→ kf ′(u)k (by Proposition 2.1) implies the limit of a (P S)c or (C)c sequence {xn} is in K(f )c. In particular, K(f )c is compact if f satisfies (P S)c or (C)c. The condition (C)c has the following equivalent form ([1, Definition 1.1]): (i) every bounded sequence (xn) in X with f (xn) → c and f ′(xn) → 0 has a convergent subsequence; (ii) there exist positive constants σ, R, α such that kf ′(x)k · kxk ≥ α for any x with c − σ ≤ f (x) ≤ c + σ and kxk ≥ R. Lemma 2.3 (First Deformation Lemma) For a strictly H-differentiable functional f : X → R on a (real) Banach space X, and c ∈ R, suppose that f satisfies the condition (C)c. Then for every ε0 > 0, every neighborhood U of K(f )c (if K(f )c = ∅ we take U = ∅), there exist an 0 < ε < ε0 and a map η ∈ C([0, 1] × X, X) satisfying (ii) η(t, x) = x if x /∈ f −1([c − ε0, c + ε0]); (i) kη(t, u) − uk ≤ e(1 + kuk)t, where e =P∞ (iii) η(cid:0){1} × (f c+ε \ U )(cid:1) ⊂ f c−ε; (iv) f (η(s, x)) ≤ f (η(t, x)) if s ≥ t; 1 n! ; n=0 (v) η(t, x) 6= x =⇒ f (η(t, x)) < f (x). In particular, (ii)-(iv) show that f satisfies the deformation condition (D)c in the sense of [5, Def. 3.1]. When f is even, η may be chosen so that η(t, ·) is odd for all t ∈ [0, 1]. When f is C 1 and satisfies the (P S)c (resp. (C)c) this lemma was proved by Palais [40] (see also [43]), (resp. Cerami [8] and Bartolo-Benci-Fortunato [1]). For a C 1−0-functional on a reflexive Banach space X, when f satisfies the (P S)c (resp. (C)c) Chang [10] (resp. Kourogenis and Papageorgiou[28]) proved this lemma. In the Proof of Lemma 2.3. The ideas are following those of [28, Theorem 4]. present case the reflexivity of X is not required because we do not need to use the Eberlein separation theorem as in the proofs of [10, Lemma 3.3] and [28, Lemma 3]. Let us reprove Lemma 3 of [28] under our assumptions as follows. By [28, Lemma 2], for each δ > 0 there exist γ > 0, ε > 0 such that for Kc = K(f )c, (1 + kxk)kf ′(x)k ≥ γ ∀x ∈ f −1([c − ε, c + ε]) \ Nδ(Kc), where Nδ(Kc) = {x ∈ X d(x, Kc) < δ}. For each x ∈ f −1([c − ε, c + ε]) \ Nδ(Kc), we have kf ′(x)k ≥ γ/(1 + kxk). Note that kf ′(x)k = sup{hf ′(x), hi h ∈ X, khk = 1}. We have hx ∈ X such that khxk = 1 and hf ′(x), hxi > 3γ 4(1 + kxk) > γ 2(1 + kxk) . By Proposition 2.1(i) we have rx > 0 such that hf ′(y), hxi > γ 2(1 + kxk) ∀y ∈ BX (x, rx). 4 Now {BX(x, rx)} is an open cover of f −1([c − ε, c + ε]) \ Nδ(Kc). Repeating the remaining arguments in the proof of [28, Lemma 2] we get a locally Lipschitz vector field V : f −1([c − ε, c + ε]) \ Nδ(Kc) → X such that kV (x)k ≤ (1 + kxk) and hf ′(x), V (x)i ≥ γ 2 . Shrinking ε > 0, γ > 0, δ > 0 so that N3δ(Kc) ⊂ U and that (8) of [28] is satisfied, and almost repeating the proof of [28, Theorem 4] we may get the desired conclusions. The unique point which should be noted is the proof of (iii). By contradiction, suppose that f (η(1, x)) > c − ε for some x ∈ f c+ε \ U . Then c − ε < f (η(t, x)) ≤ c + ε for all t ∈ [0, 1]. As in the proof (c) of [28, Theorem 4] it must hold that η([0, 1] × {x}) ∩ (K(f )c)2δ = ∅. It follows that there exist 0 ≤ t1 < t2 ≤ 1 such that d(η(t1, x), K(f )c) = 2δ, d(η(t1, x), K(f )c) = 3δ and 2δ < d(η(t, x), K(f )c) < 3δ for all t1 ≤ t ≤ t2. Repeating the remaining part of the proof (c) of [28, Theorem 4] yields (iii). ✷ Corresponding to [5, Corollary 3.3] we have Corollary 2.4 For a strictly H-differentiable functional f : X → R on a (real) Banach space X, we have (i) If f satisfies the condition (C)c for all c ∈ [a, b] and K(f )c = ∅ for c ∈ [a, b] then there exist a deformation ηt : X → X such that η0 = idX , ηt(x) = x if x /∈ f −1([a − 1, b + 1]), f (ηt(x)) is decreasing in t and η1(f b) ⊂ f a. (ii) If f satisfies the condition (C)c for all c ≥ a and K(f )c = ∅ for c ≥ a then there exist a deformation ηt : X → X such that η0 = idX , ηt(x) = x if f (x) ≤ a − 1, f (ηt(x)) is decreasing in t and η1(X) ⊂ f a. Proof. We only outline the proof of (i). For each c ∈ [a, b] Lemma 2.3 yields positive numbers ε(c) • η(c)(t, x) = x if x /∈ f −1([c − ε(c) 2 < 1 and a deformation η(c) : X → X such that η(c) 0 = idX and 1 < ε(c) 2 , c + ε(2) t 2 ]); • η(c)(cid:0){1} × (f c+ε(c) 1 (cid:1) ⊂ f c−ε(c) 1 ; • f (η(c)(s, x)) ≤ f (η(c)(t, x)) if s ≥ t. Since [a, b] is compact there exist finite numbers a ≤ c1 < · · · < ck ≤ b such that {(ci − ε(ci), ci + ε(ci))}k i=1 is an open cover of [a, b]. In particular we have c1 − ε(c1) 1 < a ≤ c1 < · · · < ck ≤ b < ck + ε(ck) 1 . Then the composition ηt = η(c1) the proof of [5, Corollary 3.3(b)] we can derive (ii) from (i). ✷ ◦ · · · ◦ η(ck) t t satisfies the desired requirements. As in Remark 2.5 Even if f is only continuous, if we replace the (C)c condition by the following (P S)c condition "f (xn) → c and df (xn) → 0 =⇒ ∃ a convergence subsequence of {xn}" 5 then Lemma 2.3 holds provided that K(f )c is replaced by LK(f )c. See [18, Theorem 2.14]. If f is F -differentiable so that ∇f (xn) = df (xn) = kf ′(xn)k ∀n, then under the (P S)c condition Lemma 2.3 is a corollary of [18, Theorem 2.14]. By the same reason, we may get the part I of the following the second deformation lemma from Theorem 2.3 of [16] or Theorem 4 and Remark 2 of [17]. Lemma 2.6 (Second Deformation Lemma) For a F-differentiable functional f : X → R on a (real) Banach space X, and −∞ < a < b ≤ +∞ suppose that f has only a finite number of critical points at the level a and has no critical values in (a, b). Then I. If f satisfies the condition (P S) on f −1([a, c]) for all c ∈ [a, b]∩ R, then there exists a deformation η : [0, 1] × f b◦ → f b◦ := {f < b} such that (a) f (η(t, u)) ≤ f (u); (b) u ∈ K(f )a =⇒ η(t, u) = u; (c) η({1} × f b◦) ⊂ f a◦ ∪ K(f )a; (d) if b ∈ R and K(f )b = ∅, then η can be extended to [0, 1] × X, still denoted by η, such that η({1} × f b) ⊂ f a◦ ∪ K(f )a. In particular, f a◦ ∪ K(f )a is a weak deformation retract of f b◦. II. If f is C 1 and satisfies the condition (C)c for all c ∈ [a, b] ∩ R, then f a is a strong deformation retract of f b \ K(f )b, i.e. there exists a map η : [0, 1] × (f b \ K(f )b) → (f b \ K(f )b), called a strong deformation retraction of f b \ K(f )b onto f b, satisfying (i) η(0, u) = u for all u ∈ f b \ K(f )b; (ii) η(t, u) = x for all (t, u) ∈ [0, 1] × f a; (iii) η({1} × (f b \ f a)) = f a. For the part II, under the condition (P S)c the proof is due to Rothe [44], Chang [11] and Wang [50]; and under the condition (C)c the proof can be found in Bartsch-Li [5], Perera-Schechter [42] and Perera-Agarwal-O'Regan [41]. Applying these two deformation lemmas and our splitting lemma, Theorem 2.1 in [36], the standard arguments as in [9, 38, 39, 40] may yield the following two theorems. Theorem 2.7 Let H be a Hilbert space and let f : H → R be a F -differentiable and strictly H-differentiable functional. Suppose: (i) for some small ε > 0 there exists a unique critical value c in [c − ε, c + ε], (ii) Kc is finite and f satisfies the conditions of Theorem 2.1 in [36] near each of Kc, (iii) Either f satisfies the (PS) condition on f −1([c−ε, c+ε]) or f is C 1 and satisfies the condition (C)d for every d ∈ [c − ε, c + ε]. 6 Then for any abelian group G one has H∗(fc+ε, fc−ε; G) ∼= Mz∈Kc C∗(f, z), which is finitely dimensional vector spaces over G if G is a field. Let Bm be the closed unit disk in Rm. By a topological embedding h : Bm → H we mean that it is continuous bijection onto h(Bm) ⊂ H and that h is a homeomorphism between Bm and h(Bm) with respect to the induced topology on h(Bm) from H. Theorem 2.8 (Handle Body Theorem). Under the assumptions of Theorem 2.7, if each of Kc = {zj}l 1 is also nondegenerate, then for some 0 < ǫ ≤ ε there exist topological embeddings hi : Bmi → H, i = 1, · · · , l, such that fc−ǫ ∩ hj(Bmj ) = f −1(c − ǫ) ∩ hj(Bmj ) = hj(∂Bmj ) for j = 1, · · · , l, and fc−ǫ ∪Sl is the Morse index of zj. j=1 hj(Bmj ) is a deformation retract of fc+ǫ, where mj Similarly, we can also give the versions on Hilbert manifolds. The following is a slight variant of [24, Prop. 2.1]. Proposition 2.9 Let H be a Hilbert space and let B(∞) : H → H be a bounded self-adjoint linear operator satisfying (C1∞), i.e. 0 is at most an isolated point of the spectrum σ(B(∞)), which implies ±(B(∞)u, u)H ≥ 2α∞kuk2 ∀u ∈ H ± (i) g : H → R is strictly H-differentiable (i.e. locally Lipschitz continuous and strictly ∞. Assume: G-differentiable) and hence ∂g(x) = {g′(x)} by Proposition 2.1; (ii) kg′(x)k is bounded, g′ is compact and ν∞ = dim H 0 (iii) For any M > 0, g′(u0 + u±) → 0 uniformly in u± ∈ ¯BH(θ, M ) ∩ H ± ∞ < ∞; ∞ as ku0k → ∞. Then L(u) = 1 2 (B(∞)u, u)H + g(u) satisfies (PS) condition on H \ CR,M , where CR,M = {u = u0 + u± ku0k > R, ku±k < M }. Consequently, for any (P S)c sequence {un} of L, either {un} has a bounded sub- sequence (and hence a converging subsequence) or c ∈ C∞(L) and there exists a subsequence {unk} such that ku0 nkk → 0 and g(unk ) → c. Here C∞(L) is a closed subset of R given by nk k → ∞, ku± C∞(L) := {c ∈ R ∃ u0 n ∈ H 0 n ∈ H ± ∞, u± ∞ with ku± ∞k → 0 such that ku0 g(u0 nk → ∞, n + u± n ) → c}. Consequently, L satisfies the (P S)c condition for c /∈ C∞. 7 Proof. Let {un} ⊂ H \ CR,M be such that L(un) → c and B(∞)un + g′(un) → 0 as n → ∞. Since kg′(x)k is bounded, and ku± n k we infer that {ku± k → ∞. Then g′(u0 n k ≤ M1 ∀n. Suppose that a subsequence ku0 nk n k} is bounded. Let ku± n k ≤ kB(∞)H ± k · kB(∞)u± ∞ nk + u± nk ) → 0 and so ku± nk k ≤ kB(∞)H ± = kB(∞)H ± ∞ ∞ k · kB(∞)u± nk k · kL′(unk ) − g′(u0 nk k + u± nk )k → 0. nk} is bounded. Since g′ is compact and ν∞ = dim H 0 nk ∈ CR,M for k large enough. This contradiction shows that ∞ < ∞ we have a subse- nk + u± Hence unk = u0 {ku0 quence {unk } such that u0 nk → u0 and g′(unk ) → v. The latter implies that ku± nk − u± nlk ≤ kB(∞)H ± = kB(∞)H ± ∞ ∞k · kB(∞)u± nk − B(∞)u± nlk k · kL′(unk ) − L′(unl) − [g′(unk ) − g′(unl)]k → 0 as k, l → ∞. Hence {unk } converges to some v. ✷ 3 Computations of critical groups 3.1 Critical groups at infinity and computations In this subsection K always denotes a commutative ring without special statements. For a strictly H-differentiable functional f : X → R on a Banach space X, suppose that the set of critical values of f is strictly bounded from below by a ∈ R, and for all c ≤ a that f satisfies the condition (C)c. By Corollary 2.4(i), for every nonnegative integer m, Cm(f, ∞; K) := Hm(X, f a; K), C m(f, ∞; K) := H m(X, f a; K) (3.1) (3.2) are independent of the choices of such a, and are called the mth critical group of f at infinity and mth cohomological critical group of f at infinity, respectively (cf. Definition 3.4 of [5]). Here H∗(−; K) and H ∗(−; K) denote the singular homology and cohomology with coefficients in K. It is well-known that Cm(f, ∞; K) = Hm(X, f a; K) ∼= Hom(Hm(X, f a; K)) ∼= H m(X, f a; K) = C m(f, ∞; K) if K is a field. Let ¯H ∗(−; K) denote Alexander-Spanier cohomology with coefficients in K, which has often some stronger excision and continuity properties. Now the Banach space X is a ANR (absolute neighborhood retract). By Section K on the page 30 of [25], every open subset of an ANR an ANR, and Hanner theorem claims that a metrizable space is an ANR if it has a countable open covers consisting of 8 ANR. Hence f a = ∪∞ follows that H m(X, f a; K) ∼= ¯H m(X, f a; K) for any field K. In particular we have n } is an ARN. From Section 9 of [48, Chapter 6] it n=1{f < a − 1 Cm(f, ∞; K) ∼= C m(f, ∞; K) ∼= ¯H m(X, f a; K) (3.3) for any field K and nonnegative integer m. These and Proposition 3.15 of [41] lead to Proposition 3.1 For a strictly H-differentiable functional f : X → R on a Banach space X, suppose that the set of critical values of f is strictly bounded from below by a ∈ R, and that f satisfies the condition (C)c for all c ∈ R. Then for any field K and nonnegative integer m it holds (i) Cm(f, ∞; K) ∼= C m(f, ∞; K) ∼= δm0K if f is bounded from below. (ii) Cm(f, ∞; K) ∼= C m(f, ∞; K) ∼= eH m−1(f a; K) if f is unbounded from below. Here eH 0(f a; K) = H 0(f a; K)/K and eH q(f a; K) = H q(f a; K) for q ≥ 1. By Proposition B.1 of [36], the continuously directional differentiability is stronger than the strict H-differentiability. We have the following generalization of Theo- rem 3.9 in [5]. Theorem 3.2 Suppose for V∞ = H: (i) the assumptions of Theorem 4.1 of [36], (S), (F1∞)-(F3∞) and (C1∞)-(C2∞), (D∞) and (E′ ∞), are satisfied; (ii) L(u) = 1 2 (B(∞)u, u)H + o(kuk2) as kuk → ∞; (iii) ∇L(u) = B(∞)u + o(kuk) as kuk → ∞, where ∇L is the gradient of L defined by dL(u)(v) = (∇L(u), v)H for all u, v ∈ H; (Note: we do not assume L ∈ C 1(H, R).) 1 (iv) the critical values of L are bounded below; (v) L satisfies the condition (C)c (or (D)c) for c ≪ 0. Then Ck(L, ∞; K) = 0 for k ∈ [µ∞, µ∞ +ν∞] even if µ∞ = ∞ or ν∞ = ∞. Moreover, if µ∞ < ∞ and ν∞ = 0 then Cµ∞(L, ∞; K) ∼= K (even if H is not complete). Proof. Step 1. Carefully checking the proof of Lemma 4.2 in [5] one easily sees that the conditions (i) and (ii) imply: for sufficiently large R > 0 and a ≪ 0 the pair ∞ (θ, R + 1) ⊕ H ± ∞, La ∩ (BH 0 ∞ (θ, R + 1) ⊕ H ± ∞)(cid:1) (θ, 1)(cid:1). is homotopy to the pair (cid:0)BH 0 (cid:0)BH 0 ∞(θ, R + 1) ⊕ ¯BH − ∞ (θ, 1), BH 0 ∞ (θ, R + 1) ⊕ ∂ ¯BH − ∞ The homotopy equivalence leaves the H 0 ∞-component fixed. 1For a possible method removing this condition, see below the end of this document. 9 Step 2. Under the assumption (i), by Theorem 4.1 of [36] we can get Lemma 4.3 of [5]: There exist a sufficiently large R > 0, a ≪ 0 and a continuous map γ : ∞ (∞, R) such BH 0 that the pair ∞(∞, R) → [0, 1] with γ(C) > 0 for C := BH 0 ∞(θ, R + 1) ∩ BH 0 ∞(∞, R) × H ± is homotopy equivalent to the pair (BH 0 (BH 0 ∞, La ∩ (BH 0 ∞ (∞, R) × H − ∞ (∞, R) × H ± ∞, Γ), where ∞)) Γ = {(z, u) ∈ BH 0 ∞(∞, R) × H − ∞ : kuk ≥ γ(z)}quadand if L(z + h∞(z)) ≤ a, if L(z + h∞(z)) ≥ a + 1, 0 1 L(z + h∞(z)) − a elsewhere. γ(z) = Moreover, the homotopy equivalence leaves the H 0 ∞-component fixed. Step 3. By the assumptions (iv) and (v), C∗(L, ∞; K) = H∗(H, La; K) for a ≪ 0 is well-defined. Using Step 1 and Step 2 we may repeat the proof on the pages 428-429 of [5] to obtain at the desired conclusion. ✷ Using Corollary 2.4 we derive (i) and (ii) the following proposition, which are corresponding with Propositions 3.5 and 3.6 in [5]. Proposition 3.3 For a strictly H-differentiable functional f : X → R on a Banach space X, we have (i) If a < inf f (K(f )) ≤ sup f (K(f )) < b and f satisfies the condition (C)c (or (D)c) for any c /∈ (a, b), then C∗(f, ∞; K) ∼= H∗(f b, f a; K). (ii) If f satisfies the condition (C)c (or (D)c) for any c ∈ R, then C∗(f, ∞; K) ∼= 0 in the case K(f ) = ∅, and C∗(f, ∞; K) ∼= C∗(f, x0; K) in the case K(f ) = {x0}. (iii) If f is F-differentiable, satisfies the condition (P S)c for every c ∈ R and has finite critical points, then for every field K it holds that dim Cm(f, ∞; K) ≤ Xu∈K(f ) dim Cm(f, u; K) ∀m ∈ N ∪ {0}. When f is C 1 and satisfies the condition (C)c for every c ∈ R, (iii) was proved in Proposition 3.16 of [41]. Proof of Proposition 3.3. We only prove (iii). It is almost standard (see [16] and [17, Remark 2]). For the reader's convenience we prove it. Since f has only finite critical points we may take −∞ < a < inf f (K(f )). Let c1 < · · · < ck be all critical values. Take numbers a < a1 < · · · < ak such that a1 < c1 < a2 < c2 < · · · < ak < ck. Take b = +∞ then f b◦ = X, and f a1◦ ⊂ f a2◦ ⊂ · · · ⊂ f ak◦ ⊂ f b◦. By [41, Lemma 3.12] we get for every nonnegative integer m that dim ¯H m(f b◦, f a1◦; K) ≤ kXi=1 10 dim ¯H m(f ai+1◦, f ai◦; K), where ak+1 = b. From the proof of Theorem 4 in [17] we may see that f a1◦ is a strong deformation retract of f b. This implies ¯H m(f b◦, f a1◦; K) ∼= ¯H m(X, f a; K), and so dim ¯H m(f b◦, f a1◦; K) = dim Cm(f, ∞; K) ∀m. Now the part I of Lemma 2.6 yields ¯H m(f ai+1◦, f ai◦; K) ∼= ¯H m(f ai+1◦, f ci◦; K) ∼= ¯H m(f ci◦ ∪ K(f )ci, f ci◦; K) for any i = 1, · · · , k and nonnegative integer m. As showed in [17, Remark 2] the subsets f ci◦ ∪ K(f )ci and f ci◦ are ARN. Hence ¯H m(f ci◦ ∪ K(f )ci, f ci◦; K) ∼= H m(f ci◦ ∪ K(f )ci, f ci◦; K) ∀m, i. Let K(f )ci = {xi1, · · · , xili}, and Ui1, · · · , Uili be mutually disjoint open neighbor- hoods of xi1, · · · , xili, respectively. Then the excision property of singular cohomology groups lead to H m(f ci◦ ∪ K(f )ci, f ci◦; K) ∼= liMj=1 H m(cid:0)(f ci◦ ∪ {xij}) ∩ Uij, f ci◦ ∩ Uij; K) for all m, i. Moreover, it is easy to prove H m(cid:0)(f ci◦ ∪ {xij}) ∩ Uij, f ci◦ ∩ Uij; K) ∼= H m(cid:0)f ci ∩ Uij, (f ci \ {xij}) ∩ Uij; K) ∼= Hm(cid:0)f ci ∩ Uij, (f ci \ {xij}) ∩ Uij; K) ∼= Cm(f, xij; K). The desirable conclusion follows from these immediately. ✷ We can also give generalizations of some computation results on critical groups at infinity such as Proposition 3.10 in [5] and some parts of [31]. We leave them intersecting readers. 3.2 Computations of critical groups at degenerate criti- cal points Definition 3.4 Let X be a Banach space and let f : X → R be a strictly H- differentiable functional. For subsets S ⊂ X and A ⊂ X ∗ we call f to satisfy (P S) condition with respect to A on S if every sequence (xn) in S with (f (xn)) bounded and f ′(xn) → y ∈ A has a convergent subsequence. Clearly, the usual (P S) condition is the (P S) condition with respect to {0} on X. We have the following generalization of Proposition 2.5 in [5]. Proposition 3.5 (i) Let X ⊂ H be as in (S) of [36, §2.1], and let L be a continu- ously directional differentiable functional defined in an open neighborhood V of x0 ∈ X in H; moreover we assume that the conditions (F1)-(F3), (C1)-(C2) and (D) in [36, §2.1] hold with θ replaced by x0. 11 (ii) x0 is an isolated critical point of L, and L is F -differentiable near x0. (iii) ∇L(u) = B(x0)(u − x0) + o(ku − x0k) as ku − x0k → 0, where ∇L is the gradient of L defined by dL(u)(v) = (∇L(u), v)H for all u, v ∈ H. (Note: we do not assume L ∈ C 1(H, R).) (iv) L satisfies the (PS) condition with respect to H 0 on a closed ball ¯BH(x0, δ). Let µ0 and ν0 be the Morse index and nullity of L at x0. Then we have: (a) Ck(L, x0; K) = δkµ0K provided that L also satisfies: (AC+) There exist ε > 0 and θ ∈ (0, π/2) such that (∇L(u + x0), u0)H ≥ 0 for any u = u0 + u± ∈ H 0 + H ± with kuk ≤ ε and ku±k ≤ kuk · sin θ. (b) Ck(L, x0; K) = δk(µ0+ν0)K provided that L also satisfies: (AC−) There exist ε > 0 and θ ∈ (0, π/2) such that (∇L(u + x0), u0)H ≤ 0 for any u = u0 + u± ∈ H 0 + H ± with kuk ≤ ε and ku±k ≤ kuk · sin θ. By Corollary 2.6 of [36] we have Cq(L, x0; K) = 0 if q /∈ [µ0, µ0 + ν0]. So Proposi- tion 3.5 may be viewed a refinement of this result. For the proof of it we also need the following stability property of critical groups for continuous functionals by Cingolani and Degiovanni [13], which is a very general generalization of the previous results due to Chang [9, page 53, Th.5.6], Chang and Ghoussoub [12] and in Mawhin and Willem [38, Th.8.8], and Corvellec and Hantoute [19]. Theorem 3.6 ([13, Th.3.6]) Let {ft : t ∈ [0, 1]} be a family of continuous functions from a metric space X to R, let U be an open subset of X and [0, 1] ∋ t 7→ ut ∈ U a continuous map. Assume: (I) if tk → t in [0, 1], then ftk → ft uniformly on U ; (II) U is complete, and for every sequence tk → t in [0, 1] and (vk) in U with dftk (tk) → 0 and (ftk (vk)) bounded, there exists a subsequence (vkj ) convergent to some v with dft(v) = 0; (III) dft(v) > 0 for every t ∈ [0, 1] and v ∈ U \ {ut} Then Cq(f0, u0; K) ∼= Cq(f1, u1; K) for every q ≥ 0. Proof of Proposition 3.5. Following the proof ideas of Proposition 2.5 in [5], we assume x0 = θ. For the case (a) (resp. (b)) we set Lt(u) = L(u) + 1 2 tku0k2 (resp. Lt(u) = L(u) − 1 2 tku0k2) for t ∈ [0, 1]. Using the assumptions (ii) and (iv), as in the proof of Proposition 2.5 in [5] we have a small ǫ ∈ (0, 2/δ) such that θ is the only critical point of each Lt in BH (θ, 2ǫ). Clearly, L1 also satisfies the assumption of Theorem 2.1 of [36], and θ is a nondegenerate critical point of L1 with Morse index µ0 (resp. µ0 + ν0) in the case (a) (resp. (b)). It follows from Corollary 2.6 of [36] that Ck(L, θ) = δkµ0 K in case (a) (resp. Ck(L, θ) = δk(µ0+ν0)K in case (b)). (3.4) The remaining is to prove Cq(L, θ) ∼= Cq(L1, θ) for every q ≥ 0 in both cases. 12 We only prove the case (a). Since L is continuously directional differentiable, and F -differentiable at θ, so is each Lt. By Proposition B.2(ii) of [36] and Proposition 2.1 every Lt is strictly H-differentiable (and thus locally Lipschitz continuous), and ∂Lt(u) = {L′ t(u)} and dLt(u) = kL′ t(u)k ∀u ∈ ¯BH(θ, ǫ) (3.5) and dLt(u) > 0 ∀u ∈ ¯BH(θ, ǫ)\{θ} because θ is only critical point of Lt in BH(θ, 2ǫ). The second equality in (3.5) implies that θ is also a lower critical point of each Lt. Because of (3.4), we hope to use Theorem 3.6 proving that Cq(L, θ) ∼= Cq(L1, θ) ∀q ≥ 0. It suffice to check the conditions of Theorem 3.6. Clearly, Ltk → Lt uniformly on ¯BH (θ, ǫ) as tk → t in [0, 1]. Now we assume: tk → t in [0, 1], (uk) ⊂ ¯BH (θ, ǫ) is such that (Ltk (uk)) is bounded and dLtk (uk) → 0. These imply that L′ (uk) = tk k → u0 (passing a subsequence L′(uk)+tku0 if necessary). Then (L(uk)) is bounded and L′(uk) → −tu0 ∈ H 0. By the assumption (iv) (uk) has a convergent subsequence uki → u0 ∈ ¯BH(θ, ǫ). Hence u0 = P 0u0. Moreover, since L is continuously directional differentiable we get k → θ. Since dim H 0 < ∞ we may assume u0 (L′(uki), v)H → (L′(u0), v)H ∀v ∈ H. Hence (−tu0, v)H = (L′(u0), v)H ∀v ∈ H, i.e. L′(u0) + tP 0u0 = L′(u0) + tu0 = θ. It follows from (3.5) that dLt(u0) = kLt(u0)k = 0. Namely {Lt t ∈ [0, 1]} satisfies the conditions of Theorem 3.6 on ¯BH(θ, ǫ). ✷ 4 Morse inequalities and some critical point theorems 4.1 Morse inequalities In this subsection, unless otherwise specified, let the functional L : H → R be as in Proposition 2.9. Then L satisfies the (P S)c condition for each c /∈ C∞(L). We also assume that L is C 1 so that it has a pseudo-gradient vector field, V : eH → H. Note that eH → H, u 7→ V (u)/kV (u)k is also a locally Lipschitz continuous map. Consider the flow η(t, u) = − V (η(t, u)) kV (η(t, u))k and η(t, 0) = u. (4.1) Our goal is to present the Morse inequality established in [5, 24] under the above weaker setting. For F ⊂ H let eF = ∪t∈Rη(t, F ). For any isolated value c in C∞(L), let Lc+ε K(L) ∩ Lc+ε c−ε. Define c−ε := L−1([c − ε, c + ε]) and K c+ε c−ε (L) := UR,M = {u = u0 + u± ku0k ≤ R} ∪ {u = u0 + u± ku0k > R, ku±k ≥ M }, CR,M = {u = u0 + u± ku0k > R, ku±k < M } = H \ UR,M , R,M = UR,M ∩ Lc+ε, C c+ε U c+ε R,M = U c+ε Ac+ε R,M = CR,M ∩ Lc+ε, 2R,M/2 ∩ C c+ε R,M . 13 The following lemma corresponds to Lemma 2.1, Proposition 2.2 and Corollary 2.4 in [24] (see also Lemma 2.7 and Theorem 2.9 in [30]). Lemma 4.1 Assume that K(L)c+ε0 c−ε0 = K(L)c is compact for some ε0 > 0. Then for R large and R > M > 0 with K(L)c ⊂ BH(θ, R/2) ∪ C3R,M/8 there exists ε1 > 0 such that for any ε ∈ (0, ε1) it holds that ^ C c+ε ^ U c+ε c−ε ∩ (i) (cid:0)Lc+ε (ii) Lc+ε ∩ R,M(cid:1) ∩(cid:0)Lc+ε ∼= Lc−ε ∩ ^ Ac+ε R,M c−ε ∩ ^ Ac+ε R,M . 2R,M/4(cid:1) = ∅, Furthermore, if K(L)c is compact then for any M > 0 there exist a large R > 0, and ε1 > 0 such that for any ε ∈ (0, ε1), Hq(Lc+ε, Lc−ε; K) ∼= Hq(Lc+ε ∩ ⊕ Hq(Lc+ε ∩ ^ 2R,M/2, Lc−ε ∩ U c+ε ^ R,M , Lc−ε ∩ C c+ε ^ C c+ε ^ U c+ε 2R,M/2; K) R,M ; K) ∀q = 0, 1, · · · . Hereafter K always denotes a commutative ring without special statements. Proof. Since U3R,M/8 \ BH(θ, R/2) = H \(cid:0)BH (θ, R/2) ∪ C3R,M/8(cid:1) is disjoint with c−ε0 = K(L)c, and L satisfies the (P S) condition in U3R,M/8 by Proposition 2.9, K(L)c+ε0 there exists an ε′ > 0 such that kL′(u)k ≥ kL′(u)k ≥ ε′ ∀u ∈ Lc+ε0 c−ε0 ∩(cid:0)U3R,M/8 \ BH(θ, R/2)(cid:1). For 0 < ε < min{ε0, ε′M/16}, suppose that η(s, x) ∈ Lc+ε c−ε ∩ C c+ε R+M/4,3M/4 for some x ∈ Lc+ε c−ε ∩ U c+ε R,M and s > 0. Then there exists t1 < t2 such that η(t1, x) ∈ Lc+ε η(t, x) ∈ ¯CR,M ∩ UR+M/4,3M/4 ∀t ∈ [t1, t2]. η(t2, x) ∈ Lc+ε c−ε ∩ ∂UR,M , c−ε ∩ ∂CR+M/4,3M/4 ⊂ Lc+ε c−ε ∩ ∂U2R,M/2, Then M/4 ≤ kη(t1, x) − η(t2, x)k ≤ t2 − t1, and L(η(t2, x)) − L(η(t1, x)) = Z t2 = −Z t2 t1 d dt L(η(t, x))dt hL′(η(t, x)), V (η(t, x))i kV (η(t, x))k dt t1 ε′ 2 ≤ − t2 − t1. Hence ε′ M 4 This contradicts to the choice of ε. So we get ≤ ε′t2 − t1 ≤ 2(cid:0)L(η(t1, x)) − L(η(t2, x))(cid:1) ≤ 4ε. c−ε ∩ UR,M(cid:1) ∩(cid:0)Lc+ε ^(cid:0)Lc+ε c−ε ∩ CR+M/4,3M/4(cid:1) = ∅. 14 Since(cid:0)Lc+ε c−ε ∩ ^ U c+ε c−ε ∩ ^ U c+ε R,M(cid:1) ⊂ ^(cid:0)Lc+ε c−ε ∩ UR,M(cid:1), (cid:0)Lc+ε R,M(cid:1) ∩(cid:0)Lc+ε (cid:0)Lc+ε 2R,M/4(cid:1) ∩(cid:0)Lc+ε ^ C c+ε c−ε ∩ c−ε ∩ CR+M/4,3M/4(cid:1) = ∅. c−ε ∩ U2R−M/4,M/2(cid:1) = ∅. Similarly, for 0 < ε < min{ε0, ε′M/16} we have (4.2) This and (4.2) together give (i). As in the proof of [24, 30] we can get the remaining conclusions. ✷ Following [5, 24], as in [9] and [38]) Lemma 4.1 may lead to Theorem 4.2 Under the assumptions of Proposition 2.9, let L be C 1, and let K(L) and C∞(L) be finite. Denote by βk(f, x) := dim Ck(f, x; K) for x ∈ K(L), and by βk(L, c) = dim Hk(Lc+ε ∩ eCR,M , Lc−ε ∩ eCR,M ; K) ∀c ∈ C∞(L), dim Hk(H, La; K)tk P (L, ∞) := ∞Xk=0 for any a < min(cid:8)x x ∈ L(K(L)) ∪ C∞(L) ∪ {0}(cid:9). Then there exists a polynomial Q(t) with nonnegative integer coefficients such that P (L, ∞) + (1 + t)Q(t) = Xx∈K(f ) P (L, x) + Xv∈C∞(f ) P (L, c), where P (L, x) :=P∞ k=0 βk(L, x)tk and P (L, c) :=P∞ k=0 βk(L, c)tk 4.2 Some critical point theorems Many critical point theorems, which were obtained by computations of critical groups, can be generalized with our methods. For example, the following is a generalization of Theorem 5.1 on the page 121 of [9]. Theorem 4.3 (I) Let the Banach space (X, k · k) and the Hilbert space (H, (·, ·)H ) Φp : Up → Φp(Up) ⊂ H with Φp(p) = θ, such that it restricts to a coordinate chart p : Up ∩ eX → Φp(Up ∩ eX) ⊂ X. satisfy the condition (S) in [36, §2.1]. Let eH (resp. eX) be a C 1 Hilbert (resp. C 2 Banach) manifold modeled on H (resp. X). Suppose that eX ⊂ eH is dense in eH, and that for each point p ∈ eX there exists a coordinate chart around p on eH, around p on eH, ΦX (II) Let L : eH → R be a continuously directional differentiable, and F -differentiable (a) L satisfies the (PS) condition, and restricts to a C 2-functional on eX; (c) ∃ a finite set {p1, · · · , pm} ⊂ eX is contained in K(L) ∩ L−1[a, b]; (b) rankHk(Lb, La; K) 6= 0 for some k ∈ N and regular values a < b; functional with the following properties. 15 (d) Around each pi there exists a chart Φpi : Upi → Φpi(Upi) ⊂ H as in (I) such that the functional L◦◦(Φpi)−1 : Φpi(Upi) → R satisfies the conditions of Theorem 2.1 of [36]; so pi has Morse index µi and nullity νi; (f ) Either µi > k or µi + νi < k, i = 1, · · · , m. Then L has at least one more critical point p0 with rankCk(L, p0; K) 6= 0. Proof. By the condition (II), the lower critical point set of L coincides with K(L), and L satisfies the (PS) condition for continuous functionals. If the conclusion is not true then K(L) = {p1, · · · , pm}. By Corollary 2.6 of [36] and (f) we have Ck(L, pi) = 0, i = 1, · · · , m. As in the proof of Theorem 5.1 on [9, page 121] we may use Corvellec's Morse theory for continuous functionals [16] to obtain a contradiction. ✷ Similarly, a suitable weaker version of Theorem 5.4 on [9, page 121] may be given. In particular, by Theorem 2.12 of [36] we have the following generalization of Corollary 5.3 therein. Theorem 4.4 Under the assumptions of Theorem 2.1 of [36], suppose V = H and the following conditions hold: (i) L is bounded below, F -differentiable and satisfies the (PS) condition; (ii) For a small ǫ > 0, either degBS(∇L, BH (θ, ǫ), θ) = ±1, or degFPR(A, BX (θ, ǫ), θ) = ±1 provided that the map A in the condition (F2) is C 1 near θ ∈ X, where the degrees degBS and degFPR are as in Theorem 2.12 of [36]. Then L has at least three critical points. Finally, we give a generalization of Theorem 3.12 in [5]. To this goal we also need the following results, which are Propositions 2.3 and 3.8 in [5]. Proposition 4.5 Let a normed vector space X have a direct sum decomposition X = X1 ⊕ X2, where k = dim X2 < ∞. For f ∈ C(X, R) we have: (i) If there exist x0 ∈ X and ε > 0 such that f (x0 +x) > f (x0) ∀x ∈ H1, 0 < kxk ≤ ε, and that f (x0 + x) ≤ f (x0) ∀x ∈ X2, kxk ≤ ε, then Ck(f, x0) 6= 0. ([32] ) (ii) If f is bounded from below on X1, and f (x) → −∞ for x ∈ X2 as kxk → ∞, then Hk(X, f a) 6= 0 for a < inf f X1. ([5, Proposition 3.8]) Theorem 4.6 Under the assumptions (i)-(iv) of Theorem 3.2, let the condition (i) of Proposition 3.5 be satisfied. (Of course the corresponding densely imbedded Banach spaces in H are not necessarily same, we denote by A∞ and B∞ the corresponding maps in (i) of Theorem 3.2). Let H split as H = H 0 ⊕ H + ⊕ H − (resp. H = H 0 ∞) according to the spectral decomposition of B(x0) (resp. B∞(∞)). Let µ0, ν0 (resp. µ∞, ν∞) be the Morse index and nullity at x0 (resp. infinity). ∞ ⊕ H − ∞ ⊕ H + (I) If (v) of Theorem 3.2 and the local linking condition as in Proposition 4.5(i) with X − = H − (resp. X − = H 0 ⊕ H −) hold, then there exists a critical point different from x0 provided µ0 /∈ [µ∞, µ∞ + ν∞] (resp. µ0 + ν0 /∈ [µ∞, µ∞ + ν∞]). 16 (II) If (ii)-(iv) and (AC+) of Proposition 3.5 hold, and L is bounded from below on H 0 ∞ ⊕ H + ∞, then there exists a nontrivial critical point provided µ∞ 6= µ0. Proof. (I) Otherwise we have K(L) = {x0}. By Proposition 3.3(ii), C∗(L, ∞) ∼= C∗(L, x0). Moreover, Proposition 4.5(i) implies Cµ0(L, x0) 6= 0. Hence Cµ0(L, ∞) 6= 0. This gives a contradiction by Theorem 3.2. (II) Similarly, assume K(L) = {x0}. By Proposition 3.3(ii), C∗(L, ∞) ∼= C∗(L, x0). Proposition 3.5(a) yields Ck(L, x0) = δkµ0K. By Proposition 4.5(ii), Cµ∞(L, ∞) 6= 0. This contradiction proves the desired conclusion. ✷ . 5 Critical groups of sign-changing critical points In this section we shall present the corresponding version of the results on critical groups of sign-changing critical points in [4] and [34] in our framework and sketch how to prove them with our results in the previous sections. It is also possible to generalize some of [2]. They are left to the interested reader. Let the Hilbert space (H, (·, ·)H ) and the Banach space (X, k · kX ) satisfy the △ = PH ∩ X, condition (S) as in [36, §2.1]. Let PH be a closed convex in H so that P a closed convex cone in X, satisfies: (i) intX(P ) 6= ∅, (ii) ∃ e ∈ intX(P ) with (u, e)H > 0 for all u ∈ P \ {θ}. (5.1) Then H (resp. X) is partially ordered by by PH (resp. P ). For u, v ∈ H (resp. X) we write: u ≥ v if u − v ∈ PH (resp. u − v ∈ P ); u > v if u − v ∈ PH \ {θ} (resp. u − v ∈ P \ {θ}). When u, v ∈ X we also write u ≫ v if u − v ∈ intX(P ). A map f : H → H (resp. f : X → X) is called order preserving if u ≥ v ⇒ f (u) ≥ f (v) ∀u, v ∈ H (resp. X). In particular, f : X → X is said to be strongly order preserving if u > v ⇒ f (u) ≫ f (v) ∀u, v ∈ X. Recall that above Proposition 2.1 we have showed that a continuously directional differentiable map is locally Lipschitz continuous and strictly G-differentiable. Let the assumptions of Theorem 2.1 of [36] hold for V = H. We also assume: (L0) The assumptions of Proposition 2.26 of [36] hold for V = H, that is, the as- sumptions of Theorem 2.1 of [36] hold for V = H, the map A : X → X in the condition (F2) of [36] is Fr´echet differentiable, and there exist positive constants 0 and C ′ η′ 1 such that 2 > C ′ C ′ 2kuk2 ≥ (P (x)u, u) ≥ C ′ 1kuk2 ∀u ∈ H, ∀x ∈ BH(θ, η′ 0) ∩ X. (L1) L : H → R is C 2−0, and satisfies the (P S)c condition for any c ∈ R. Moreover, all critical points of L are contained in X, and H and X induce an equivalent topology on the critical set of L 2 2The final assumption is used in the proof of Claim 3, cf. the proof of Theorem 5.1. There is no such a assumption in [4]. 17 (L2) The map idX − A : X → X is strongly order preserving. (L3) The smallest eigenvalue of B(θ), which is equal to inf kuk=1(B(θ)u, u)H by [6, Prop.6.9] and Proposition B.2 of [36], is simple and its eigenspace (contained in X by (D1)) is spanned by a positive eigenvector (i.e. sitting in IntX(P )). (L4) One of the following holds: (i) L is bounded below; (ii) For every u ∈ H \ {θ} it holds that L(tu) → −∞ as t → ∞. Moreover, there exists ♭ < 0 such that L(u) ≤ ♭ implies DL(u)(u) < 0; (iii) There exist a compact self-adjoint linear operator Q∞ ∈ L(H) and a posi- tive definite operator P∞ ∈ L(H) such that ∇L(u) = P∞u − Q∞u + o(kuk) as kuk → ∞, where ∇L is the gradient of L. The smallest eigenvalue of B∞ := P∞ − Q∞ is simple and its eigenspace is spanned by a positive eigenvector e∞ ∈ IntX (P ) such that (u, e∞)H > 0 for every u ∈ P \ {θ}. Moreover, B∞X ∈ L(X), and if a subset S ⊂ X is bounded in H then A(S) is also bounded in H. Let µ0 := dim H − and ν0 := dim H 0. Define µ∞ = ν∞ = 0 in the case (i) of (L4), and µ∞ = ∞ and ν∞ = 0 in the case (ii) of (L4). For the case (iii) of (L4), let µ∞ be the number of negative eigenvalues of B∞ (counted with multiplicities) and ν∞ = dim Ker(B∞). An element x ∈ X is called a subcritical (resp. supercritical) critical point of L if ∇L(x) ≤ 0 (resp. ∇L(x) ≥ 0). By (F2) the map A : X → X is continuously directional differentiable. It follows from Proposition B.1(ii) of [36] that A is a locally Lipschitz map from X to X. Note that ∇L is C 1−0, and equal to A on X by (L2), i.e. ∇L(x) = A(x) ∀x ∈ X. Consider the negative gradient flow ϕ(t, x) of L on H defined by d dt ϕ(t, x) = −∇L(ϕ(t, x)) and ϕ(0, x) = x ∈ H. (5.2) Let ϕt(·) = ϕ(t, ·). As in [4] it restricts to a continuous local flow on X, still denoted by ϕt(·), and t 7→ L(ϕt(x)) is strictly decreasing for any x /∈ K(L). Let D = P ∪ (−P ) and D∗ = D \ {θ}, and let Ld = {L(x) ≤ d} for d ∈ R. The following result is a generalization of Theorem 3.4 in [4]. Theorem 5.1 Under the above assumptions (L0)-(L4), if µ0 ≥ 2 and µ∞ + ν∞ ≤ 1 then L has a sign-changing critical point x1 with L(x1) < 0. If all sign changing critical points (i.e. those in X \D ) with critical values contained in a bounded interval of (−∞, 0) are isolated, then there exists a sign changing critical point x1 with L(x1) < 0 and C1(L, x1; K) 6= 0. Furthermore, if near x1 the conditions of Theorem 2.1 of [36] hold and x1 has nullity 1 then x1 is of mountain type and Ck(L, x1; K) ∼= δk1F. (Note: It is sufficient that L satisfies the (P S)c condition for each c < 0). Proof. Firstly, we prove: 18 Claim 1. Suppose that all sign changing critical points with critical values contained in a bounded interval of (−∞, 0) are isolated. Then for any interval [a, b] ⊂ (−∞, 0) there exist only finitely many sign changing critical points with critical values in [a, b]. Otherwise, let {un} be infinite such points with {L(un)} ⊂ [a, b] ⊂ (−∞, 0). We may assume L(un) → c (by passing to a subsequence if necessary). By the (P S)c condition we may assume un → u0 in H, and un → u0 in X because of the final assumption in (L1). Clearly, u0 6= θ because of L(u0) ≤ b < 0. Since each un sits in an open subset H \(PH ∪(−PH)), it belongs to X∩(H \(PH ∪(−PH))) = X\(P ∪(−P )) as well. Then either u0 ∈ X \ (P ∪ (−P )) or u0 ∈ ∂D \ {θ}. In the first case u0 also sits in H \ (PH ∪ (−PH )). This contradicts the assumption that all sign changing critical points with negative critical values are isolated. In the latter case either u0 ∈ ∂P \ {θ} or u0 ∈ (−∂P ) \ {θ}. Note that idX − A is strongly order preserving. If u0 ∈ ∂P \ {θ} then u0 = u0 − A(u0) ≫ θ − A(θ) = θ, i.e. u0 ∈ IntX (P ). This leads to a contradiction. If u0 ∈ ∂(−P ) \ {θ}, i.e. −u0 ∈ ∂P \ {θ} then θ = u0 + (−u0) − A(u0 + (−u0)) ≫ u0 − A(u0) = u0 because u0 + (−u0) ≥ u0. Hence −u0 ∈ IntX(P ), which yields a contradiction again. Claim 1 is proved. Next, following the methods of proof of [4, Th.3.4], since idX − A : X → X is strongly order preserving by (L2) we get that ϕt(x) ∈ IntX (D) for all x ∈ D∗ and t > 0. For d ∈ R let (jd)1 be the homomorphism from H1(X, D∗; K) to H1(X, Ld ∪D∗; K) induced by the inclusion jd : (X, D∗) ֒→ (X, Ld ∪ D∗). Let Γ := {d ∈ R (jd)1 6= 0} and c := sup Γ. Since µ∞ + ν∞ ≤ 1, either (i) of (L4) or (iii) of (L4) occurs. Let v∞ = e/kek in case (i), and v∞ = e∞/ke∞k in case (iii). We conclude: Claim 2. L is bounded below on H1 := H ∩ hv∞i⊥. We only need to prove this in the latter case. Since the smallest eigenvalue of B∞, given by inf kuk=1(B∞u, u)H = (B∞v∞, v∞)H , is simple, we have λ2 := inf{(B∞u, u)H kuk = 1, u ∈ H1} > λ1 := (B∞v∞, v∞)H . Obverse that e∞ ∈ X implies that X1 := X ∩ hv∞i⊥ is dense in H1. Since µ∞ + ν∞ ≤ 1 we have three cases: (a) µ∞ = 0 = ν∞, (b) µ∞ = 0 and ν∞ = 1, (c) µ∞ = 1 and ν∞ = 0. They corresponds to λ1 > 0, λ1 = 0, and λ1 < 0 but λ2 > 0, respectively. So we always have λ2 > 0. Take a large N > 0 so that k∇L(u)− B∞uk < λ2 4 kuk as kuk ≥ N . For any u ∈ H1 in H. So there exists a M > 0 such that for any u ∈ X ∩ BH(θ, N ), with kuk > N let ¯u = N · u/kuk. By the assumption A(cid:0)X ∩ BH(θ, N )(cid:1) is bounded (A(tu), u)H dt(cid:12)(cid:12)(cid:12)(cid:12) ≤ M N. L(u) = L(u) − L(θ) =(cid:12)(cid:12)(cid:12)(cid:12)Z 1 DL(tu)(u)dt(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)Z 1 0 0 19 This also holds for all u ∈ BH(θ, N ) because X ∩ BH(θ, N ) is dense in BH(θ, N ). On the other hand (∇L(tu + (1 − t)¯u), (u − ¯u))H dt (B∞(tu + (1 − t)¯u), u − ¯u)Hdt − λ2 4 ku − ¯uk · ktu + (1 − t)¯uk L(u) − L(¯u) = Z 1 ≥ Z 1 0 0 1 2 λ2 2 ≥ ≥ (B∞(u − ¯u), u − ¯u)H + (B∞(¯u), u − ¯u)H − ku − ¯uk2 + (B∞(¯u), u − ¯u)H − λ2 4 ku − ¯uk · k¯uk. λ2 4 ku − ¯uk · k¯uk Claim 2 follows immediately. This implies that any d < inf LX1 = inf LH1 belongs to Γ. Hence c is finite. Since µ0 ≥ 2 the two smallest eigenvalues of B(θ) are negatives and the corresponding eigenspaces are contained in X by (D1). Let e1 and e2 two normalized eigenvectors belonging to the two smallest eigenvalues of B(θ). By (L3), e1 ∈ IntX (P ) ⊂ IntX(D). Let Sρ be the sphere of radius ρ in Span{e1, e2}. It easily follows from Proposition 2.26 of [36] that max L(Sρ) < 0. 3 As in the proof of [4, Lem.4.2] we can prove c < 0. Claim 3. negative critical value. c is a critical value of L, and hence L has a sign-changing critical point with This may be proved by a standard deformation argument. In view of Claim 1 let us suppose that there exist only finitely many sign changing critical points x1, · · · , xq at the level c. Note that Claim 1 also implies that there exist a η > 0 such that no number in [c − η, c + η] \ {c} is a critical value of sign changing critical points. Repeating the remainder of proof of [4, Th.3.4] we get some i ∈ {1, · · · , q} such that C1(L, xi; K) 6= 0. The final conclusions follow from Corollary 2.9(ii) of [36]. ✷ Corresponding with Theorem 3.5 in [4] we have: Theorem 5.2 Under the assumptions (L0)-(L4) above, if µ0 ≥ 2 and there exist a subcritical critical point x and a supercritical critical point ¯x of L such that x ≪ θ ≪ ¯x, then the conclusions of Theorem 5.1 is still true. Similarly, we can get the corresponding results with Theorems 3.6, 3.8 in [4] as follows. Theorem 5.3 Under the assumptions (L0)-(L4) above, if µ∞ ≥ 2 and µ0 + ν0 ≤ 1 then L has a sign-changing critical point x1 with L(x1) > 0. If all sign changing critical points (i.e. those in X \D ) with critical values contained in a bounded interval of [0, ∞) are isolated, then there exists a sign changing critical point x1 with L(x1) > 0, Morse index µ ∈ {1, 2} and C0(L, x1; K) = 0 = C1(L, x1; K), C2(L, x1; K) 6= 0. Furthermore, if near x1 the conditions of Theorem 2.1 of [36] hold then x1 is neither a local minimum nor of mountain pass type, and Ck(L, x1; K) = δk2K holds for all k provided that µ = 2 or the nullity ν ≤ 1. 3This is only place where Proposition 2.26 of [36] is used. The assumptions of Theorem 2.1 of [36] and (L1)-(L4) are sufficient to other arguments. 20 Theorem 5.4 Under the assumptions (L0)-(L4) above, let ν0 = 0 = ν∞, µ∞ ≥ 1 and µ0 6= µ∞. Suppose also that all sign changing critical points are isolated. Then (i) If µ0 ≥ 1 then L has a sign changing critical point x1 which satisfies either L(x1) > 0 and Cµ0+1(L, x1; K) 6= 0 or L(x1) < 0 and Cµ0−1(L, x1; K) 6= 0. (ii) If (L4) (iii) applies with µ∞ ≥ 2 then L has a sign changing critical point x1 with Cµ∞(L, x1; K) 6= 0. The last theorem can be obtained by completely repeating the proof of Theorem 3.8 in [4]. Proof of Theorem 5.3. For reader's convenience we follow the proof ideas of [4, Theorem 3.6] to give necessary details. Since µ∞ ≥ 2 either (L4)-(ii) or (L4)-(iii) occurs. Claim 4. There exist two orthogonal unit vectors v∞ ∈ IntX (P ) and u∞ ∈ X such that L(u) < 0 for u ∈ span{v∞, u∞} with kuk ≥ R. In fact, In the latter case, the two smallest eigenvalue of B∞, λ1 := (B∞v∞, v∞)H < λ2 := inf{(B∞u, u)H kuk = 1, u ∈ H1} are negative, where v∞ = e∞/ke∞k and H1 := H ∩ hv∞i⊥. Since X is dense in H, X ∩ H1 6= ∅. Let u∞ be a unit vector in X ∩ H1. Then (v∞, u∞)H = 0. As in the arguments below Claim 2 in the proof of Theorem 5.1 we have N > 0 and M > 0 such that k∇L(u) − B∞uk < λ2 4 kuk as kuk ≥ N, L(u) ≤ M N ∀u ∈ BH(θ, N ). For any u ∈ span{v∞, u∞} with kuk > N let ¯u = N · u/kuk. Then (∇L(tu + (1 − t)¯u), (u − ¯u))H dt + L(¯u) (B∞(tu + (1 − t)¯u), u − ¯u)H dt + λ2 4 ku − ¯uk · ktu + (1 − t)¯uk + M N L(u) = Z 1 ≤ Z 1 0 0 1 2 λ2 2 ≤ ≤ (B∞(u − ¯u), u − ¯u)H + (B∞(¯u), u − ¯u)H + λ2 4 ku − ¯uk · k¯uk + M N ku − ¯uk2 + (B∞(¯u), u − ¯u)H + λ2 4 ku − ¯uk · k¯uk + M N. So Claim 4 follows from this in this case. In the former case take any unit vector v∞ ∈ IntX(P ). As above we can choose another unit vector u∞ ∈ X which is orthogonal to v∞. Since for any u ∈ H with L(u) ≤ a it holds that DL(u)(u) < 0, La is a manifold with C 1-boundary L−1(a), and L−1(a) is transversal to the radial vector field. Moreover, for any u ∈ H \ {θ}, L(tu) → −∞ as t → ∞. So for each u ∈ ∂BH (θ, 1) there exists a tu ∈ (0, ∞) such that L(tu · u) = a and that u → tu is continuous by the implicit function theorem. This shows that the map ∂BH (θ, 1) → L−1(a), u → tu · u is a homeomorphism. It follows 21 that ∂BH(θ, 1) ∩ span{v∞, u∞} is compact and hence ∂BH (θ, 1) ∩ span{v∞, u∞} ⊂ BH(θ, R) for some R > 0. We have still L(u) < 0 for u ∈ span{v∞, u∞} with kuk ≥ R. That is, Claim 4 holds. For R in Claim 4 let us set BR := {sv∞ + tu∞ : ∂BR := {sv∞ + tu∞ : s ≤ R, 0 ≤ t ≤ R}, s = R, or t ∈ {0, 1}}. Then ∂BR ⊂ L0 ∪ D. Let β := max L(BR) and ξβ ∈ H2(Lβ ∪ D, L0 ∪ D; K) be the image of 1 ∈ K ∼= H2(BR, ∂BR; K) under the homomorphism K ∼= H2(BR, ∂BR; K) → H2(Lβ ∪ D, L0 ∪ D; K) induced by the inclusion (BR, ∂BR) ֒→ BR, ∂BR). For γ ≤ β let (jγ)2 : H2(Lγ ∪ D, L0 ∪ D; K) → H2(Lβ ∪ D, L0 ∪ D; K) be the homomorphism induced by the inclusion. Set Γ := {γ ≤ β ξβ ∈ Image(jγ )2)} and c := inf Γ. (5.3) Let e1 ∈ IntX(P ) be the eigenvector of B(θ) belonging to the first eigenvalue λ1 = inf kuk=1(B(θ)u, u)H , and let X1 = he1i and X2 := X ⊥ 1 ∩ X. Since µ0 + ν0 ≤ 1. There are three cases: (a) µ0 = 0 = ν0, (b) µ0 = 1 and ν0 = 0, (c) µ0 = 0 and ν0 = 1. For the first two case we have X2 ⊂ H +. In the third case, λ1 = 0 and X1 = Ker(B(θ)), and H − = {θ}. We also get X2 ⊂ H +. It follows from [36, (2.74)] that L(u) ≥ kuk2 a1 4 for all u ∈ BH(θ, ρ0) ∩ H +. Hence for some small ρ > 0 it always holds that inf{L(u) u ∈ X2, kuk = ρ} > 0. This is what is needed in the proof of [4, Lem.4.3]. It leads to ξβ 6= 0 and hence β ∈ Γ. Moreover, (j0)2 = 0 implies that 0 /∈ Γ. Since we have assumed that θ is an isolated critical point4, and that all sign changing critical points with critical values in a bounded interval of [0, ∞), by the proof of Claim 3 in the proof of Theorem 5.1 we can derive that there exist only finitely many sign changing critical points with critical values in [0, β]. It follows that L0 ∪ D is a strong deformation retract of Lγ ∪ D for γ > 0 small enough. So c > 0. The remained arguments are the same as those of [4, Th.3.6] (as long as slightly modifications as in the proof of Theorem 5.1. ✷ Remark 5.5 Let us outline a possible way to weaken the conditions of Theorem 5.1, that is, removing the assumption that L is C 2−0, but adding the condition (6.5.1) "For any subset S ⊂ X, which is bounded in H, the image A(S) is bounded in X" in case (L4)-(i); 4In [4] it was claimed that since µ0 + ν0 ≤ 1 the sign changing solutions cannot accumulate at 0. 22 (6.5.2) "For any c ∈ R and small ε > 0, A(X ∩ L−1([c − ε, c + ε])) is bounded in X" in case (L4)-(iii). Since (F2) and Proposition B.1(ii) of [36] imply that the map A : X → X is locally Lipschitz, we get a (local) flow on X \ K(L), d dt σ(t, x) = −A(σ(t, x)) σ(0, x) = x ∈ X \ K(L), ) (5.4) where K(L) is the critical set of L. By (L2) we get that σ(t, x) ∈ IntX(D) for all x ∈ D∗ and t > 0. The key is how to prove Claim 3 in the present assumptions. Note that we have proved c < 0 above Claim 3. Then Claim 1 implies that L−1(c) contains at most finitely many sign changing critical points x1, · · · , xq and that there exist a η > 0 such that no number in [c − η, c + η] \ {c} is a critical value of sign changing critical points. By contradiction, suppose that c is not a critical value of L. Then the PS condition implies that there exist ε > 0 and δ > 0 such that k∇L(u)k = kDL(u)k ≥ δ ∀u ∈ L−1[c − ε, c + ε]. We may assume ε < η. By (L2) and kukX ≥ kuk ∀u ∈ X we get kA(u)kX ≥ kA(u)k = kDL(u)k ≥ δ ∀u ∈ X ∩ L−1[c − ε, c + ε]. (5.5) For x ∈ X ∩ L−1[c − ε, c + ε] let [0, Tx) be the maximal existence interval of the flow in (5.4) on X ∩ L−1[c − ε, c + ε]. Then η + ε ≥ L(x) − L(σ(t, x)) =Z t 0 kA(σ(s, x))k2ds ≥ δ2t for any t ∈ [0, Tx). So Tx ≤ (η + ε)/δ2. In case (L4)-(i), L is coercive by a result of [7]. It follows that Lc+ε is bounded in H and hence A(X ∩ Lc+ε) is bounded in X by the assumption (6.5.1). Namely, there exists a N > 0 such that kA(x)kX ≤ N for all x ∈ X ∩ Lc+ε. Then d dt t1 (cid:13)(cid:13)(cid:13)(cid:13) σ(t, x)(cid:13)(cid:13)(cid:13)(cid:13)X distX(σ(t2, x), σ(t1, x)) ≤Z t2 dt =Z t2 σ(t, x)(cid:19) = c − ε. L(cid:18) lim t→Tx− for any 0 ≤ t1 < t2 < Tx. It follows that the limit limt→Tx−0 σ(t, x) exists in X and kA(σ(t, x))kX dt ≤ N (t2 − t1) t1 As usual we can use σ to construct a deformation retract from Lc+ε where Ld X ∪D∗, X := X ∩ Ld for d ∈ R. This is a contradiction. Hence c is a critical value. X ∪D∗ to Lc−ε In case (L4)-(iii), by the assumption (6.5.2), A(X ∩L−1([c−ε, c+ε])) is bounded in X for some small ε > 0. Then the same method leads to a contradiction yet.✷ Finally, we are going to generalize the following result, which is an abstract sum- mary of the arguments in [34]. 23 Theorem 5.6 Let H be a Hilbert space, and let PH 6= H be a closed cone, i.e. PH = ¯PH is convex and satisfies R+ · PH ⊂ PH , PH ∩ (−PH) = {θ}. Suppose that a C 2-functional L : H → R has critical point θ and satisfies the following properties. (i) L is bounded from below, and satisfies the (PS) condition. 5 (ii) PH is positively invariant under the negative gradient flow ϕt of L, d dt ϕt(u) = −∇L(ϕt(u)) and ϕ0(u) = u ∈ H. (iii) There exists a positive element e ∈ PH \ {θ} such that the cone D := {u ∈ H u ≥ e} ⊂ PH (resp. −D) contains all positive (resp. negative) critical points of Φ. (Note that D ∩(−D) = ∅). Let Dε := {u ∈ H dist(u, D) ≤ ε} for ε > 0. (iv) There exists a ε0 > 0 such that for each ε ∈ (0, ε0], Dε0 ∩ (−Dε0) = ∅6, and Dε and −Dε are strictly positively invariant for the the flow ϕt. Let Wε := Dε ∪ (−Dε) and let ic : (Lc ∪ Wε, Wε) → (H, Wε) be the inclusion. Then c1 := inf{c ∈ R i∗ c : ¯H 1(H, Wε; Z2) → ¯H 1(Lc ∪ Wε, Wε; Z2) is a monomorphism} (5.6) is finite and K ∗ c1 := {u ∈ H \ Wε, L(u) = c1, L′(u) = 0} 6= ∅ for 0 < ε ≤ ε0, where i∗ c is induced by the inclusion ic. Moreover, L has at least two nontrivial critical points in H \ Wε provided that c1 < L(θ) and each u ∈ K ∗ Cq(L, θ; Z2) ∼= δqnZ2 for some n ≥ 2 and any q ∈ Z. c1 with Morse index µ(u) = 0 has nullity ν(u) ≤ 1, (5.7) (5.8) Remark 5.7 (i) Actually, the assumptions (i)-(ii) in Theorem 5.6 imply that L has a critical point u ∈ PH with L(u) = infv∈PH L(v) ([33, Theorem 2.1]). In the same way the assumption (iv) yields a critical point sitting in −D ⊂ PH. (ii) Note that (5.8) holds if θ is a nondegenerate critical point of L with Morse index n ≥ 2. (iii) If K := id − ∇L : H → H maps Dε (resp. −Dε) into int(Dε) (resp. int(−Dε)), then (iv) holds by the proof of Proposition 3.2(ii) of [34]. We shall generalize Theorem 5.6 as follows. Theorem 5.8 Let H be a Hilbert space, and let PH 6= H be a closed cone, i.e. PH = ¯PH is convex and satisfies R+ · PH ⊂ PH , PH ∩ (−PH) = {θ}. Suppose that a C 1-functional L : H → R has critical point θ and satisfies the following properties. 5By a result of [7] these two conditions imply the coercivity of L. 6By considering the functional f (x) = (x, e)H we obtain D ⊂ {f > 0} and −D ⊂ {f < 0}. This implies dist(D, −D) ≥ 2kek and thus dist(Dε0 , −Dε0) ≥ 2kek − 2ε0 for ε < kek! 24 (i) L is bounded from below, and satisfies the (PS) condition. (ii) There exists a positive element e ∈ PH \ {θ} such that the cone D := {u ∈ H u ≥ e} ⊂ PH (resp. −D) contains all positive (resp. negative) critical points of L. (iii) There exists a ε0 ∈ (0, kek) such that for each ε ∈ (0, ε0], K := id − ∇L : H → H maps Dε (resp. −Dε) into int(Dε) (resp. int(−Dε)). ) (5.9) Then there exists ε1 ∈ (0, ε0] such that for any 0 < ε ≤ ε1, c1 defined by (5.6) is finite c1 := {u ∈ H \ Wε, L(u) = c1, L′(u) = 0} 6= ∅. Moreover, if c1 < L(θ) and the and K ∗ conditions of Theorem 2.1 of [36] are satisfied near each u ∈ K ∗ c1 and θ then L has at least two nontrivial critical points in H \ Wε provided that (5.7) and (5.8) hold. Since Dε and −Dε are positive invariant under the pseudo-gradient flow eϕt de- fined by (5.11) (see proof below (5.12)), as in Remark 5.7(i) we may show that L has a critical point u ∈ Dε (resp. u ∈ −Dε) with L(u) = infv∈Dε L(v) (resp. L(u) = infv∈−Dε L(v)). Remark 5.7(iii) shows that Theorem 5.8(iii) is stronger than Theorem 5.6(iv). Proof of Theorem 5.8. The basic ideas is almost the same as in [34]. However, since we only assume L to be C 1, the negative gradient flow of it cannot be used. We shall overcome this difficulty by some methods in [3]. Note that V := ∇L is a C 0 pseudo-gradient vector field for L in the sense of [3], i.e. V : H → H is a continuous map satisfying kV (x)k < 2k∇L(x)k and (∇L(x), V (x))H > k∇L(x)k2 ∀x ∈ H \ K(L). 1 2 Moreover, (5.9) means that D and −D are K-attractive in the sense of [3, Definition 3.3]. By [3, Lemma 3.4] there exists ε1 ∈ (0, ε0] such that for every σ ∈ (0, ε1) there for any u ∈ ∂Dε, ∃ ǫ0 > 0 such that ∀ǫ ∈ (0, ǫ0], is a pseudo-gradient vector field eVσ of L such that for all ε ∈ [σ, ε1] the sets Dε and −Dε are strongly positive invariant under −eVσ in the following sense: u + ǫeVσ(u) ∈ Int(Dε) and −u + ǫeVσ(−u) ∈ Int(−Dε). ) dteϕt(u) = −eVσ(eϕt(u)) eϕ0(u) = u ∈ H. strictly positive invariant for the flow eϕt. ) Let eKσ := id −eVσ and let eϕt be the flow of −eVσ, i.e. for any ε ∈ [σ, ε0] the sets Dε and −Dε are (5.10) (5.11) (5.12) d and We claim: 25 d ds f (eϕs(u))(cid:12)(cid:12)(cid:12)s=t = f(cid:0)−eV (eϕt(u))(cid:1) = = 1 ǫ 1 ǫ 1 ǫ f(cid:0)eϕt(u) − ǫeV (eϕt(u))(cid:1) − f(cid:0)eϕt(u) − ǫeV (eϕt(u))(cid:1) − f(cid:0)−ǫeV (eϕt(u))(cid:1) f(cid:0)eϕt(u)(cid:1) β > 0. 1 ǫ 1 ǫ = Indeed, from (5.10) and Theorem 5.2 in [23] we deduce that the sets Dε and −Dε As in the proof of [34, Proposition 3.2(ii)], suppose by contradiction that there are positive invariant for the flow eϕt for any ε ∈ [σ, ε0]. exist u ∈ Dε and t > 0 such that eϕt(u) ∈ ∂Dε. Then Mazur's separation theorem yields f ∈ H ∗ and β > 0 such that f (eϕt(u)) = β and f (u) > β for any u ∈ Int(Dε). By (5.10) we have ǫ0 > 0 such that eϕt(u) + ǫeV (eϕt(u)) ∈ Int(Dε) ∀ǫ ∈ (0, ǫ0]. Hence This leads to a contradiction as in [34]. The assertion in (5.11) is proved. Having (5.12) we may follow the lines of [34] to outline the remaining proof. Let ¯H ∗ denote Alexander-Spanier cohomology with coefficients in the field Z2. For a critical point u of L the cohomological critical groups of L at u are defined by C q(L, u) := ¯H q(Lc, Lc \ {u}) ∀q ≥ 0, where c = L(u). We conclude C q(L, u) ∼= Cq(L, u; Z2) ∀q ∈ Z. (5.13) This can be obtained from the proof of Proposition 3.3. We can also prove it as follows. By excision property it holds that C q(L, u) = ¯H q(Lc ∩ U, (Lc \ {u}) ∩ U ) ∀q ≥ 0 for any open neighborhood U of u. In particular, one can find a U so that both Lc ∩ U and (Lc \ {u}) ∩ U are absolute neighborhood retracts ( [21, Th.1.1] and [17, Remark 2]). It follows that ¯H q(Lc ∩ U, (Lc \ {u}) ∩ U ) ∼= H q(Lc ∩ U, (Lc \ {u}) ∩ U ) ∀q ≥ 0. Moreover, H q(Lc ∩ U, (Lc \ {u}) ∩ U ; G) ∼= Hom(Hq(Lc ∩ U, (Lc \ {u}) ∩ U ; G), G) for any divisible group G. If C 1(L, u) 6= 0 we say u to be of mountain-pass type. Lemma 5.9 ([34, Lemma 4.4]). If C 1(L, u) 6= 0, and ν(u) ≤ 1 in case µ(u) = 0, then C q(L, u) = δq1Z2 for q ∈ Z. Proof. Note that ¯H q(Bn, Sn−1; G) = δqnG. If ν(u) = Ker(L′′(u)) = 0, i.e. u is nondegenerate, by Morse lemma we get that C q(L, u) ∼= H q(Bµ(u), Sµ(u)−1; Z2) = δqµ(u)Z2. The desired conclusion follows. If ν(u) = Ker(L′′(u)) > 0, from Corollary 2.6 of [36] it follows that µ(u) ≤ 1 ≤ µ(u) + ν(u) and C 1(L, u) ∼= C 1−µ(u)(L◦, θ), where L◦ is a function defined near the origin of a ν(u)-dimensional space Ker(L′′(u)). If µ(u) = 1 then Corollary 2.9(iii) of [36] gives the conclusion. If µ(u) = 0 then ν(u) = 1 by the fact that 1 ≤ µ(u)+ν(u) = ν(u). Corollary 2.9(ii) of [36] yields the conclusion. ✷ The following is a special case of the strong excision property (Theorem 5) on the page 318 of [48]. 26 Lemma 5.10 ([34, Lemma 4.4]). Let X be paracompact Hausdorff space and let Y, Z ⊂ X be closed subsets such that X = Y ∪ Z. Then the inclusion (Y, Y ∩ Z) → (X, Z) induces an isomorphism ¯H ∗(X, Z) → ¯H ∗(Y, Y ∩ Z). By the assumption (i) in Theorem 5.8, L is bounded from below, for each u ∈ H the flow t 7→ eϕt(u) exists in an open interval containing [0, ∞). For −∞ ≤ c ≤ ∞, Lc is positively invariant for the flow eϕt, and strictly positively invariant if c is a regular value of L. Lemma 5.11 ([34, Lemma 4.3]). Let u be a critical point of L with L(u) = c and such that BH(u, 2ε) contains no other critical point of L. Then for δ > 0 sufficiently small there exists a closed neighborhood N ⊂ BH(u, ε)∩Lc+δ of of u such that N ∪Lc−δ is positively invariant for the flow eϕt. Moreover, C q(L, u) ∼= ¯H q(N ∪ Lc−δ, Lc−δ) ∼= ¯H q(N, Lc−δ ∩ N ) ∀q ∈ Z. Fix ε ∈ [σ, ε1], then Wε := Dε ∪ (−Dε) is closed and strictly positively invariant under eϕt by (5.11). In particular, ∂Wε contains no critical points of L. So the (PS) condition implies that K ∗ of H \ Wε for every c ∈ R. Clearly, every non-trivial critical point in K ∗ c := {u ∈ H \ Wε, L(u) = c, L′(u) = 0} is a compact subset c changes sign. Lemma 5.12 ([34, Lemma 4.5]). Suppose that −∞ ≤ a < b ≤ c ≤ d ≤ ∞ satisfy: (i) K ∗ c′ = ∅ for any c′ ∈ [b, d] \ {c} (ii) There is a neighborhood N ⊂ Ld of K ∗ c such that Lb ∪ N is is positively invariant under eϕt. Then the inclusion (N ∪ Lb ∪ Wε, La ∪ Wε) → (Ld ∪ Wε, La ∪ Wε) is a homotopy equivalence. inclusion (Lb ∪ Wε, La ∪ Wε) → (Ld ∪ Wε, La ∪ Wε) is a homotopy equivalence. c = ∅ for any c ∈ [b, d], then the In particular, if K ∗ From Lemmas 5.10, 5.11 and 5.12 we have Lemma 5.13 ([34, Lemma 4.7]). If the compact set K ∗ points u1, · · · , um for some c ∈ R, then c consists of isolated critical ¯H q(Lc+δ ∪ Wε, Lc−δ ∪ Wε) ∼= C q(L, ui) mMi=1 for q ∈ Z and δ > 0 sufficiently small. Note that ¯H q(H, Wε; Z2) ∼= Hom(Hq(H, Wε; Z2); Z2) ∼= δq1Z2. (5.14) By the definition of c1 and the assumption (i) we get c1 > −∞. 27 Lemma 5.14 ([34, Lemma 4.8]). (i) K ∗ (ii) If c1 < L(θ) then K ∗ (iii) If K ∗ c1 6= ∅, and c1 ≤ L(θ) provided that L(te) ≤ L(θ) ∀t ∈ [−1, 1]. c1 consists of non-trivial critical points. c1 consists of only one isolated critical point u, then C q(L, u) ∼= ¯H q(Lc1+δ ∪ Wε, Lc1−δ ∪ Wε) ∼= δq1Z2 for small δ > 0. The claim that c1 ≤ L(θ) in (i) can be proved by (5.14). Lemma 5.12 leads c1 6= ∅. Then the assumption and Lemma 5.13 imply C q(L, u) ∼= ¯H q(Lc+δ ∪ to K ∗ Wε, Lc−δ ∪ Wε) for small δ > 0. From this and the definition of c1 it follows that C 1(L, u) 6= 0. Combing it with (5.7) together, we may use Lemma 5.9 to infer (iii). Now (5.8), (5.14) and Lemmas 5.11, 5.12 and 5.14 lead to Lemma 5.15 ([34, proposition 4.9]). Under the assumption (5.8), if c1 < L(θ) and then there exists c 6= c1 such that K ∗ c contains a non-trivial critical point of L. Summarizing these two lemmas we complete the proof for ε ∈ [σ, ε1]. Since σ ∈ (0, ε0) is arbitrary Theorem 5.8 is proved. ✷ Part II Applications In progress! References [1] P. Bartolo, V. Benci, and D. Fortunato, Abstract critical point theorems and applications to some nonlinear problems with strong resonance at infinity, Non- linear Anal., 7(1983), 981-1012. [2] T. Bartsch, Critical Point Theory on Partially Ordered Hilbert Spaces, J. Funct. Anal., 186(2001), 117C152. [3] Bartsch, Thomas Linking, positive invariance and localization of critical points. Morse theory, minimax theory and their applications to nonlinear differential equations, 9 -- 21, New Stud. Adv. Math., 1, Int. Press, Somerville, MA, 2003. [4] T. Bartsch, K.-C. Chang, Z.-Q. Wang, On the Morse indices of sign changing solutions of nonlinear elliptic problems, Math. Z. 233(2000), 655C677. [5] T. Bartsch, S.-J. Li, Critical point theory for asymptotically quadratic function- als and applications to problems with resonance, Nonlinear Analysis, Theory, Methods and Applications, 28(1997), no. 3, 419 - 441. 28 [6] H. Brezis, Analyse Fonctionnelle -- Th´eorie et applications, Masson, Paris (1983). [7] L. Caklovic, S. Li and M. Willem, A note on Palais-Smale condition and coer- civity, Differential Integral Equations, 3(1990), 799-800. [8] G. Cerami, Un criterio de esistenza per i punti critici su variet'a ilimitate, Rc. Ist. Lomb. Sci. Lett., 112(1978), 332-336. [9] K. C. Chang, Infinite Dimensional Morse Theory and Multiple Solution Problem. Birkhauser, 1993. [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] K. C. Chang, Solutions of asymptotically linear operator equations via Morse theory, Comm. Pure. Appl. Math., 34(1981), 693-712. [12] K. C. Chang and N. Ghoussoub, The Conley index and the critical groups via extension of Gromoll-Meyer theory, Topol. Methods Nonlinear Anal., 7(1996), 77-93. [13] S. Cingolani and M. Degiovanni, On the Poincar-Hopf theorem for functionals defined on Banach spaces, Adv. Nonlinear Stud. 9(2009), no. 4, 679 -- 699. [14] F. H. Clarke, Optimization and Nonsmooth Analysis, Wiley, New York, 1983. [15] J. B. Conway, A Course in Functional Analysis, Springer, New York, 1990. [16] J. N. Corvellec, Morse theory for continuous functionals, J. Math. Anal. Appl. 196(1995), 1050-1072. [17] J. N. Corvellec, On the second deformation lemma, Topological Methods in Non- linear Analysis, 17(2001), 55-66. [18] J. N. Corvellec and M. Degiovanni, Deformation properties for continuous func- tionals and critical point theory, Topological methods in nonlinear analysis. 1(1993), 151-171. [19] J. N. Corvellec and A. Hantoute, Homotopical stablity of isolated critical points of continuous functionals, Set-Valued Anal., 10(2002), 143-164. Calculus of vari- ations, nonsmooth analysis and related topics. [20] E. De Giorgi, A. Marino, M. Tosques, Problemi di evoluzione in spazi metrici e curve di massima pendenza, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur., 68(1980), 180-187. [21] M. Degiovanni, Critical groups of finite type for functionals defined on Banach spaces, Commun. Appl. Anal., 13(2009), no.3, 395-410. [22] M. Degiovanni, M. Marzocchi, A critical point theory for nonsmooth functionals, Ann. Mat. Pura Appl., 167(1994), 73-100. [23] K. Deimling, Ordinary Differential Equations in Banach Spaces, Lecture Notes in Math., Vol. 596, Springer-Verlag, New York, 1977. 29 [24] N. Hirano, S.-J. Li, Z.Q.Wang, Morse theory without (PS) condition at isolated values and strong resonance problem, Calc. Var. Partial Differential Equations, 10(2000), 223 - 247. [25] Sze-Tsen Hu, Homotopy Theory, Academic Press, New York and London 1959. [26] A. Ioffe, E. Schwartzman, Parametric Morse lemmas for C 1,1-functions. (English summary) Recent developments in optimization theory and nonlinear analysis (Jerusalem, 1995), 139C147, Contemp. Math., 204, Amer. Math. Soc., Provi- dence, RI, 1997. [27] G. Katriel, Mountain pass theorems and global homeomorphism theorems, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire, 11(1994), 189-209. [28] N.C. Kourogenis and N. S. Papageorgiou, Nonsmooth critical point theory and nonlinear elliptic equations at resonance, Kodai Math. J., 23(2000), no. 1, 108 -- 135. [29] A. Krist´aly, Nonsmooth critical point theories with applications in elliptic prob- lems and theory of geodesics, University Debrecen Institute of Mathematics Debrecen, 2003. [30] S.-J. Li, A new Morse theory and strong resonance problems, Topological Meth- ods in Nonlinear Analysis, 21(2003), 81-100. [31] S.-J. Li, Some advances in Morse theory and minimax theory. Morse theory, minimax theory and their applications to nonlinear differential equations, 91 -- 115, New Stud. Adv. Math., 1, Int. Press, Somerville, MA, 2003. [32] J. Q. Liu, The Morse index of a saddle point, Syst. Sc. and Math. Sc., 2(1989), 32-39. [33] Z. L. Liu and J. X. Sun. Invariant sets of descending flow in critical point theory with applications to nonlinear differential equations, J. Diff. Eqns., 172(2001), 257-299. [34] Z. L. Liu, Z.-Q. Wang and T. Weth, Multiple solutions of nonlinear Schrodinger equations via flow invariance and Morse theory, Proc. Roy. Soc. Edinburgh Sect, A 136(2006), 945-969. [35] G. Lu, The Conley conjecture for Hamiltonian systems on the cotangent bundle and its analogue for Lagrangian systems, J. Funct. Anal. 259(2009), 2967 -- 3034. [36] G. Lu, The splitting lemmas for nonsmooth functionals on Hilbert spaces, Preprint 2011. [37] G. Lu, Corrigendum: The Conley conjecture for Hamiltonian systems on the cotangent bundle and its analogue for Lagrangian systems. arXiv:0909.0609 v2. [38] Jean Mawhin and Michel Willem, Critical Point Theory and Hamiltonian Sys- tems, Applied Mathematical Sciences Vol.74., Springer-Verlag, 1989. [39] John Milnor, Morse theory, Princeton University Press, Princeton, N.J., 1963. [40] R. S. Palais, Morse theory on Hilbert manifolds, Topology, 2(1963), 299-340. 30 [41] K. Perera, R. P. Agarwal, Donal O'Regan, Morse Theoretic Aspects of p- Laplacian Type Operators, Mathematical Surveys and Monographs Vol.161, American Mathematical Society, Providence Rhode Island 2010. [42] K. Perera and M. Schechter, Double Resonance Problems with Respect to The Fuc´ık Spectrum, Indiana Univ. Math. J., 52(2003), 156-164. [43] P. H. Rabinowitz, Variational methods for nonlinear eigenvalue problems, Eigen- values of Nonlinear Problems, Ed. Cremonese, Roma (1974), 141-195. [44] E. H. Rothe, Morse theory in Hilbert space, Rocky Mountain J. Math., 3(1977), 251-274, Rocky Mountain Consortium Symposium on Nonlinear Eigenvalue Problems (Santa Fe, N.M., 1971). [45] Winfried Schirotzek, Nonsmooth Analysis, Springer 2007. [46] I. V. Skrypnik, Nonlinear Elliptic Equations of a Higher Order [in Russian], Naukova Dumka, Kiev (1973). [47] I. V. Skrypnik, Nonlinear Elliptic Boundary Value Problems, Teubner, Leipzig, 1986. [48] E. H. Spanier, Algebraic Topology, Springer-Verlag, 1981. [49] S. A. Vakhrameev, Critical point theory for smooth functions on Hilbert mani- folds with singularities and its application to some optimal control problems, J. Sov. Math., 67(1993), No. 1, 2713-2811. [50] Z. Q. Wang, A note on the second variation theorem, Acta. Math. Sinica, 30(1987), 106-110. 31
1208.6590
1
1208
2012-08-31T19:44:48
On a connection between a generalised modulus of smoothness of order~$r$ and the best approximation by algebraic polynomials
[ "math.FA" ]
In this paper an asymmetrical operator of generalised translation is introduced, the generalised modulus of smoothness is defined by its means and the direct and inverse theorems in approximation theory are proved for that modulus. ----- V danno\v{i} rabote vvoditsya nesimmetrichny\v{i} operator obobshchennogo sdviga, s ego pomoshchyu opredelyaetsya obobshchenny\v{i} modul' gladkosti i dlya nego dokazyvaetsya pryamaya i obratnaya teoremy teorii priblizheni\v{i}.
math.FA
math
Фундаментальная и прикладная математика 1999, том 5, No. 2, 563 -- 587 О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ И НАИЛУЧШИМИ ПРИБЛИЖЕНИЯМИ АЛГЕБРАИЧЕСКИМИ МНОГОЧЛЕНАМИ M. K. Потапов AND Ф. M. Бериша Аннотация. В данной работе вводится несимметричный оператор обоб- щенного сдвига, с его помощью определяется обобщенный модуль гладко- сти и для него доказывается прямая и обратная теоремы теории прибли- жений. Abstract. In this paper an asymmetrical operator of generalised translation is introduced, the generalised modulus of smoothness is defined by its means and the direct and inverse theorems in approximation theory are proved for that modulus. 1. Введение Для 2π -- периодических функций хорошо известны связи между r-ым обыч- ным модулем гладкости ωr(f, δ)p∗ функции f ∈ Lp∗ с ее наилучшими приближе- ниями En(f )p∗ тригонометрическими полиномами порядка не выше чем, n− 1: C1En(f )p∗ ≤ ωr(cid:18)f, 1 n(cid:19)p∗ ≤ C2 1 nr n νr−1Eν (f )p∗, (1.1) Xν=1 где C1 и C2 (cid:22) положительные постоянные, не зависящие от f и n (n ∈ N). При рассмотрении непериодических функций, заданных на конечном отрез- ке вещественной оси, уже не удается получить такие же связи между обыч- ными модулями гладкости этих функций и их наилучшими приближениями алгебраическими многочленами. Однако полная аналогия с 2π периодическим случаем имеет место тогда, когда обычный модуль гладкости заменен обобщенным модулем гладкости (см. например [3, 2, 5, 7]). В этой работе доказывается аналог неравенства (1.1) для r-го обобщенного модуля гладкости, определяемого при помощи одного несимметричного опера- тора обобщенного сдвига. 2. Определение обобщенного модуля гладкости Обозначим через Lp, 1 ≤ p < ∞, множество функций f , измеримых по Лебегу и суммируемых в p-й степень на отрезке [−1, 1], а через L∞ обозначим 1991 Mathematics Subject Classification. Primary 41A35, Secondary 41A50, 42A16. (UDK 517.5.) Key words and phrases. Generalised modulus of smoothness, asymmetric operator of gener- alised translation, Jackson theorem, converse theorem, best approximations by algebraic polyno- mials. Работа выполнена при поддержке Российского Фонда Фундаментальных Исследования (грант No. 97 -- 01 -- 00010) и программы поддержки ведущих научных школ (грант No. 96/97 -- 15 -- 96073). 1 где Bcos t(x, cos ϕ, R)f (R) dϕ, 1 2 Z π 0 τt (f, x) = π(1 − x2) cos4 t R = x cos t −p1 − x2 sin t cos ϕ, By(x, z, R) = 2(cid:16)p1 − x2y + xzp1 − y2 +p1 − x2(1 − y)(1 − z2)(cid:17)2 2 M. K. Потапов и Ф. M. Бериша множество функций, непрерывных на отрезке [−1, 1], причем если 1 ≤ p < ∞, если p = ∞. −1 f (x)p dx(cid:17)1/p (cid:16)R 1 max−1≤x≤1 f (x), , kfkp =   Через Lp,α обозначим множество функций f , таких, что f (x)(1 − x2)α ∈ Lp, причем Через En(f )p,α обозначим наилучшее приближение функций f при помощи алгебраических многочленов степени не выше, чем n − 1, в метрике Lp,α, т.е. kfkp,α = (cid:13)(cid:13)f (x)(1 − x2)α(cid:13)(cid:13)p . En(f )p,α = inf Pn kf − Pnkp,α , где Pn (cid:22) алгебраический многочлен степени не выше, чем n − 1. вилу Для суммируемой функции f введем оператор обобщенного сдвига по пра- (2.1) − (1 − R2). При помощи этого оператора обобщенного сдвига определим r-ю обобщен- ную разность по правилу ∆1 t (f, x) = ∆t (f, x) = τt (f, x) − f (x), ∆r t1,...,tr (f, x) = ∆tr (cid:16)∆r−1 t1,...,tr−1 (f, x) , x(cid:17) (r = 2, 3, . . . ). и, для функции f ∈ Lp,α, r-й обобщенный модуль гладкости по правилу ωr(f, δ)p,α = sup tj ≤δ j=1,2,...,r(cid:13)(cid:13)∆r t1,...,tr (f, x)(cid:13)(cid:13)p,α (r = 1, 2, . . . ). Полагая y = cos t, z = cos ϕ в операторе τt (f, x), обозначим его через τy (f, x) и запишем в виде τy (f, x) = 4 π(1 − x2)(1 + y)2 Z 1 −1 By(x, z, R)f (R) где R и By(x, z, R) определены формулами (2.1). dz √1 − z2 , Определим r-й оператор обобщенного сдвига по правилу τ 1 y (f, x) = τy (f, x) , τ r y1,...,yr (f, x) = τyr (cid:16)τ r−1 t1,...,tr−1 (f, x) , x(cid:17) (r = 2, 3, . . . ). Обозначим через Dx,ν,µ оператор дифференцирования, определяемый по правилу Ясно, что Dx,ν,µ = (1 − x2) d2 dx2 + (µ − ν − (ν + µ + 2)x) d dx . Dx,ν,µ = (1 − x)−ν (1 + x)−µ d dx (1 − x)ν+1(1 + x)µ+1 d dx . О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 3 Будем обозначать D1 Dr x,ν,µf (x) = Dx,ν,µf (x), x,ν,µf (x) = Dx,ν,µ(Dr−1 x,ν,µf (x)) (r = 2, 3, . . . ). Будем писать, что f (x) ∈ ADr(p, α), если f ∈ Lp,α, f (x) имеет на каждом dx2r−1 f (x) отрезке [a, b] ⊂ (−1, 1) абсолютно непрерывную 2r − 1 производную d2r−1 и Dl x,2,2f (x) ∈ Lp,α (l = 0, 1, . . . , r). Обозначим через Kr(f, δ)p,α = inf g∈ADr (p,α)(cid:16)kf − gkp,α + δ2r(cid:13)(cid:13)Dr x,2,2g(x)(cid:13)(cid:13)p,α(cid:17) K -- функционал типа Петре, интерполирующий между пространствами Lp,α и ADr(p, α). Для f ∈ L1,2 обозначим через H (f, x) и Hδ (f, x) следующие операторы 0 (1 − y2)−3Z 1 H (f, x) = −Z x −1 f (z)(1 − z2)2 dz, c0 = R 1 2(cid:17)−1(cid:16)cos κ(δ) Z δ 0 (cid:16)sin y (cid:18)f (z) − −1(1 − z2)2 dz; и 2(cid:17)−9 1 v v где c1 = R 1 Hδ (f, x) = c1 c0(cid:19) (1 − z2)2 dz dy, ×Z v 0 τu (f, x)(cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv, где κ(δ) = Z δ 0 (cid:16)sin v 2(cid:17)−1(cid:16)cos v 2(cid:17)−9Z v 0 (cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv. Определим r-ую степень оператора H по правилу H 1 (f, x) = H (f, x) , H r (f, x) = H (cid:0)H r−1 (f, x) , x(cid:1) = −Z x y (cid:18)H r−1 (f, z) − ×Z 1 0 (1 − y2)−3 cr c0(cid:19) (1 − z2)2 dz dy (r = 2, 3, . . . ), где cr = R 1 −1 H r−1 (f, z) (1 − z2)2 dz; и r-ую степень оператора Hδ по правилу δ (f, x) = Hδ (f, x) , H 1 H r δ (f, x) = Hδ (cid:0)H r−1 δ (f, x) , x(cid:1) (r = 2, 3, . . . ). n Будем обозначать через P (ν,µ) (x) (n = 0, 1, . . . ) многочлены Якоби, т.е. мно- гочлены степени n ортогональные друг другу с весом (1−x)ν(1+x)µ на отрезке [−1, 1] и нормированные условием P (ν,µ) Через an(f ) обозначим коэффициенты Фурье -- Якоби функции f ∈ L1,2 по (x)o∞ системе многочленов Якоби nP (2,2) (x)(1 − x2)2 dx (n = 0, 1, . . . ). an(f ) = Z 1 (1) = 1 (n = 0, 1, . . . ). f (x)P (2,2) , т.е. n=0 n n n −1 4 M. K. Потапов и Ф. M. Бериша 3. Вспомогательные утверждения Лемма 3.1 ([4]). Пусть Pn(x) -- алгебраический многочлен степени не выше, чем n − 1, 1 ≤ p ≤ ∞, ρ ≥ 0; α > − α ≥ 0 Тогда справедливы неравенства 1 p при 1 ≤ p < ∞, при p = ∞. kP ′ n(x)kp,α+ 1 2 ≤ C1nkPnkp,α , kPnkp,α ≤ C2n2ρ kPnkp,α+ρ , где постоянные C1 и C2 не зависят от n (n ∈ N). Следствие. Пусть Pn(x) (cid:22) алгебраический многочлен степени не выше, чем n − 1, 1 ≤ p ≤ ∞; 1 p α > − α ≥ 0 при 1 ≤ p < ∞, при p = ∞. Тогда kDx,2,2Pn(x)kp,α ≤ Cn2 kPn(x)kp,α , где постоянная C не зависит от n (n ∈ N). Доказательство. Так как, kDx,2,2Pn(x)kp,α ≤ kP ′′ n (x)kp,α+1 + 6 kP ′ n(x)kp,α , то, применяя дважды лемму 3.1, получаем утверждение следствия. (cid:3) Лемма 3.2 ([6]). Оператор τy обладает следующими свойствами 1) Оператор τy (f, x) линеен по f ; 2) τ1 (f, x) = f (x); n n 4) τy (1, x) = 1; , x(cid:17) = P (2,2) 3) τy (cid:16)P (2,2) 5) an(cid:0)τy (f, x)(cid:1) = an(f )P (0,4) n (x)P (0,4) n (y) (n = 0, 1, . . . ); (y) (n = 0, 1, . . . ). Лемма 3.3 ([6]). Пусть g(x)τy (f, x) ∈ L1,2 для любого y. Тогда справедливо равенство Z 1 f (x)τy (g, x) (1 − x2)2 dx = Z 1 g(x)τy (f, x) (1 − x2)2 dx. Лемма 3.4 ([6]). Пусть даны числа p и α такие, что 1 ≤ p ≤ ∞; −1 −1 1 2 1 2p 1 − < α ≤ 1 3 < α < 2 − 3 2 1 ≤ α < при p = 1, 1 2p при 1 < p < ∞, при p = ∞. Пусть f ∈ Lp,α. Тогда справедливо неравенство kτt (f, x)kp,α ≤ C где постоянная C не зависит от f и t. 1 cos4 t 2 kfkp,α , О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 5 Лемма 3.5 ([6]). Пусть функция f (x) имеет абсолютно непрерывную на каждом отрезке [a, b] ⊂ (−1, 1) производную f ′(x). Тогда для почти всех x ∈ [−1, 1] выполнены следующие равенства τt (f, x) − f (x) = Z t 0 (cid:16)sin v 2(cid:17)−1(cid:16)cos и v 2(cid:17)−9Z v 0 τu (Dx,2,2f, x)(cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv τt (f, x) − τπ/2 (f, x) v π/2(cid:16)sin = −Z t 2(cid:17)−1(cid:16)cos v 2(cid:17)−9Z π v τu (Dx,2,2f, x)(cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv. Лемма 3.6. Пусть функция f (x) имеет на каждом отрезке [a, b] ⊂ (−1, 1) абсолютно непрерывную 2r − 1 производную d2r−1 dx2r−1 f (x). Тогда 1) При фиксированном y функция τy (f, x) имеет на каждом отрезке [c, d] ⊂ dx2r−1 τy (f, x). 2) При фиксированном x функция τy (f, x) имеет на каждом отрезке [c, d] ⊂ dy2r−1 τy (f, x). (−1, 1) абсолютно непрерывную 2r − 1 производную по x d2r−1 (−1, 1) абсолютно непрерывную 2r − 1 производную по y d2r−1 3) Для почти всех x и y справедливы равенства τy (Dx,2,2f, x) = Dx,2,2τy (f, x) = Dy,0,4τy (f, x) . Доказательство. Докажем утверждение 1). Для r = 1 оно доказано в рабо- те [6]. Обозначим ϕ(x) = By(x, z, R) (1 − x2)(1 + y)2√1 − z2 f (R), где R и By(x, z, R) даны формулами (2.1). Применяя индукцию, можно дока- зать, что для l = 1, . . . , 2r − 1 имеем dl dxl ϕ(x) = ϕ(l)(x) = 1 (1 + y)2√1 − z2 l Xk=0 k(cid:19)(cid:18) dl−k (cid:18)r dxl−k By(x, z, R) 1 − x2 (cid:19) dk dxk f (R), где dk dxk f (R) = k Xν=1 dνf (R) dRν Xµ1≥···≥µν ≥0 µ1+···+µν =k αk k Yj=1 dµj R dxµj . Аналогичным рассуждением как в случае r = 1 (см. [6]) доказывается, что функция ϕ(l)(x) абсолютно непрерывна на каждом отрезке [c, d] ⊂ (−1, 1) (l = 1, . . . , 2r − 1). Воспользовавшись теоремой Лебега, при фиксированных y и z, получаем, что существует конечная производная d2r−1 dx2r−1 τy (f, x) (cid:22) абсолютно непрерывная на каждом отрезке [c, d] ⊂ (−1, 1). зывается абсолютная непрерывность функции d2r−1 ном x. Используя симметричность R по x и y, аналогичным рассуждением дока- dy2r−1 τy (f, x) при фиксирован- Утверждение 3) доказано в работе [6]. Лемма 3.6 доказана. (cid:3) 6 M. K. Потапов и Ф. M. Бериша Лемма 3.7. Пусть функция f (x) имеет на каждом отрезке [a, b] ⊂ (−1, 1) абсолютно непрерывную 2l−1 производную d2l−1 dx2l−1 f (x). Тогда для почти всех x и y справедливы равенства x,2,2τ r y1,...,yr (f, x) (r = 1, 2, . . . ). τ r y1,...,yr (cid:0)Dl x,2,2f, x(cid:1) = Dl Пусть l ≥ 2, r = 1. Ясно, что Dl−1 каждом отрезке [a, b] ⊂ (−1, 1) производную d мы 3.6 следует, что Доказательство. При r = l = 1 равенство леммы следует из леммы 3.6. x,2,2f (x) имеет абсолютно непрерывную на x,2,2f (x). Поэтому, из лем- dx Dl−1 Применяя это равенство l раз получим, что τy1 (cid:0)Dl x,2,2f, x(cid:1) = Dx,2,2τy1 (cid:0)Dl−1 x,2,2f, x(cid:1) . τy1 (cid:0)Dl x,2,2f, x(cid:1) = Dl x,2,2τy1 (f, x) . Значит, равенство леммы справедливо при любых l ∈ N и r = 1. любых натуральных r и l. Теперь, применяя индукцию, нетрудно доказать утверждение леммы при Лемма 3.7 доказана. (cid:3) Лемма 3.8. Пусть даны числа p, α и r такие, что 1 ≤ p ≤ ∞, r ∈ N; 1 2 1 2p 1 − < α ≤ 1 3 2 − < α < 3 2 1 ≤ α < при p = 1, 1 2p при 1 < p < ∞, при p = ∞. Пусть g(x) ∈ ADr(p, α). Для 0 ≤ δ < π справедливо неравенство ωr (g, δ)p,α ≤ C 1 (cos δ/2)4r δ2r (cid:13)(cid:13)Dr x,2,2g(x)(cid:13)(cid:13)p,α , где постоянная C не зависит от g и δ. Доказательство. Докажем сначала, что при ti < π (i = 1, . . . , r) справедливо неравенство (cid:13)(cid:13)∆r t1,...,tr (g, x)(cid:13)(cid:13)p,α ≤ C1 Qr i=1 cos4 ti 2 t2 1 . . . t2 r (cid:13)(cid:13)Dr x,2,2g(x)(cid:13)(cid:13)p,α , (3.1) где постоянная C1 не зависит от g и ti (i = 1, . . . , r). Для r = 1 неравенство (3.1) доказано в работе [6]. Предположим, что справедливо неравенство (3.1). Так как и при доказа- t1,...,tr (g, x) вместо g(x), учитывая, что из тельстве для r = 1, только взяв ∆r лемм 3.6 и 3.4 вытекает ∆r t1,...,tr (g, x) ∈ ADr+1(p, α), получим Применяя лемму 3.7 и предположение леммы, получаем, что (cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) ∆r+1 t1,...,tr+1 (g, x)(cid:13)(cid:13)(cid:13)p,α ≤ t1,...,tr+1 (g, x)(cid:13)(cid:13)(cid:13)p,α ≤ ∆r+1 C2 cos4 tr+1 2 C3 cos4 tr+1 2 t2 r+1(cid:13)(cid:13)Dx,2,2∆r r+1(cid:13)(cid:13)∆r t2 t1,...,tr (g, x)(cid:13)(cid:13)p,α . . t1,...,tr (Dx,2,2g, x)(cid:13)(cid:13)p,α что неравенство (3.1) справедливо. На основании индукции, учитывая, что Dx,2,2g(x) ∈ ADr(p, α), получаем, Переходя в (3.1) к точной верхней грани по всем ti, ti < δ (i = 1, . . . , r), Лемма 3.8 доказана. получим неравенство леммы. (cid:3) О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 7 Лемма 3.9. Пусть даны числа p и α такие, что 1 ≤ p ≤ ∞; −1 < α ≤ 2 1 < α < 3 − − p 0 ≤ α < 3 1 p при p = 1, при 1 < p < ∞, при p = ∞. Тогда если f ∈ Lp,α, то H (f, x) ∈ Lp,α. Доказательство. Нетрудно заметить, что при условиях леммы f ∈ L1,2. Зна- чит существует H (f, x). Для 1 ≤ p < ∞ обозначим I = kH (f, x)kp p,α = Z 1 −1 (1 − x2)pαH (f, x) p dx. Пусть p = 1. Рассмотрим I1 = Z 1 0 (1 − x2)αH (f, x) dx ≤ Z 1 0 (1 − x2)αZ x 0 Ясно, что I1 ≤ C1Z 1 0 (1 − x)αZ x 0 (1 − y)−3Z 1 0 0 C1 (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) f (z) − α + 1 Z 1 Z z c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) ≤ C2Z 1 Поскольку f ∈ L1,α и α > −1, то = 0 0 y y c1 f (z) − f (z) − c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z2)2(cid:12)(cid:12)(cid:12)(cid:12) (1 − y2)−3Z 1 (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) (1 − y)−3Z 1 (1 − x)α dx dy dz Z z c0(cid:12)(cid:12)(cid:12)(cid:12) f (z) − (1 − z)αz(cid:12)(cid:12)(cid:12)(cid:12) f (z) − (1 − y)α−2 dy dz dz ≤ C2(cid:13)(cid:13)(cid:13)(cid:13) c0(cid:12)(cid:12)(cid:12)(cid:12) c1 c1 y 0 I1 < ∞. dz dy dx. c1 c0(cid:12)(cid:12)(cid:12)(cid:12) dz dy dx. f − . c1 c0(cid:13)(cid:13)(cid:13)(cid:13)1,α Поменяв пределы интегрирования, учитывая, что −1 < α ≤ 2, имеем I1 ≤ C1Z 1 c1 Рассмотрим I2 = Z 0 −1 (1 − x2)αH (f, x) dx = Z 0 Z x −1 0 (1 − x2)α(cid:12)(cid:12)(cid:12)(cid:12) Z 1 −1 (1 − y2)−3Z 1 y (1 − z2)2(cid:18)f (z) − dx. c1 c0(cid:19) dz dy(cid:12)(cid:12)(cid:12)(cid:12) (1 − z2)2(cid:18)f (z) − c1 c0(cid:19) dz = 0. Из определения c1 и c0 следует, что Поэтому Z 1 y (1 − z2)2(cid:18)f (z) − Отсюда вытекает, что c1 c0(cid:19) dz = −Z y −1 I2 ≤ C3Z 0 −1 (1 + x)αZ 0 x (1 + y)−3Z y −1 c1 (1 − z2)2(cid:18)f (z) − c0(cid:19) dz. c0(cid:12)(cid:12)(cid:12)(cid:12) (1 + z)2(cid:12)(cid:12)(cid:12)(cid:12) f (z) − c1 dz dy dx. (3.2) 8 M. K. Потапов и Ф. M. Бериша Меняя пределы интегрирования получаем I2 ≤ C3Z 0 −1 (1 + z)2(cid:12)(cid:12)(cid:12)(cid:12) f (z) − c1 c0(cid:12)(cid:12)(cid:12)(cid:12) Z 0 z (1 + y)−3Z y −1 (1 + x)α dx dy dz. Отсюда, аналогичным рассуждением как в предыдущем случае получаем, что I2 < ∞. Таким образом, при p = 1 доказано, что I = I1 + I2 < ∞. Значит, H (f, x) ∈ L1,α. Пусть 1 < p < ∞. Имеем H (f, x) ≤ Z x 0 (1 − y2)−3Z 1 y Рассмотрим (1 − z2)2(cid:12)(cid:12)(cid:12)(cid:12) f (z) − c1 c0(cid:12)(cid:12)(cid:12)(cid:12) dz dy. I3 = Z 1 0 (1 − x2)pαH (f, x)p dx. Пусть 0 ≤ x ≤ 1. Выберем число γ такое, что max(cid:26)α − 3 + 1 p ,−2 − 1 p(cid:27) < γ < min{0, α − 2}. p , получаем Применяя к внешнему интегралу неравенство Гельдера, учитывая, что γ > −2 − 1 H (f, x)p ≤ C4Z x ×(cid:26)Z x f (z) − dz(cid:27) (1 − y)pγ (cid:26)Z 1 (1 − y)(−3−γ) p ×Z x (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) p−1 dy(cid:27)p−1 (1 − y)pγ (cid:26)Z 1 ≤ C5(1 − x)p(−2−γ)−1 c1 p dz(cid:27) dy. f (z) − dy p y 0 0 0 y c1 c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) p , находим, что Применяя теперь неравенство Гельдера к внутреннему интегралу, учитывая, что γ > α − 3 + 1 H (f, x)p ≤ C5(1 − x)p(−2−γ)−1Z x (1 − y)pγ dz(cid:26)Z 1 ×Z 1 c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)p(α−γ)(cid:12)(cid:12)(cid:12)(cid:12) f (z) − ≤ C6(1 − x)p(−2−γ)−1Z x (1 − y)pγ Z 1 (1 − z)(2−α+γ) p p−1 dz(cid:27) f (z) − dz dy. dy p−1 0 p c1 p y y 0 y (1 − z)p(α−γ)(cid:12)(cid:12)(cid:12)(cid:12) c1 c0(cid:12)(cid:12)(cid:12)(cid:12) c0(cid:12)(cid:12)(cid:12)(cid:12) Отсюда получаем, что I3 ≤ C6Z 1 0 (1 − x)p(α−2−γ)−1Z x 0 (1 − y)pγ ×Z 1 y (1 − z)p(α−γ)(cid:12)(cid:12)(cid:12)(cid:12) f (z) − p c1 c0(cid:12)(cid:12)(cid:12)(cid:12) dz dy dx. О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 9 Поменяв пределы интегрирования, учитывая, что γ < α − 2 и γ < 0, имеем I3 ≤ C6Z 1 (1 − y)pγ pZ z c1 0 (1 − z)p(α−γ)(cid:12)(cid:12)(cid:12)(cid:12) f (z) − ×Z 1 ≤ C7Z 1 Учитывая, что f ∈ Lp,α и α > − 1 y 0 0 c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)pαz(cid:12)(cid:12)(cid:12)(cid:12) p имеем (1 − x)p(α−2−γ)−1 dx dy dz c1 p f (z) − c0(cid:12)(cid:12)(cid:12)(cid:12) dz ≤ C7(cid:13)(cid:13)(cid:13)(cid:13) f − . c1 c0(cid:13)(cid:13)(cid:13)(cid:13)p,α Обозначим −1 Учитывая равенство (3.2) имеем I3 < ∞. I4 = Z 0 (1 − x2)pαH (f, x) p dx. H (f, x) = Z x 0 (1 − y2)−3Z y −1 (1 − z2)2(cid:18)f (z) − c1 c0(cid:19) dz dy. Отсюда, при −1 ≤ x ≤ 0 имеем H (f, x)p ≤ C8Z 0 x (1 + y)−3Z y −1 (1 + z)2(cid:12)(cid:12)(cid:12)(cid:12) f (z) − dz dy. c1 c0(cid:12)(cid:12)(cid:12)(cid:12) Рассуждая как и при оценке I3, а именно применяя дважды неравенство Гельдера, потом меняя пределы интегрирования, получим, что I4 < ∞. Теперь I = kH (f, x)kp p,α = I3 + I4 < ∞. Таким образом доказано, что при 1 ≤ p < ∞ H (f, x) ∈ Lp,α. Пусть p = ∞. Обозначим J = max −1≤x≤1 (1 − x2)αH (f, x). Пусть Тогда J1 = max 0≤x≤1 (1 − x2)αH (f, x). J1 ≤ max 0≤x≤1 0 (1 − x2)αZ x (1 − y2)−3Z 1 c0(cid:13)(cid:13)(cid:13)(cid:13)∞,α ≤ (cid:13)(cid:13)(cid:13)(cid:13) max 0≤x≤1 f − c1 y c1 (1 − z2)2(cid:12)(cid:12)(cid:12)(cid:12) (1 − x2)αZ x c0(cid:12)(cid:12)(cid:12)(cid:12) f (z) − (1 − y2)−3Z 1 0 y dz dy J1 ≤ C9 max 0≤x≤1 (1 − x)αZ x 0 (1 − y)−3Z 1 y (1 − z)2−α dz dy. Учитывая, что f ∈ L∞,α, имеем (1 − z2)2−α dz dy. Отсюда, при 0 ≤ α < 3, находим J1 ≤ C10 max 0≤x≤1 (1 − x)αZ x 0 (1 − y)−α dy < ∞. Пусть J2 = max −1≤x≤0 (1 − x2)αH (f, x). 10 M. K. Потапов и Ф. M. Бериша Тогда по аналогии с оценкой для J1, учитывая равенство (3.2), имеем f − J2 ≤ (cid:13)(cid:13)(cid:13)(cid:13) c1 c0(cid:13)(cid:13)(cid:13)(cid:13)∞,α max −1≤x≤0 (1 + x)αZ 0 x (1 + y)−3Z y −1 (1 + z)2−α dz dy < ∞. Таким образом, для p = ∞ вытекает, что т.е. H (f, x) ∈ L∞,α. Лемма 3.9 полностью доказана. J = max{J1, J2} < ∞, Лемма 3.10. Пусть даны числа p и α такие, что 1 ≤ p ≤ ∞; − 1 p 1 < α < 3 − p 0 ≤ α < 3 dx H (f, x) ∈ Lp,α. при 1 ≤ p < ∞, при p = ∞. Тогда если f ∈ Lp,α, то d Доказательство. Из определения H (f, x) имеем d dx H (f, x) = −(1 − x2)−3Z 1 x (1 − z2)2(cid:18)f (z) − c1 c0(cid:19) dz. Пусть 1 ≤ p < ∞ и (cid:3) dx. Сначала рассмотрим случай p = 1. Рассмотрим d dx I = (cid:13)(cid:13)(cid:13)(cid:13) H (f, x)(cid:13)(cid:13)(cid:13)(cid:13) p p,α I1 = Z 1 0 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) Z 1 x (1 − z2)2(cid:18)f (z) − p c1 c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) x −1 = Z 1 Z 1 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) (1 − z2)2(cid:18)f (z) − dx c1 c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) (1 − x)α−3Z 1 ≤ C1Z 1 f (z) − Меняя пределы интегрирования, при −1 < α < 2 и f ∈ L1,α, имеем I1 ≤ C1Z 1 (1 − x)α−3 dx dz c1 x 0 0 (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) dz dx. c1 c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) 0 0 f (z) − ≤ C2Z 1 Z z c0(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)α(cid:12)(cid:12)(cid:12)(cid:12) Z 1 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) x Пусть I2 = Z 0 −1 < ∞. c1 c0(cid:13)(cid:13)(cid:13)(cid:13)1,α f (z) − c1 f (z) − c0(cid:12)(cid:12)(cid:12)(cid:12) dz ≤ C2(cid:13)(cid:13)(cid:13)(cid:13) c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) (1 − z2)2(cid:18)f (z) − c1 dx. Аналогично как и при оценке I1, учитывая равенство (3.2), получим I2 < ∞. Из того, что I1 < ∞ и I2 < ∞, следует, что I = I1 + I2 < ∞, т.е. d dx H (f, x) ∈ L1,α. О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 11 Пусть 1 < p < ∞. Рассмотрим I3 = Z 1 (1 − x2)p(α−3)(cid:12)(cid:12)(cid:12)(cid:12) Z 1 ≤ C3Z 1 x 0 p c1 dx (1 − z2)2(cid:18)f (z) − c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) (1 − x)p(α−3)(cid:26)Z 1 (1 − z)2(cid:12)(cid:12)(cid:12)(cid:12) x 0 p dz(cid:27) dx. f (z) − c1 c0(cid:12)(cid:12)(cid:12)(cid:12) Пусть α < γ < 3 − 1 интегрирования, получаем, что p . Применяя неравенство Гельдера, потом меняя пределы I3 ≤ C3Z 1 0 dz p−1 dx f (z) − p x x c1 f (z) − (1 − z)(2−γ) p (1 − x)p(α−3)Z 1 (1 − z)pγ (cid:12)(cid:12)(cid:12)(cid:12) c0(cid:12)(cid:12)(cid:12)(cid:12) ×(cid:26)Z 1 p−1 dz(cid:27) (1 − x)p(α−γ)−1Z 1 (1 − z)pγ (cid:12)(cid:12)(cid:12)(cid:12) pZ z (1 − z)pγ (cid:12)(cid:12)(cid:12)(cid:12) c0(cid:12)(cid:12)(cid:12)(cid:12) f (z) − = C5Z 1 (1 − z)pα(cid:12)(cid:12)(cid:12)(cid:12) f (z) − = C4Z 1 = C4Z 1 x c1 0 0 0 0 p , имеем (1 − x)p(α−γ)−1 dx dz dz dx p c1 c0(cid:12)(cid:12)(cid:12)(cid:12) dz ≤ C5(cid:13)(cid:13)(cid:13)(cid:13) p c1 c0(cid:12)(cid:12)(cid:12)(cid:12) f − . c1 c0(cid:13)(cid:13)(cid:13)(cid:13)p,α Отсюда, учитывая, что f ∈ Lp,α и α > − 1 I3 < ∞. Рассмотрим I4 = Z 0 −1 (1 − x2)p(α−3)(cid:12)(cid:12)(cid:12)(cid:12) Z 1 x (1 − z2)2(cid:18)f (z) − p dx. c1 c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) Аналогично как и при оценке I3, учитывая равенство (3.2), получим Теперь I4 < ∞. Таким образом доказано, что при 1 ≤ p < ∞ d Пусть теперь p = ∞. Рассмотрим dx H (f, x) ∈ Lp,α. I = I3 + I4 < ∞. J = max −1≤x≤1 (1 − x2)α(cid:12)(cid:12)(cid:12)(cid:12) При α < 3 имеем d dx H (f, x)(cid:12)(cid:12)(cid:12)(cid:12) = max J1 = max 0≤x≤1 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) c0(cid:13)(cid:13)(cid:13)(cid:13)∞,α ≤ (cid:13)(cid:13)(cid:13)(cid:13) f − c1 Отсюда, при f ∈ L∞,α и α ≥ 0 имеем x −1≤x≤1 Z 1 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) (1 − z2)2(cid:18)f (z) − (1 − x)α−3Z 1 c1 x Z 1 x max 0≤x≤1 (1 − z2)2(cid:18)f (z) − . c1 c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) (1 − z)2−α dz = C6(cid:13)(cid:13)(cid:13)(cid:13) f − . c1 c0(cid:13)(cid:13)(cid:13)(cid:13)∞,α J1 < ∞. 12 M. K. Потапов и Ф. M. Бериша Аналогично, с использованием равенства (3.2), при f ∈ L∞,α и 0 ≤ α < 3 находим, что J2 = max −1≤x≤0 Таким образом Z 1 (1 − x2)α−3(cid:12)(cid:12)(cid:12)(cid:12) x (1 − z2)2(cid:18)f (z) − c1 c0(cid:19) dz(cid:12)(cid:12)(cid:12)(cid:12) < ∞. J = max{J1, J2} < ∞, значит d dx H (f, x) ∈ L∞,α. Лемма 3.10 доказана. Лемма 3.11. Пусть f ∈ L1,2. Справедливые следующие равенства (l = 1, . . . , r − 1) c1 c0 Dl x,2,2H r (f, x) = H r−l (f, x) − Dr x,2,2H r (f, x) = f (x) − cr−l+1 c0 и , где cr−l+1 = R 1 −1(1 − z2)2H r−l (f, z) dz. Доказательство. Докажем сначала равенство (3.3). Для r = 1 имеем Dx,2,2H (f, x) = f (x) − c1 c0 . (cid:3) (3.3) равенство (3.3) доказывается по индукции. Теперь, учитывая, что по утверждении леммы 3.9 следует H r (f, x) ∈ L1,2, Из доказанного равенства (3.3) следует, что для l = 1, . . . , r − 1 имеем . x,2,2H r (f, x) = Dl cr−l+1 x,2,2H l(cid:0)H r−l (f, x) , x(cid:1) = H r−l (f, x) − Dl c0 Лемма 3.11 доказана. (cid:3) Лемма 3.12. Пусть даны числа p, α и r такие, что 1 ≤ p ≤ ∞, r ∈ N; − 1 p 1 < α < 3 − p 0 ≤ α < 3 при 1 ≤ p < ∞, при p = ∞. Тогда если f ∈ Lp,α, то H r (f, x) ∈ ADr(p, α). Доказательство. По лемме 3.9 имеем, что H r (f, x) ∈ Lp,α. Из условий леммы следует, что f ∈ L1,2 и H r (f, x) ∈ L1,2. Значит, постоянные cr (r = 1, 2, . . . ) в определении H r (f, x) определены. Рассмотрим сначала случай r = 1. По определению оператора H (f, x) ясно, dx H (f, x) (cid:22) абсолютно непрерывная функция на каждом отрезке [a, b] ⊂ что d (−1, 1). Далее, из леммы 3.11 вытекает, что Dx,2,2H (f, x) = f (x) − c1 c0 , и следовательно Dx,2,2H (f, x) ∈ Lp,α. Из лемм 3.9 и 3.10 следует, что H (f, x) ∈ Lp,α. Таким образом, H (f, x) ∈ AD1(p, α). Теперь, применяя формулу Лейбница и Лемму 3.11, утверждение леммы доказывается на основании индукции. (cid:3) Лемма 3.13. Пусть f ∈ L1,2. Справедливы равенства H r δ (f, x) = δ (H r (f, x) , x) + (r = 1, 2, . . . ), (3.4) 1 κ(δ)r ∆r c1 c0 где κ(δ) = Z δ 0 (cid:16)sin v 2(cid:17)−1(cid:16)cos v 2(cid:17)−9Z v 0 (cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv. О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 13 Доказательство. Докажем сначала равенство (3.4) для r = 1. По лемме 3.11 имеем f (x) = Dx,2,2H (f, x) + c1 c0 . Поэтому Hδ (f, x) = 1 κ(δ) Z δ 0 (cid:16)sin ×Z v 0 v v 2(cid:17)−9 2(cid:17)−1(cid:16)cos τu(cid:18)Dx,2,2H (f, x) + c1 c0 , x(cid:19)(cid:16)sin u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv. Поскольку из свойств оператора τu (f, x), отмеченных в лемме 3.2, следует, что τu(cid:18)Dx,2,2H (f, x) + c1 c0 , x(cid:19) = τu (Dx,2,2H (f, x) , x) + c1 c0 , то Hδ (f, x) = 1 κ(δ) Z δ v v 0 (cid:16)sin ×Z v 2(cid:17)−9 2(cid:17)−1(cid:16)cos τu (Dx,2,2H (f, x) , x)(cid:16)sin 0 u 2(cid:17)(cid:16)cos u 2(cid:17)9 du dv + c1 c0 . Применяя лемму 3.5, учитывая, что по лемме 3.12 функция H (f, x) имеет абсо- лютно непрерывную на каждом отрезке [a, b] ⊂ (−1, 1) производную d dx H (f, x), получаем Hδ (f, x) = 1 κ(δ) ∆δ (H (f, x) , x) + c1 c0 . Теперь для любого натурального r справедливость равенства (3.4) доказы- (cid:3) вается индукции, применяя леммы 3.11, 3.7 и 3.5. Следствие. Пусть f ∈ L1,2, тогда справедливы равенства Dr κ(δ)r ∆r Доказательство. По лемме 3.13 имеем δ (f, x) = x,2,2H r 1 δ (f, x) (r = 1, 2, . . . ). H r δ (f, x) = 1 κ(δ)r ∆r δ (H r (f, x) , x) + c1 c0 . Так как, из леммы 3.12 вытекает H r (f, x) ∈ ADr(p, α), то по лемме 3.7 полу- чаем, что Dr x,2,2H r δ (f, x) = 1 κ(δ)r Dr x,2,2∆r δ (H r (f, x) , x) Применяя лемму 3.11, находим = 1 κ(δ)r ∆r δ (cid:0)Dr x,2,2H r (f, x) , x(cid:1) . Dr x,2,2H r δ (f, x) = 1 κ(δ)r ∆r δ(cid:16)f − c1 c0 , x(cid:17) = 1 κ(δ)r ∆r δ (f, x) . Следствие доказано. (cid:3) 14 M. K. Потапов и Ф. M. Бериша Лемма 3.14. Пусть даны числа p, α, r и δ такие, что 1 ≤ p ≤ ∞, r ∈ N, 0 ≤ δ < π; 1 2 1 2p 1 − < α ≤ 1 3 2 − < α < 3 2 1 ≤ α < при p = 1, 1 2p при 1 < p < ∞, при p = ∞. Если f ∈ Lp,α, то H r Доказательство. Так как, в условиях леммы имеем f ∈ L1,2, то по лемме 3.13 δ (f, x) ∈ ADr(p, α). H r δ (f, x) = δ (H r (f, x) , x) + 1 κ(δ)r ∆r c1 c0 . По лемме 3.12 H r (f, x) ∈ ADr(p, α). Из леммы 3.6 следует, что H r δ (f, x) име- ет абсолютно непрерывную на каждом отрезке [a, b] ⊂ (−1, 1) производную d2r−1 dx2r−1 H r δ (f, x). Применяя теорему Лебега о предельном переходе под знаком интеграла, имеем, что для l = 1, . . . , r 1 δ (cid:0)Dl x,2,2H r (f, x) , x(cid:1) , δ (f, x) (cid:22) абсолютно непрерывная функ- Dl x,2,2H r δ (f, x) = κ(δ)r ∆r x,2,2H r т.е., опять по леммам 3.13 и 3.6, Dl ция на каждом отрезке [a, b] ⊂ (−1, 1). Из леммы 3.11 для l = 1, . . . , r имеем 1 Dl x,2,2H r δ (f, x) = κ(δ)r ∆r δ(cid:0)H r−l (f, x) , x(cid:1) . Теперь, применяя лемму 3.4 при фиксированном δ, учитывая, что по лемме 3.9 H r−l (f, x) ∈ Lp,α, имеем что Dl δ (f, x) ∈ Lp,α. δ (f, x) ∈ ADr(p, α). Тем самым лемма 3.14 доказана. (cid:3) Следовательно, H r x,2,2H r Лемма 3.15. Пусть даны числа p, α и r такие, что 1 ≤ p ≤ ∞, r ∈ N; 1 − 2 1 − 2p < α ≤ 2 5 < α < 2 − 5 2 при p = 1, 1 2p при 1 < p < ∞, 0 ≤ α < при p = ∞. Пусть f ∈ ADr(p, α). Тогда справедливо неравенство x,2,2f (x)(cid:13)(cid:13)p,α где постоянная C не зависит от f и n. n2r (cid:13)(cid:13)Dr En(f )p,α ≤ C 1 , Доказательство. Для r = 1 лемма доказана в работе [6]. Пусть Pn(x) (cid:22) алгебраический многочлен наилучшего приближения функ- ции Dx,2,2f (x), степени не выше, чем n − 1. Ясно, что многочлен Pn(x) можно представить в виде Пусть Pn(x) = g(x) = f (x) + n−1 Xk=0 n−1 Xk=0 λkP (2,2) k (x). λk k(k + 5) P (2,2) k (x). О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 15 Тогда по уже доказанному для r = 1 случаю леммы имеем [1, с.171] En (g)p,α ≤ C1 1 n2 kDx,2,2g(x)kp,α = C1 Dx,2,2f (x) − 1 n2(cid:13)(cid:13)(cid:13)(cid:13) n−1 Xk=0 λkP (2,2) k (x)(cid:13)(cid:13)(cid:13)(cid:13) = C1 1 n2 En (Dx,2,2f )p,α . Отсюда, учитывая, что f (x) − g(x) (cid:22) алгебраический многочлен степени не выше, чем n − 1, получаем En(f )p,α ≤ En (f − g)p,α + En (g)p,α ≤ C1 1 n2 En (Dx,2,2f )p,α . Теперь, применяя это неравенство r раз, получим, что En(f )p,α ≤ C2 Лемма 3.15 доказана. 1 n2r En(cid:0)Dr x,2,2f(cid:1)p,α ≤ C2 1 n2r (cid:13)(cid:13)Dr x,2,2f (x)(cid:13)(cid:13)p,α . (cid:3) 4. Основные утверждения Теорема 4.1. Пусть даны числа p, α и r такие, что 1 ≤ p ≤ ∞, r ∈ N; 1 2 1 2p 1 − < α ≤ 1 3 < α < 2 − 3 2 1 ≤ α < при p = 1, 1 2p при 1 < p < ∞, при p = ∞. Пусть f ∈ Lp,α. Тогда при всех δ ∈ [0, π) имеют место неравенства 2(cid:1)r Kr(f, δ)p,α, Kr(f, δ)p,α ≤ ωr(f, δ)p,α ≤ C2 C1(cid:18)cos4 δ 2(cid:19)r(r−1) где положительные постоянные C1 и C2 не зависят от f и δ. Доказательство. Для любой функции g(x) ∈ ADr(p, α) имеем (cid:0)cos4 δ 1 ωr(f, δ)p,α ≤ ωr (f − g, δ)p,α + ωr (g, δ)p,α . Применяя лемму 3.4, находим, что ωr (f − g, δ)p,α ≤ C3 Далее, в силу леммы 3.8 ωr (g, δ)p,α ≤ C4 1 2(cid:1)r kf − gkp,α . x,2,2g(x)(cid:13)(cid:13)p,α (cid:0)cos4 δ 2(cid:1)r δ2r (cid:13)(cid:13)Dr . 1 (cid:0)cos4 δ 2(cid:1)r (cid:16)kf − gkp,α + δ2r (cid:13)(cid:13)Dr 1 (cid:0)cos4 δ x,2,2g(x)(cid:13)(cid:13)p,α(cid:17) . Поэтому ωr(f, δ)p,α ≤ C5 Переходя в этом неравенстве к точной нижней грани по g(x) ∈ ADr(p, α), получаем правое неравенство теоремы. Для доказательства левого неравенства для данной функции f ∈ Lp,α рас- смотрим функцию где 1(f, x) = f (x). gδ,r(x) = (1 − (1 − H r δ )r) (f, x), 16 M. K. Потапов и Ф. M. Бериша Из леммы 3.14 следует, что H l δ (f, x) ∈ ADl(p, α) (l ∈ N). Поскольку δ )r = (cid:18)r k(cid:19)(−1)kH kr δ , r Xk=1 1 − (1 − H r то, учитывая, что ADkr(p, α) ⊆ ADr(p, α) (k = 1, . . . , r), получаем, что Оценим выражение gδ,r(x) ∈ ADr(p, α). (cid:13)(cid:13)Dr x,2,2gδ,r(x)(cid:13)(cid:13)p,α . Для этого, замечаем, что, поскольку H kr−l (f, x) (k = 2, . . . , r; l = 0, 1, . . . , r − 1) имеет на каждом отрезке [a, b] ⊂ (−1, 1) абсолютно непрерывную 2r − 1 производную, то применяя сначала теорему Лебега о предельном переходе под знаком интеграла, потом лемму 3.7, обобщенное неравенство Минковского и, наконец, лемму 3.4, получаем, что δ v v 1 (cid:13)(cid:13)Dr κ(δ) Z δ x,2,2H kr−1 x,2,2H kr δ (f, x)(cid:13)(cid:13)p,α ≤ ×Z v 0 (cid:13)(cid:13)τu(cid:0)Dr 2(cid:17)−9 2(cid:17)(cid:16)cos 2 (cid:13)(cid:13)Dr cos4 δ Применяя это неравенство k − 1 раз, получим, что 2(cid:17)−1(cid:16)cos 0 (cid:16)sin (f, x) , x(cid:1)(cid:13)(cid:13)p,α(cid:16)sin ≤ C6 u 1 δ du dv u 2(cid:17)9 x,2,2H kr−1 δ . (f, x)(cid:13)(cid:13)p,α (cid:13)(cid:13)Dr x,2,2H kr δ (f, x)(cid:13)(cid:13)p,α ≤ C7 1 2(cid:1)r(k−1) (cid:13)(cid:13)Dr (cid:0)cos4 δ x,2,2H r . δ (f, x)(cid:13)(cid:13)p,α Так как, gδ,r(x) представляет собой сумму членов содержащих H kr δ (f, x) (k = 1, . . . , r), то по последнему неравенству находим (cid:13)(cid:13)Dr x,2,2gδ,r(x)(cid:13)(cid:13)p,α ≤ C8 1 2(cid:1)r(r−1) (cid:13)(cid:13)Dr (cid:0)cos4 δ x,2,2H r . δ (f, x)(cid:13)(cid:13)p,α Применяя следствие из леммы 3.13, получаем (cid:13)(cid:13)Dr x,2,2gδ,r(x)(cid:13)(cid:13)p,α ≤ C8 Легко оценить, что при 0 < δ ≤ π 2 1 κ(δ)r (cid:0)cos4 δ 2(cid:1)r(r−1) k∆r δ (f, x)kp,α . κ(δ) ≥ C9δ2. Отсюда следует, что при 0 < δ ≤ π 2 δ2r(cid:13)(cid:13)Dr x,2,2gδ,r(x)(cid:13)(cid:13)p,α ≤ C10 С другой стороны 1 2(cid:1)r(r−1) ωr(f, δ)p,α. (cid:0)cos4 δ kf (x) − gδ,r(x)kp,α = kf (x) − (1 − (1 − H r δ )r) (f, x)kp,α = k(1 − H r δ )r(f, x)kp,α . Заметим, что (4.1) (4.2) 1 − H r δ = (1 − Hδ)(1 + Hδ + H 2 δ + ··· + H r−1 δ ). (4.3) О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 17 Теперь, применяя обобщенное неравенство Минковского и лемму 3.4, для l = 1, . . . , r − 1, имеем (cid:13)(cid:13)H l δ (f, x)(cid:13)(cid:13)p,α ≤ ×Z v 2(cid:17)−1(cid:16)cos 2(cid:17)−9 (f, x) , x(cid:1)(cid:13)(cid:13)p,α(cid:16)sin κ(δ) Z δ 0 (cid:16)sin 0 (cid:13)(cid:13)τu(cid:0)H l−1 u u 1 v v δ 2(cid:17)(cid:16)cos ≤ C11 du dv 2(cid:17)9 2 (cid:13)(cid:13)H l−1 1 δ cos4 δ (f, x)(cid:13)(cid:13)p,α . Применяя это неравенство l раз, получаем, что (cid:13)(cid:13)H l δ (f, x)(cid:13)(cid:13)p,α ≤ C12 1 (cid:0)cos4 δ 2(cid:1)l kf (x)kp,α . Поэтому, из равенства (4.3), применяя обобщенное неравенство Минковского, имеем, что k(1 − H r δ )(f, x)kp,α ≤ C13 1 v 1 (cid:0)cos4 δ κ(δ)(cid:0)cos4 δ 2(cid:1)r−1 k(1 − Hδ)(f, x)kp,α 2(cid:1)r−1 Z δ 2(cid:17)−1(cid:16)cos 0 (cid:16)sin ≤ C13 ×Z v 2(cid:17)9 2(cid:17)(cid:16)cos 0 kτu (f, x) − f (x)kp,α(cid:16)sin sup 0≤u≤δ k∆u (f, x)kp,α . 2(cid:1)r−1 (cid:0)cos4 δ 2(cid:17)−9 ≤ C14 du dv u u 1 v Применяя неравенство (4.4), из равенства (4.2) получим, что kf (x) − gδ,r(x)kp,α ≤ C14 1 (cid:0)cos4 δ 2(cid:1)r−1 sup 0≤t1≤δ(cid:13)(cid:13)∆t1 (cid:0)(1 − H r δ )r−1(f, x), x(cid:1)(cid:13)(cid:13)p,α . Замечаем, что меняя пределы интегрирования получаем (4.4) (4.5) τt (Hδ (f, x) , x) = Hδ (τt (f, x) , x) . Применяя это равенство r раз получим τt (H r δ (f, x) , x) = H r δ (τt (f, x) , x) . Отсюда очевидно, что ∆t ((1 − H r δ )(f, x), x) = (1 − H r δ )(∆t (f, x) , x). Применяя сначала это равенство, затем неравенство (4.5), потом неравен- ство (4.4), получим, что kf (x) − gδ,r(x)kp,α ≤ C15 1 (cid:0)cos4 δ 2(cid:1)2(r−1) sup 0≤t1≤δ sup 0≤t2≤δ(cid:13)(cid:13)∆2 t1,t2 (cid:0)(1 − H r δ )r−2(f, x), x(cid:1)(cid:13)(cid:13)p,α . 18 M. K. Потапов и Ф. M. Бериша Теперь, применяя r − 1 раз эту процедуру, получим, что kf (x) − gδ,r(x)kp,α ≤ C16 sup 1 (cid:0)cos4 δ 2(cid:1)r(r−1) t1,...,tr (f, x)(cid:13)(cid:13)p,α 0≤ti≤δ i=1,...,r(cid:13)(cid:13)∆r ≤ C16 2(cid:1)r(r−1) ωr(f, δ)p,α. (cid:0)cos4 δ 2 , из этого неравенства и неравенства (4.1) следует, что Таким образом, для 0 < δ ≤ π Iδ = kf (x) − gδ,r(x)kp,α + δ2r (cid:13)(cid:13)Dr x,2,2gδ,r(x)(cid:13)(cid:13)p,α ≤ C17 1 1 Поскольку для π Тем самим доказано левое неравенство теоремы для 0 < δ ≤ π 2 . Iδ ≤ π2r(cid:16)kf (x) − g1,r(x)kp,α + 1 ·(cid:13)(cid:13)Dr 2 ≤ δ < π имеем δ2 < π2 · 1 и 1 < π x,2,2g1,r(x)(cid:13)(cid:13)p,α(cid:17) ≤ C18 ωr (f, 1)p,α ≤ C18 1 2(cid:1)r(r−1) ωr(f, δ)p,α. (cid:0)cos4 δ 2 , то 2(cid:1)r(r−1) ωr(f, δ)p,α. (cid:0)cos4 δ Для δ = 0 левое неравенство теоремы тривиально. Итак для любого 0 ≤ δ < π доказано левое неравенство теоремы. Теорема 4.1 полностью доказана. (cid:3) Теорема 4.2. Пусть даны числа p, α и r такие, что 1 ≤ p ≤ ∞, r ∈ N; 1 2 1 2p 1 − < α ≤ 1 3 < α < 2 − 3 2 1 ≤ α < при p = 1, 1 2p при 1 < p < ∞, при p = ∞. Пусть f ∈ Lp,α. Тогда для любого натурального n справедливы неравенства C1En(f )p,α ≤ ωr (cid:18)f, 1 n(cid:19)p,α ≤ C2 1 n2r n ν2r−1Eν (f )p,α , Xν=1 где положительные постоянные C1 и C2 не зависят от f и n. Доказательство. Для любой функции g(x) ∈ ADr(p, α), применяя лемму 3.15, имеем En(f )p,α ≤ En (f − g)p,α + En (g)p,α ≤ kf − gkp,α + C3 где постоянная C3 не зависит от g и n. Отсюда, переходя к точной нижней граны по всем g(x) ∈ ADr(p, α), получим 1 n2r (cid:13)(cid:13)Dr x,2,2g(x)(cid:13)(cid:13)p,α , En(f )p,α ≤ C4Kr(cid:18)f, 1 n(cid:19)p,α . Применяя теорему 4.1, получаем, что En(f )p,α ≤ C5(cid:18)cos4 1 2n(cid:19)−r(r−1) ωr(cid:18)f, 1 n(cid:19)p,α ≤ C6 ωr(cid:18)f, 1 n(cid:19)p,α . Левое неравенство теоремы доказано. О СВЯЗИ МЕЖДУ r-ЫМ ОБОБЩЕННЫМ МОДУЛЕМ ГЛАДКОСТИ. . . 19 Докажем правое неравенство теоремы. Пусть Pn(x) алгебраический много- член наилучшего приближения для f , степени не выше, чем n − 1. Пусть k выбрано так, что Из теоремы 4.1, учитывая, что P2k (x) ∈ ADr(p, α), следует, что ωr(cid:18)f, n(cid:19)p,α ≤ C7(cid:18)cos 2n(cid:19)−4r 1 1 2k ≤ n < 2k+1. 1 Kr(cid:18)f, n(cid:19)p,α ≤ C8(cid:18)E2k (f )p,α + 1 n2r (cid:13)(cid:13)Dr x,2,2P2k (x)(cid:13)(cid:13)p,α(cid:19) . (4.6) Так как, Dr x,2,2P2k (x) = k−1 Xν=0 Dr x,2,2(P2ν+1(x) − P2ν (x)), то применяя r раз следствие из леммы 3.1, получаем (cid:13)(cid:13)Dr x,2,2P2k (x)(cid:13)(cid:13)p,α ≤ C9 k−1 Xν=0 22(ν+1)r kP2ν+1 − P2νkp,α ≤ 2C9 k−1 Xν=0 22(ν+1)rE2ν (f )p,α . Поэтому, учитывая неравенство (4.6), ωr (cid:18)f, 1 n(cid:19)p,α ≤ C10 1 n2r k Xν=0 22(ν+1)rE2ν (f )p,α . Теперь, замечая, что для ν = 1, . . . , k 2ν −1 Xj=2ν−1 j2r−1Ej (f )p,α ≥ 22(ν+1)r−4rE2ν (f )p,α , находим ωr(cid:18)f, 1 n(cid:19)p,α ≤ C11 1 n2r(cid:18)22rE1 (f )p,α + k 2ν −1 Xν=1 Xj=2ν−1 j2r−1Ej (f )p,α(cid:19) Теорема 4.2 доказана. ≤ C12 1 n2r n Xν=1 ν2r−1Eν (f )p,α . (cid:3) Список литературы [1] Г. Бейтмен, А. Эрдейи, Высшие трансцендентные функции, Москва, 1969. [2] P. L. Butzer, R. L. Stens, M. Wehrens, Higher order moduli of continuity based on the Jacobi translation operator and best approximation, C. R. Math. Rep. Acad. Sci. Canada 2 (1980), no. 2, 83 -- 88. [3] S. Pawelke, Ein Satz vom Jacksonschen Typ fur algebraische Polynome, Acta Sci. Math. (Szeged) 33 (1972), no. 3 -- 4, 323 -- 336. [4] М. К. Потапов, Некоторые неравенства для полиномов и их производных, Вестник МГУ, сер. мат. (1960), N. 2, 10 -- 20. [5] , О приближении алгебраическими многочленами в интегральной метрике с ве- сом Якоби, Вестник МГУ, сер. мат. (1983), N. 4, 43 -- 52. [6] M. K. Potapov, F. M. Berisha, Direct and inverse theorems of approximation theory for a generalized modulus of smoothness, Anal. Math. 25 (1999), no. 3, 187 -- 203. [7] М. К. Потапов, В. M. Федоров, О теоремах Джексона для обобщенного модуля гладко- сти, Тр. мат. ин.-та АН СССР, 172 (1985), 291 -- 298, 355. 20 M. K. Потапов и Ф. M. Бериша M. K. Потапов, Механико-математический факультет, Московский Государ- ственный Университет им. Ломоносова, Москва 117234, Россия Ф. M. Бериша, Механико-математический факультет, Московский Государствен- ный Университет им. Ломоносова, Москва 117234, Россия Current address: F. M. Berisha, Faculty of Mathematics and Sciences, University of Prishtina, Nena Tereze 5, 10000 Prishtina, Kosovo E-mail address: [email protected]
1711.03715
1
1711
2017-11-10T07:13:36
Bounded Point Evaluations For Certain Polynomial And Rational Modules
[ "math.FA" ]
Let $K$ be a compact subset of the complex plane $\mathbb C.$ Let $P(K)$ and $R(K)$ be the closures in $C(K)$ of analytic polynomials and rational functions with poles off $K,$ respectively. Let $A(K) \subset C(K)$ be the algebra of functions that are analytic in the interior of $K$. For $1\le t <\infty,$ let $P^t(1, \phi_1,...,\phi_N,K)$ be the closure of $P(K)+P(K)\phi_1+...+P(K)\phi_N$ in $L^t(dA|_K),$ where $dA|_K$ is the area measure restricted to $K$ and $\phi_1,...,\phi_N\in L^t(dA|_K).$ Let $HP(\phi_1,...,\phi_N,K)$ be the closure of $P(K)\phi_1+...+P(K)\phi_N +R(K)$ in $C(K),$ where $\phi_1,...,\phi_N\in C(K).$ In this paper, we prove if $R(K)\ne C(K),$ then there exists an analytic bounded point evaluation for both $P^t(1, \phi_1,...,\phi_N,K)$ and $HP(\phi_1,...,\phi_N,K)$ for certain smooth functions $\phi_1,...,\phi_N,$ in particular, for $\bar z,\bar z^2,...,\bar z^N.$ We show that $A(K)\subset HP(\bar z,\bar z^2,...,\bar z^N,K)$ if and only if $R(K) = A(K).$ In particular, $C(K) \ne HP(\bar z,\bar z^2,...,\bar z^N,K)$ unless $R(K) = C(K).$ We also give an example of $K$ showing the results are not valid if we replace $\bar z^n$ by certain $\phi_n,$ that is, there exist $K$ and a function $\phi\in A(K)$ such that $R(K) \ne A(K),$ but $A(K) = HP (\phi ,K).$
math.FA
math
Bounded Point Evaluations For Certain Polynomial And Rational Modules Liming Yang Department of Mathematics Virginia Polytechnic and State University Blacksburg, VA 24061 [email protected] Abstract Let K be a compact subset of the complex plane C. Let P (K) and R(K) be the closures in C(K) of analytic polynomials and rational functions with poles off K, respectively. Let A(K) ⊂ C(K) be the algebra of functions that are analytic in the interior of K. For 1 ≤ t < ∞, let P t(1, φ1, ..., φN , K) be the closure of P (K) + P (K)φ1 + ... + P (K)φN in Lt(dAK), where dAK is the area measure restricted to K and φ1, ..., φN ∈ Lt(dAK). Let HP (φ1, ..., φN , K) be the closure of P (K)φ1 + ... + P (K)φN + R(K) in C(K), where φ1, ..., φN ∈ C(K). In this paper, we prove if R(K) 6= C(K), then there exists an analytic bounded point evaluation for both P t(1, φ1, ..., φN , K) and HP (φ1, ..., φN , K) for certain smooth functions φ1, ..., φN , in particular, for ¯z, ¯z2, ..., ¯zN . We show that A(K) ⊂ HP (¯z, ¯z2, ..., ¯zN , K) if and only if R(K) = A(K). In particular, C(K) 6= HP (¯z, ¯z2, ..., ¯zN , K) unless R(K) = C(K). We also give an example of K showing the results are not valid if we replace ¯zn by certain φn, that is, there exist K and a function φ ∈ A(K) such that R(K) 6= A(K), but A(K) = HP (φ, K). 1 Introduction Let P denote the set of polynomials in the complex variable z. For a compact subset K of the complex plane C, let Rat(K) be the set of all rational functions with poles off K and let C(K) denote the Banach algebra of complex-valued continuous functions on K with customary norm k.kK . Let P (K) and R(K) denote the closures in C(K) of P and Rat(K), respectively. Let A(K) ⊂ C(K) be the algebra of functions that are analytic in the interior of K. For φ1, ..., φN ∈ C(K), let HP (φ1, ..., φN , K) denote the closure of P (K)φ1 + ... + P (K)φN + R(K) in C(K). For 1 ≤ t < ∞, let Lt(K) = Lt(dAK), where dAK is the area measure restricted to K. For φ1, ..., φN ∈ Lt(K), let P t(1, φ1, ..., φN , K) be the closure of P (K) + P (K)φ1 + ... + P (K)φN in Lt(K). For a subset A ⊂ C, we set Int(A) for its interior, ¯A or clos(A) for its closure, Ac for its complement, and χA for its characteristic function. For a subspace A of C(K) and a function f ∈ C(K), we define the distance from f to A by dist(f, A) = inf g∈A kf − gkK . For a subspace B of Lt(K) and a function f ∈ Lt(K), we define the distance from f to B by dist(f, B) = inf g∈B kf − gkLt(K). Set The open unit disk is denoted by D = B(0, 1). The constants used in the paper such as C, C0, C1, CN , δ0, δ1, δN , ǫ0, ǫ1, ǫN , ... may change from one step to the next. B(λ0, δ) = {z : z − λ0 < δ}. We denote the Riemann sphere C∞ = C ∪ {∞}. For a compact subset E ⊂ C, we define the analytic capacity of E by γ(E) = sup f∈A(E)f′(∞), 1 where A(E) consists of those functions f analytic in C∞ \ E for which f (∞) = 0, f (z) ≤ 1 for all z ∈ C∞ \ E, and f′(∞) = lim z→∞ z(f (z) − f (∞)). The analytic capacity of a general E1 ⊂ C is defined to be The continuous analytic capacity for a compact subset E is defined similarly as γ(E1) = sup{γ(E) : E ⊂ E1, E compact}. α(E) = sup f∈AC(E)f′(∞), where AC(E) = A(E) ∩ C(C∞). For a general E1 ⊂ C, α(E1) = sup{α(E) : E ⊂ E1, E compact}. (see Gamelin (1969) and Conway (1991) for basic information of rational approximation and analytic capacity). Let ν be a compactly supported finite measure on C. The Cauchy transform of ν is defined by Cν(z) =Z 1 w − z dν(w) for all z ∈ C for whichR dν(w) loc(C) for 0 < s < 2, in particular that it is defined for Area almost all z, and clearly Cν is analytic in C∞ \ sptν. < ∞. A standard application of Fubini's Theorem shows that Cν ∈ Ls w−z We denote the map Et(λ) : p0 + N Xi=1 piφi →  p0(λ) p1(λ) ... pN (λ)   , (1-1) where p0, p1, ..., pN ∈ P. If E(λ) is bounded from P t(1, φ1, ..., φN , K) to (CN+1,k.kN+1), where kxkN+1 = PN i=0 xi for x ∈ CN+1, then every component in the right hand side extends to a bounded linear functional on P t(1, φ1, ..., φN , K) and we will call λ a bounded point evaluation for P t(1, φ1, ..., φN , K). A bounded point evaluation λ0 is called an analytic bounded point evaluation for P t(1, φ1, ..., φN , K) if there is a neighborhood B(λ0, δ) of λ0 such that every λ ∈ B(λ0, δ) is a bounded point evaluation and Et(λ) is analytic as a function of λ on B(λ0, δ) (equivalently (1-1) is uniformly bounded for λ ∈ B(λ0, δ)). Similarly, we can define a bounded point evaluation (or analytic bounded point evaluation) λ for HP (φ1, ..., φN , K)by replacing (1-1) with the following map: E(λ) : r + N Xi=1 piφi →  p1(λ) p2(λ) ... pN (λ)   , (1-2) where p1, p2, ..., pN ∈ P and r ∈ Rat(K). Notice that the rational function r ∈ Rat(K) does not appear on the right hand side of definition (1-2). For an arbitrary finite compactly supported positive measure µ, Thomson (1991) describes com- pletely the structure of P t(µ), the closed subspace of Lt(µ) spanned by P. Conway and Elias (1993) extends some results of Thomson's Theorem to the space Rt(K, µ), the closure of Rat(K) in Lt(µ), while Brennan (2008) expresses Rt(K, µ) as a direct sum that includes both Thomson's theorem and results of Conway and Elias (1993). For a compactly supported complex Borel measure ν of C, by estimat- ing analytic capacity of the set {λ : Cν(λ) ≥ c}, Brennan (2006. English), Aleman et al. (2009), and Aleman et al. (2010) provide interesting alternative proofs of Thomson's theorem. Both their proofs rely on X. Tolsa's deep results on analytic capacity. Yang (2018) extends some results to a rationally multi- cyclic subnormal operator (restriction of a normal operator on a separable Hilbert space to an invariant subspace). However, even for µ = dAK, it is difficult to obtain necessary and sufficient conditions under which P t(1, K) = Lt(K) or P t(1, K) has a bounded point evaluation. Brennan and Militzer (2011) proves if R(K) 6= C(K), then P t(1, K) has a bounded point evaluation. Yang (2016) shows that there exists a compact subset K ⊂ C with R(K) = C(K), but P t(1, K) still has bounded point evaluations. The first part of this paper is to extend the above result of Brennan and Militzer (2011) to P t(1, φ1, ..., φN , K) and 2 HP (φ1, ..., φN , K) for some smooth functions φ1, ..., φN . The following theorem, which connects analytic capacity with special types of bounded point evaluation estimations, is essential for our main results. Notice that Theorem 1 (1) extends Lemma B in Aleman et al. (2009). Theorem 1. There exist absolute constants ǫN , CN > 0 that only depend on N . If then (1) γ(B(λ0, δ) \ K) < ǫN δ, (2) pN (λ) ≤ δN+2 ZK∩ ¯B(λ0,δ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) δN (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=1 where r ∈ Rat(K ∩ ¯B(λ0, δ)) and pk ∈ P. pk ¯zk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) pk ¯zk + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩ ¯B(λ0 ,δ) pN (λ) ≤ N Xk=0 CN N CN dA, λ ∈ ¯B(cid:18)λ0, 1 2 δ(cid:19) , , λ ∈ ¯B(cid:18)λ0, 1 2 δ(cid:19) , (1-3) (1-4) (1-5) The proof of Theorem 1 depends on a careful modification of Thomson's coloring scheme on dyadic squares. Thomson's coloring scheme, for a point a ∈ C and a positive integer m, starts with a dyadic square of side length 2−m containing a and either terminates at some finite stage or produces an infinite sequence of annuli surrounding a. These annuli are made up of dyadic squares colored red (heavy square). When the scheme terminates at some finite stage, one can find a path consisting of many dyadic squares colored green (light square). The definition of a light square in Thomson (1991) (see also page 168 of Thomson (1993) or page 461 of Aleman et al. (2009)) only works for P t(1, K) and HP (¯z, K). Our definition of a light square (2-2) allows us to recursively extend (1-4) and (1-5) for N > 1. BA(λ0, δ) = {f ∈ C(C∞) : ∃n, f ∈ C (n)( ¯B(λ0, δ)), ¯∂nf ¯B(λ0,δ) = 0}, where ¯∂ is the Cauchy-Riemann operator. We now state our first main result. Theorem 2. Let λ0 ∈ K be a nonpeak point for R(K) and 1 ≤ t < ∞. If there exist δ > 0 and F1, F2, ..., FN ∈ BA(λ0, δ) such that det[ ¯∂iFj (λ0)]n×n 6= 0, Define then (1) λ0 is an analytic bounded point evaluation for P t(1, F1, ..., FN , K). In particular, λ0 is an analytic bounded point evaluation for P t(1, ¯z, ..., ¯zN , K). (2) λ0 is an analytic bounded point evaluation for HP (F1, ..., FN , K). In particular, λ0 is an analytic bounded point evaluation for HP (¯z, ..., ¯zN , K). Let Λ be a constant coefficient elliptic differential operator in R2. For a compact K ⊂ C, let H(K, Λ) and h(K, Λ) denote the uniform closures in C(K) of the set {f ∈ K : Λf = 0 in some neighborhood of K} and the set respectively. Notice that, for Λ = ¯∂, the space H(K, ¯∂) = R(K) and h(K, ¯∂) = A(K). For Λ = ¯∂n, the nth power of Cauchy-Riemann operator, the space C(K) ∩ {f ∈ K : Λf = 0 in the interior of K} H(K, ¯∂n) = clos(R(K) + ¯zR(K) + ... + ¯zn−1R(K)) and h(K, ¯∂n) = clos(A(K) + ¯zA(K) + ... + ¯zn−1A(K)). One of uniform approximation problems is the following: Problem 1. Find necessary and sufficient conditions for K so that H(K, Λ) = h(K, Λ). A complete solution for Λ = ∆ was obtained by Deny (1949) and Keldysh (1966) using a duality argument relying on classical potential theory. Let Cap denote the Wiener capacity in potential theory. Deny and Keldysh show that the identity H(K, ∆) = h(K, ∆) occurs if and only if for each open ball B one has Cap(B \ intK) = Cap(B \ K). 3 Using a constructive scheme for uniform approximation (based on a localization operator), Vitushkin (1967) proves that the identity H(K, ¯∂) = h(K, ¯∂) occurs if and only if for each open disc O one has α(O \ intK) = α(O \ K). The inner boundary of K, denoted by ∂iK, is the set of boundary points which do not belong to the boundary of any connected component of C \ K. The remarkable paper Tolsa (2004) proves that the continuous analytic capacity is semiadditive. The result implies an affirmative answer to the so called inner boundary conjecture (see Vitushkin and Melnikov (1984), Conjecture 2). That is, if α(∂iK) = 0, then R(K) = A(K). For Λ = ¯∂2, Trent and Wang (1981) show if K is a compact subset without interior, then H(K, ¯∂2) = h(K, ¯∂2) = C(K). Verdera (1993) proves that each Dinicontinuous function in h(K, ¯∂2) belongs to H(K, ¯∂2). Finally, the excellent paper Mazalov (2004) completely solved the problem by proving H(K, ¯∂2) = h(K, ¯∂2) for any compact subset K. In Baranova et al. (2016), the authors consider an interesting analogous problem: find necessary and sufficient conditions so that P (K) + P (K)¯zn is dense in A(K) + A(K)¯zn. The paper obtained some results for a Caratheodory compact set K with n ≥ 2 (see Baranova et al. (2016), Theorem 1). We define hP (φ1, φ2, ..., φN , K) = clos N Xi=1 P (K)φi + A(K)! . As analogous to Problem 1, we are interested in the following problem: Problem 2. Find necessary and sufficient conditions so that HP (φ1, φ2, ..., φN , K) = hP (φ1, φ2, ..., φN , K). (1-6) For N = 1, the problem was studied by several authors. In Thomson (1993), Thomson proves if R(K) 6= C(K), then HP (¯z, K) is not equal to C(K). Yang (1995) and Yang (1994) study the generalized space HP (g, K) and prove that for a smooth function g with ¯∂g 6= 0, then HP (g, K) = hP (g, K) if and only if A(K) = R(K). In the second part of this paper, we will study Problem 2 when N > 1. Our second main theorem is the following: Theorem 3. Let q1(z, ¯z), q2(z, ¯z), ..., qN (z, ¯z)) be polynomials in two variables z and ¯z. If A(K) ⊂ HP (q1, ..., qN , K), then A(K) = R(K). In particular, if R(K) 6= C(K), then HP (q1, ..., qN , K) 6= C(K). Notice that an important special case is that q1(z, ¯z) = ¯z, q2(z, ¯z) = ¯z2, ...,qN (z, ¯z) = ¯zN . By Stone-Weierstrass theorem,P∞k=1 P (K)¯zk +R(K) is dense in C(K). So it is critical here to assume In Baranova et al. (2016), the authors are interested in the question of finding necessary and sufficient that N is a finite integer. conditions so that clos A(K) + N Xi=1 A(K)¯zdi! = clos P (K) + P (K)¯zdi! , N Xi=1 (1-7) where d1, ..., dN are positive integers. Our theorem implies if (1-7) holds, then A(K) = R(K). So (1-7) is equivalent to clos R(K) + R(K)¯zdi! = clos P (K) + N Xi=1 P (K)¯zdi! . N Xi=1 The following result shows that Theorem 3 will not hold if we replace qn by certain functions. Proposition 1. There exist a compact subset K ⊂ C and a function φ ∈ A(K) such that R(K) 6= A(K), but A(K) = HP (φ, K). The proposition raises the following question: Problem 3. For a compact subset K of C, is there a function φ ∈ A(K) such that A(K) = HP (φ, K)? It seems that one might be able to use Tolsa (2004) characterization of continuous analytic capacity to find a proper finite Borel measure ν supported on the inner boundary of K such that A(K) = HP (Cν, K). We prove Theorem 1 and Theorem 2 in Section 2. In section 3, we prove Theorem 3 and construct a compact subset K to prove Proposition 1. 4 2 Bounded Point Evaluations In this section, we will prove Theorem 1 and Theorem 2. Tolsa (2003) proves the astounding result about analytic capacity γ that implies the semiadditivity of analytic capacity. That is, γ( m [i=1 Ei) ≤ CT γ(Ei) m Xi=1 (2-1) where CT is an absolute constant. For a square S (also denoted by S(cS, dS)), whose edges are parallel to x-axis and y-axis, let cS denote the center and dS denote the side length. For a > 0, aS is a square with the same center of S (caS = cS) and the side length daS = adS. For a given ǫ > 0, a closed square S is defined to be light ǫ if (2-2) A square is called heavy ǫ if it is not light ǫ. Let R = {z : −1/2 < Re(z), Im(z) < 1/2} and Q = ¯D \ R. We now sketch our version of Thomson's coloring scheme for Q with a given ǫ and a positive integer m. We refer the reader to Thomson (1991) and Thomson (1993) section 2 for details. γ(Int(S) \ K)) > ǫdS. For each integer k > 3 let {Skj} be an enumeration of the closed squares contained in C with edges of length 2−k parallel to the coordinate axes, and corners at the points whose coordinates are both integral multiples of 2−k (except the starting square Sm1, see (3) below). In fact, Thomson'scoloring scheme is just needed to be modified slightly as the following: (1) Use our definition of a light ǫ square (2-2). (2) A path to ∞ means a path to any point that is outside of Q (replacing the polynomially convex (3) The starting yellow square Sm1 in the m-th generation is R. Notice that the length of Sm1 in hull of Φ by Q). m-th generation is 1 (not 2−m). We will borrow notations that are used in Thomson's coloring scheme such as {γn}n≥m and {Γn}n≥m, etc. We denote Two things can happen (depending on m): Y ellowBuf f erm = ∞ Xk=m+1 k22−k. Case I. The scheme terminates, in our setup, this means Thomson's coloring scheme reaches a square S in n-th generation that is not contained in Q. One can construct a polygonal path P, which connects the centers of adjacent squares, from the center of a square (contained in Q) adjacent to S to the center of a square adjacent to R so that the orange (non green in Thomson's coloring scheme) part of length is no more than Y ellowBuf f erm. Let GP = ∪Sj , where {Sj} are all light ǫ squares with P ∩ Sj 6= ∅. By Tolsa's Theorem (2-1), we see γ(P ) ≤ CT (γ(Int(GP )) + Y ellowBuf f erm). Since P is a connected set, γ(P ) ≥ 0.1 to be large enough so that 4 (Theorem 2.1 on page 199 of Gamelin (1969)). We can choose m γ(Int(GP )) ≥ 1 40CT − Y ellowBuf f erm = ǫm > 0. (2-3) Case II. The scheme does not terminate. In this case, one can construct a sequence of heavy ǫ barriers inside Q, that is, {γn}n≥m and {Γn}n≥m are infinite. coloring scheme. For simplicity, we will use scheme(Q, ǫ, m, γn, Γn, n ≥ m) to stand for our version of Thomson's If a function f is analytic at ∞, then f can be represented by its Laurent series f (z) = f (∞) + a1(z − z0)−1 + a2(z − z0)−2 + ... in a neighborhood of infinite. We define f′(∞) to be a1 and β(f, z0) to be a2. The number f′(∞) does not depend on z0, but β(f, z0) does depend on the choice of z0. Lemma 1. For a square T, if then for two complex numbers α ≤ 1 and β ≤ 1, there exists a function f in C(C∞) such that the following hold: γ(Int(T ) \ K) ≥ ǫ1dT , 5 ; ǫ3 1 (1) kfk ≤ 54 (2) f ∈ R(C∞ \ (Int(T ) \ K)) ; (3) f (∞) = 0; (4) f′(∞) = αdT ; (5) β(f, cT ) = βd2 T . Proof: There exists a function f1 in C(C∞) such that kf1k ≤ 1, f1 is analytic off a compact subset of Int(T )\ K, f1(∞) = 0, and f′1(∞) > ǫ1dT /2. Set f2 = dT f1/f′1(∞). Then by Theorem 2.5 of Gamelin (1969) on page 201, we get β(f2, 0) ≤ 12 ǫ1 d2 T . Let 2(cid:19) + βf 2 then f ∈ R(C∞ \ (Int(T ) \ K)) and satisfies the conditions (1)-(5). f = α(cid:18)f2 − β(f2, 0) d2 T f 2 2 , Let ϕ be a smooth function with compact support. The localization operator Tϕ is defined by (Tϕf )(λ) = 1 π Z f (z) − f (λ) z − λ ¯∂ϕ(z)dA(z), where f is a continuous function on C∞. One can easily prove the following norm estimation for Tϕ : Lemma 2. Suppose Case I of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true, then kTϕfk ≤ 4kfkdiameter(suppϕ)k ¯∂ϕk. where and ǫm is from (2-3). γ(D \ K) ≥ ǫ1, ǫ1 = 10−8ǫ3ǫm (2-4) (2-5) Proof: we will follow the second part of proof of Lemma B in Aleman et al. (2009) on pages 464- 465 with slight modifications. Let GP = ∪Sj, where {Sj} are light ǫ squares discussed above, so that γ(Int(GP )) ≥ ǫm. For each j let zj be the center of Sj , dj be the edge length of Sj, Qj, Rj be the closed squares with center zj and sides parallel to the coordinate axes of lengths 7 3 dj respectively. The collection {Sj} has the following properties (see (2.16)-(2.18) on page 464 of Aleman et al. (2009)): , φj = 1 on Rj , and (a) No point lies in more than four Qj 's. (b) There are C∞ functions φj with 0 ≤ φj ≤ 1, spt(φj) ⊂ Int(Qj), k ¯∂φjk ≤ 50 (c) For each z ∈ C, P φj = 1 on GP. 6 dj = δj, 2 dj for all z ∈ C and Tφj f (∞) = 0, kTφj fk ≤ 400, (Tφj f )′(∞) ≤ 400dj , and β(Tφj f, zj) ≤ 400d2 j . f )′(∞) 400dj , using Lemma 1 for the light ǫ square Sj, we find a function and β = β(Tφj f,zj ) (Tφj 400d2 j For α = fj in C(C∞) such that kfjk ≤ 54 Therefore, ( if z ∈ C∞ \ Qj, then Tφj Tφj 400 − fj has triple zeros at ∞. 400 − fj )(z − zj)3 is analytic on C∞ \ Qj. Using the Maximum Modulus Theorem, we see ǫ3 , fj ∈ R(C∞ \ (Int(Sj) \ K)), and f f (cid:12)(cid:12)(cid:12)(cid:12) Tφj f (z) 400 − fj (z)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (1 + 54 ǫ3 ) δ3 j z − zj3 ≤ 55 ǫ3 δ3 j z − zj3 . 6 Let f ∈ C(C∞) such that f is analytic off a compact subset of Int(GP ), f (∞) = 0, kfk∞ = 1, and f′(∞) > ǫm 2 . Then from (b), we see that f −P Tφj f is zero on GP and analytic off GP. Hence, X min(cid:26)1, δ3 j z − zj3(cid:27) ≤ 10, 000. f (z) =X Tφj f (z) Set F = 400P fj, then F is analytic off a compact subset of Int(GP ) \ K, F (∞) = 0, F ′(∞) = f′(∞), and Hence, for z ∈ C∞, (cid:12)(cid:12)(cid:12)(cid:12) Tφj f (z) δ3 j 55 ≤ z − zj3(cid:19) . ǫ3 min(cid:18)1, 400 − fj(z)(cid:12)(cid:12)(cid:12)(cid:12) kFk∞ ≤kfk∞ +X kTφj f − 400fjk∞ z − zj3(cid:19) ǫ3 X min(cid:18)1, 2200 δ3 j ≤1 + 5 × 107 where the last step follows from (c). Therefore, ≤ ǫ3 . γ(D \ K) ≥ γ(Int(GP ) \ K) ≥ F ′(∞) 5×107 ǫ3 > 10−8ǫ3ǫm = ǫ1. Lemma 3. Suppose Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true. (1) If there exists ǫN > 0 such that for every heavy ǫ square S in scheme(Q, ǫ, m, γn, Γn, n ≥ m), then there exists a constant Cm,N (depends on m and N ) such that S dist(cid:16)¯zN , P 1(1, ¯z, ..., ¯zN−1, K ∩ S)(cid:17) ≥ ǫN dN+2 pN (λ) ≤ Cm,N(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩Q) , λ ∈ R, Xj=0 N , (2-6) (2-7) (2-8) where pj ∈ P and N = 1, 2, .... For N = 0, if (2-6) is replaced by Area(K ∩ S) ≥ ǫ0d2 S, then (2-7) holds for N = 0. (2) If there exists ǫN > 0 such that for every heavy ǫ square S in scheme(Q, ǫ, m, γn, Γn, n ≥ m), then there exists a constant Cm,N (depends on m and N ) such that S , dist(cid:16)¯zN , HP (¯z, ¯z2, ..., ¯zN−1, K ∩ S)(cid:17) ≥ ǫN dN pN (λ) ≤ Cm,N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pj ¯zj + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩Q , λ ∈ R, N Xj=1 where r ∈ Rat(K ∩ Q), pj ∈ P, and N = 1, 2, .... (2-9) (2-10) Proof: The proofs of section 4 in Thomson (1991) and Thomson (1993) will work if we make the following modifications: (a) For w ∈ γn ∩ S, where S is a heavy ǫ square, by the Hahn-Banach theorem, there exists a finite Borel measure σw supported in K∩S and kσwk = 1 such that for (1), σw = ΦwdA with Φw ∈ L∞(K∩S), (2-6) becomes dist(cid:16)¯zN , P 1(1, ¯z, ..., ¯zN−1, K ∩ S)(cid:17) =Z ¯zN dσw ≥ ǫN dN+2 S , where R f dσw = 0 for f ∈ P 1(1, ¯z, ..., ¯zN−1, K ∩ S), and (2-8) becomes σw = χK∩SdA, Area(K ∩ S) =Z dσw ≥ ǫ0d2 S; and for (2), (2-9) becomes dist(cid:16)¯zN , HP (¯z, ¯z2, ..., ¯zN−1, K ∩ S)(cid:17) =Z ¯zN dσw ≥ ǫN dN S , where R f dσw = 0 for f ∈ HP (¯z, ¯z2, ..., ¯zN−1, K ∩ S). Define τw = ¯zN dσw R ¯zN dσw . 7 (b) Set L = eλ (eλ(f ) = f (λ)) for λ ∈ R. Use the same argument as in section 4 in Thomson (1993), one can construct µn+q defined by linear combination of τw, then for (1) and kµn+qk ≤ ǫ−q−1 N 2(N+2)(n+q)4q((n + q − 1)...n)−2 for (2). Let µ = µn + ... + µn+q + ..., then kµk ≤ Cm,N , for (1) kµn+qk ≤ ǫ−q−1 N 2N(n+q)4q((n + q − 1)...n)−2 pN (λ) =Z N Xj=0 pj ¯zj dµ, and for (2) pN (λ) =Z ( N Xj=1 Clearly, the support of µ is outside R. The proof is completed. pj ¯zj + r)dµ. The idea to prove our Theorem 1 is to find sufficient small ǫ so that Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true. Then we use mathematical induction to show that for every heavy ǫ square, we can find ǫN so that (2-6), (2-8), and (2-9) all hold. We will demonstrate the idea by proving the following corollary, which is the case for N = 0 in (1) of Theorem 1 and Lemma B in Aleman et al. (2009). Corollary 1. There are absolute constants ǫ0 > 0 and C0 < ∞ with the following property. For a compact subset K ⊂ C, let R > 0 and γ(RD \ K) < ǫ0R. Then R2 Z(RD)∩K p p(λ) ≤ dA π C0 2 and all p ∈ P. Proof: Since γ(RD \ K) = Rγ(D \ K for λ ≤ R R ), by a simple changing of variables from z to Rz, we assume R = 1. Let ǫ1 = ǫ0 = 10−8ǫ3ǫm in (2-5) with ǫ < q 1 4π . Then from Lemma 2, we conclude that Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) must be true. Let S be a heavy ǫ square, then γ(Int(S) \ K) ≤ ǫdS. By Theorem 3.2 on page 204 of Gamelin (1969), we get Area(S \ K) ≤ 4πγ(Int(S) \ K)2 ≤ 4πǫ2d2 S. Therefore, Area(S ∩ K) ≥ (1 − 4πǫ2)d2 S. So (2-8) holds and the corollary follows from Lemma 3. Let φ be a smooth function supported in D such that: 0 ≤ φ ≤ 1, φ(z) = φ(z), k ¯∂N φk < C φ N , Z φdA = 1. (2-11) Proof of Theorem 1 (1): We only need to prove the case that λ0 = 0 and δ = 1. In fact, using the elementary properties of analytic capacity (see p. 196 of Gamelin (1969)), one sees that condition (1-3) is equivalent to The inequality (1-4) N CN pN (λ) ≤ γ(cid:18)B(0, 1) \ δN+2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=0 δ2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=0 p0N (λ) ≤ CN (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p0k ¯zk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1( Xk=0 qk( CN = N N δ K − λ0 (cid:19) < ǫN . pk ¯zk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) )k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) , λ ∈ ¯B(cid:18)0, ¯z − ¯λ0 K−λ0 δ ∩ ¯B(0,1)) δ , 8 1 2(cid:19) , for λ ∈ ¯B(λ0, 1 to 2 δ), where qN = pN , qk(1 ≤ k ≤ N − 1) are certain linear combinations of pk, is equivalent where p0k(z) = qk(δz + λ0). We will assume that λ0 = 0 and δ = 1 in the rest of the proof. We use mathematical induction for N. The case N = 0 is directly implied by Corollary 1. Now we assume that Theorem 1 (1) holds for k = 0, 1, 2, ..., N. Set ǫ = min(c1, ǫN 2 , ǫN−1 22 , ..., ǫ1 2N , ǫ0 2N+1 ), where c1 > 0 will be determined later, and ǫN+1 as in (2-5), that is, ǫN+1 = 10−8ǫ3ǫm. Since γ(B(0, 1) \ K) < ǫN+1 (assumption (1-3)), from Lemma 2, we conclude that Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true. For every heavy ǫ square S, we need to prove (2-6) holds. We assume otherwise (2-6) already holds. Without loss of generality, we assume the center of S is zero. There are polynomials p0, p1, ..., pN such that 1 2 dN+3 S , N dist(cid:16)¯zN+1, P 1(1, ¯z, ..., ¯zN , K ∩ S)(cid:17) < pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩S) pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩S) Xj=0 ≤ dN+3 Xj=0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N S + k¯zN+1kL1(K∩S) ≤ 2dN+3 S . ≤ 2dist(cid:16)¯zN+1, P 1(1, ¯z, ..., ¯zN , K ∩ S)(cid:17) < dN+3 S . (2-12) ¯zN+1 + (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Hence, Since S is a heavy ǫ square, we get γ(B(0, dS 2 ) \ K) ≤ γ(S \ K) < ǫdS ≤ ǫN dS 2 . Applying the induction assumption for N, from (1-4), we have pN (λ) ≤ CN dN+2 S for λ ∈ ¯B(0, dS 4 ). This implies N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xj=0 pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(0, dS 2 )) < 2CN dS, k¯zN+1 + pN (z)¯zNkL1(K∩ ¯B(0, From (2-12), we conclude dS 4 )) ≤ 3CN dN+3 S . dS 4 )) In general, there is an absolute constant CN+1 > 0 so that N−1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xj=0 pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(0, Xj=0 pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(0, N−k (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dS 2k+1 )) ≤ 4CN dN+3 S . ≤ CN+1dN+3 S . Since S is a heavy ǫ square, we get γ(B(0, dS 2k+1 ) \ K) ≤ γ(S \ K) < ǫdS ≤ ǫN−k dS 2k+1 . Apply the induction assumption for N − k, from (1-4), we get pN−k(λ) ≤ CN−k dN−k+2 S ≤ CN−kCN+1dk+1 S , dS 2k+1 )) for λ ∈ ¯B(0, dS 2k+2 ). So there is an absolute constant which we still use CN+1 > 0 such that N−k Xj=0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(0, pj(z)¯zj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0 N 9 ¯zN+1 + (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CN+1dN+1 S for z ∈ ¯B(0, dS Let φN+1(z) = φ( 2N +1z 2N +1 ). Let gN+1 = ¯zN+1 +PN j=0 pj ¯zj, then gN+1(z) ≤ CN+1dN+1 2N +1 ), ), where φ is in (2-11), then spt(φN+1) ⊂ B(0, dS , Z φN+1dA = φN +1 N+1 /dN+1 0 ≤ φN+1 ≤ 1, k ¯∂N+1φN+1k < C dS S S d2 S 4N+1 . on z ∈ ¯B(0, dS 2N +1 ). Then (N + 1)! d2 S 4N+1 ≤ =(N + 1)!Z φN+1dA Z gN+1 ¯∂N+1φN+1dA(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)kgN+1kL1(K∩ ¯B(0, 2N +1 )\K)(cid:19) dS 2N+1 ) \ K) kgN+1kL1(K∩S) + CN+1C dist(cid:16)¯zN+1, P 1(1, ¯z, ..., ¯zN , K ∩ S)(cid:17) + 4πCN+1C φ C φN +1 N+1 dN+1 S φN +1 C N+1 dN+1 S + kgN+1kL1( ¯B(0, φN +1 N+1 Area(B(0, φN +1 C N+1 dN+1 S 2N +1 )) ≤2 ≤ dS dS N+1γ(Int(S) \ K)2 where the last step follows from (2-12) and Theorem 3.2 on page 204 of Gamelin (1969). Now choose then since S is a heavy ǫ square, we have c2 1 = (N + 1)! 22N+5πCN+1C φN +1 N+1 , dist(cid:16)¯zN+1, P 1(1, ¯z, ..., ¯zN , K ∩ S)(cid:17) ≥ (N + 1)! 22N+4C φN +1 N+1 dN+3 S . So (2-6) holds and the theorem follows from Lemma 3 (1). For a smooth function ϕ with compact support, Tϕ is a bounded linear operator on C(K). Let M be the space of finite complex Borel measures supported on K ( = C(K)∗), then T ∗ϕ is a bounded linear operator on M. Moreover, for µ ∈ M, Z Tϕf dµ =Z f dT ∗ϕµ and kT ∗ϕµk ≤ 4diameter(suppϕ)k ¯∂ϕkkµk. (2-13) Consequently, T ∗ϕµ ⊥ R(K) for each µ ⊥ R(K). Lemma 4. Let S be a square with center 0 and dS < 1. Let gN = ¯zN +PN−1 polynomials and kgNk ≤ CN dN S (CN is an absolute constant depending on N ) on S. If k=1 pk ¯zk, where pk are where and C φ N is in (2-11). Then dist(gN , R(K ∩ S)) ≤ cN dN S , cN = N ! 2N+8πC φ N γ(Int(S) \ K) ≥ 4cN CN + 4cN dS. Proof: Let φ1 = ¯∂N−1φ( 2z dS ), where φ is in (2-11). Let f1 = Tφ1 gN , then and f′1(∞) = N CN dS kf1k ≤ 2N+3C φ π Z ¯∂gN φ1dA = (−1)N−1 N !d2 4π 1 S . 10 We have the following computation: dist(f1, R(C∞ \ (Int(S) \ K))) = = sup µ⊥R(C∞\(Int(S)\K)) kµk=1 sup µ⊥R(C∞\(Int(S)\K)) kµk=1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Z f1dµ(cid:12)(cid:12)(cid:12)(cid:12) Z gN dT ∗φ1 µ(cid:12)(cid:12)(cid:12)(cid:12) ≤ sup kµk=1 µ⊥R(C∞\(Int(S)\K)) ≤8dSk ¯∂N φ1kdist(gN , R(K ∩ S)) ≤2N+3C φ N cN dS, kT ∗φ1 µkdist(gN , R(K ∩ S)) where we use the fact that T ∗φ1 µ ⊥ R(K ∩ S) for µ ⊥ R(C∞ \ (Int(S) \ K)). Let f = f1 (−1)N−12N+3C φ N CN dS , then f is analytic off S, kfk ≤ 1, f′(∞) = 8 cN CN dS, and dist(f, R(C∞ \ (Int(S) \ K))) ≤ cN CN . Then there is a compact subset F of Int(S) \ K and a rational function r with poles in F such that Hence, r(∞) < 2 cN CN , krkC∞\F ≤ 1 + 2 cN CN kf − rkC∞\F < 2 , and by the maximum modulus principle, . cN CN f′(∞) − r′(∞) = lim z→∞ z(f (z) − r(z) + r(∞)) ≤ max z=dS z(f (z) − r(z) + r(∞)) < 4 cN CN dS. Therefore, γ(Int(S) \ K) ≥ γ(F ) ≥ r′(∞) krkC∞\F + r(∞) ≥ 4cN CN + 4cN dS. Proof of Theorem 1 (2): We assume λ0 = 0 and δ = 1 (same argument used in the first paragraph of the proof of Theorem 1 (1)). We use mathematical induction for N. First we consider the case that N = 1. Let ǫ = 4c1 C1 + 4c1 in Lemma 4. Let ǫ1 be as in (2-5). Then from assumption (1-3) and Lemma 2, we conclude that Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true. Let S be a heavy ǫ square and g1 = ¯z − cS, then by Lemma 4, we must have Hence, (2-9) holds for N = 1, by Lemma 3, we prove (1-5) for N = 1. dist(¯z, R(S ∩ K)) = dist(g1, R(S ∩ K)) ≥ c1dS. Now we assume that (1-5) holds for k = 1, 2, ..., N. The proof is similar to that of Theorem 1 (1). Set ǫ = min(c0, ǫN 2 , ǫN−1 22 , ..., ǫ1 2N , ǫ0 2N+1 ), where c0 > 0 will be determined later, and ǫN+1 as in (2-5). Since γ(B(0, 1) \ K) < ǫN+1 (assumption (1-3)), from Lemma 2, we conclude that Case II of scheme(Q, ǫ, m, γn, Γn, n ≥ m) is true. Let S be a heavy ǫ square. We assume dist(cid:16)¯zN+1, HP (1, ¯z, ..., ¯zN , K ∩ S)(cid:17) ≤ 1 2 dN+1 S , otherwise (2-9) already holds. Without loss of generality, we assume the center of S is zero. There are polynomials p1, ..., pN and a rational function r with poles off K ∩ S such that ¯zN+1 + (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N Xj=1 pj ¯zj + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩S 11 ≤ dN+1 S . Using the same argument of the paragraph (in the proof of Theorem 1) under (2-12), we get (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¯zN+1 + N Xj=1 pj ¯zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩ ¯B(0, 2N +1 ) j=1 pj ¯zj, then kgN+1k ≤ CN+1dN+1 S dS ≤ CN+1dN+1 S . on 1 2N +1 S ⊂ B(0, dS 2N +1 ). By Lemma 4 for Let gN+1 = ¯zN+1 +PN 2N +1 S, if we choose 1 c0 = 4cN+1 CN+1 + 4cN+1 , we must have dist(cid:16)¯zN+1, HP (1, ¯z, ..., ¯zN , K ∩ S)(cid:17) = dist(cid:16)gN+1, HP (1, ¯z, ..., ¯zN , K ∩ S)(cid:17) ≥ cN+1dN+1 S Therefore, (2-9) holds and the theorem now follows from Lemma 3. . Proof of Theorem 2: Assume that λ0 is not a peak point for R(K), then by Curtis's Criterion (see Theorem 4.1 in Gamelin (1969), p. 204), we get, Then together with the assumptions for F1, ..., FN , we can choose δ > 0 such that M γ(B(λ0, δ) \ K) δ = 0. lim sup δ→0 and Fj (z) = Xi=0 γ(B(λ0, δ) \ K) < ǫδ where M ≥ N, gij ∈ A( ¯B(λ0, δ)), j = 1, ..., N, i = 0, 1, ..., M, , ..., ǫM−1 gij (z)¯zi, z ∈ B(λ0, δ), ǫ = min(cid:16)ǫM , 2 ǫ1 2M−1(cid:17) , and ǫ1, ..., ǫM are in Theorem 1. Set where g00 = 1 and gi0 = 0 for i = 1, ..., M. Then F0(z) = gij (z)¯zi, z ∈ B(λ0, δ), M Xi=0 (2-14) g00 g10 ... gM 0 ... ... ... ... g0N g1N ... gM N     F0 0 ... 0   F1 ¯∂F1 ... ¯∂N F1 ... ... ... ... FN ¯∂FN ... ¯∂N FN   ¯z 1 0 ... 0 ¯z2 2¯z 2 ... 0 ¯zM M ¯zM−1 ... ... ... M (M − 1)¯zM−2 ... (M−N)! ¯zM−N ... ... M !   Since the matrix [ ¯∂iFj(λ0)]1≤i,j≤N is invertible, we may choose above δ small enough and i0 = 1 < i1 < i2 < ... < iN so that 1 0 0 ... 0 =   G(λ) =  gi00 gi10 ... giN 0 is analytic and invertible on B(λ0, δ). Moreover, gi01 gi11 ... giN 1 ... ... ... ... gi0N gi1N ... giN N   kG(λ)−1k ≤ C, λ ∈ B(λ0, δ). So G(λ) and G(λ)−1 are linear and uniform bounded operators from (CN+1, k.kN+1) to (CN+1,k.kN+1), where kxkN+1 =PN+1 i=1 xi. For polynomials p0, p1, ..., pN , we get N N N Xj=0 M Xi=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) pj(λ)gij(λ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) pj(λ)gikj (λ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) pj (λ) Xj=0 ≥ ≥ Xk=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0 1 C N 12 =k(p0(λ), p1(λ), ..., pN (λ))G(λ)k (2-15) on B(λ0, δ following calculation 2 ) and q0, q1, ..., qM are polynomials. Using (2-14) and Theorem 1 (1) again, we have the + kqMkK∩ ¯B(λ0, δ 2 )k¯zMkL1(K∩ ¯B(λ0, δ 2 ))  qi ¯zi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) on B(λ0, δ 4 ). Therefore, there is a constant CN > 0 such that δ 2M +1 ). From (2-15) and (2-16), we see that on B(λ0, on B(λ0, N Xj=0 pj (λ) ≤ C N pjFj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) δ 2M +1 ). So λ0 is an analytic bounded point evaluation for P t(1, F1, ..., FN , K ∩ ¯B(λ0, δ)). The proof of (2) is the same. (2-16) becomes (2-16) (2-17) on B(λ0, δ). Now let us prove (1). From (2-14) and Theorem 1 (1), we have qM−1(λ) ≤ ≤ ≤ CM−1 ( δ CM−1 ( δ CM−1 ( δ M M M M 2 )) 2 )) CM CM M−1 Xi=0 qM (λ) ≤ qi ¯zi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) δM +2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 )M +1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) qi ¯zi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0, δ Xi=0 2 )M +1  (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) qi ¯zi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0 , δ Xi=0  δM +2 k¯zMkL1(K∩ ¯B(λ0,δ))(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 )M +1 (cid:18)1 + Xi=0 qi ¯zi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(K∩ ¯B(λ0,δ)) qi(λ) ≤ CN(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=0 Xi=0 Xi=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CCN(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pj(λ)gij(λ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=0 Xj=0 qi(λ) ≤ CN(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) qi ¯zi + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩ ¯B(λ0,δ) Xi=1 Xi=1 Xi=1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CCN(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pj(λ)gij(λ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj=1 Xj=1 M M M M M N N N N Xj=1 pj (λ) ≤ C where r is a rational function with poles off K ∩ ¯B(λ0, δ). (2-17) becomes pjFj + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩ ¯B(λ0,δ) on B(λ0, δ 2M +1 ). So λ0 is an analytic bounded point evaluation for HP (F1, ..., FN , K ∩ ¯B(λ0, δ)). 3 Uniform Rational Approximation In this section, we will prove Theorem 3 and Proposition 1. To prove Theorem 3, we need to prove several Lemmas. Lemma 5. If R(K) 6= C(K), then HP (¯z, ¯z2, ..., ¯zN , K) 6= C(K). Proof: Notice, by Lemma 3 (2-10), N CN δN (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) , λ ∈ ¯B(cid:18)λ0, Then there exists a finite Borel measure µ supported on K ∩ ¯B(λ0, δ) \ B(λ0, δ pN (λ) ≤ Xk=1 δ(cid:19) . pk ¯zk + r(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K∩ ¯B(λ0,δ)\B(λ0, δ pN (λ0) =Z ( 2 )) and pk ∈ P. Therefore, the non zero measure (z − λ0)µ ⊥ 2 ) such that pk ¯zk + r)dµ, Xk=1 where r ∈ Rat(K ∩ ¯B(λ0, δ) \ B(λ0, δ HP (¯z, ¯z2, ..., ¯zN , K) and HP (¯z, ¯z2, ..., ¯zN , K) 6= C(K). 1 2 2 ) N 13 Lemma 6. Let T be a closed square and γ(B(cT ,√2dT )\K) < ǫN√2dT . If A(K) ⊂ HP (¯z, ¯z2, ..., ¯zN , K), φ is a smooth function supported inside T, and f ∈ A(K), then Tφf ∈ R(K). Proof: Case 1: Suppose Int(K ∩ T ) = ∅. Then A(K ∩ T ) = C(K ∩ T ) ⊂ HP (¯z, ¯z2, ..., ¯zN , K ∩ T ) since each smooth function with support in K ∩ T belongs to A(K). From Lemma 5, we get C(K ∩ T ) = R(K ∩ T ). Hence, Tφf ∈ R(K). Case 2: Int(K ∩ T ) 6= ∅. There are sequences of {pij}1≤i≤N,1≤j<∞ ⊂ P and {rj} ⊂ Rat(K) such that ( lim j→∞ N Xi=1 pij ¯zi + rj ) = Tφf uniformly on K since TφA(K) ⊂ A(K). By Theorem 1 (2), for each i , the sequence {pij}1≤j<∞ converges to an analytic function fi uniformly on T ⊂ ¯B(cT ,√2dT /2). Hence {rj} converges to r uniformly on K ∩ T that is analytic on Int(K ∩ T ) and Tφf (z) = PN i=1 fi(z)¯zi + r(z) on Int(K ∩ T ). This implies fi(z) = 0 on T and Tφf ∈ R(K ∩ T ). The Lemma is proved. Lemma 7. If A(K) ⊂ HP (¯z, ¯z2, ..., ¯zN , K), then A(K) = R(K). Proof: We use standard Vitushkin approximation scheme (see Gamelin (1969) for example). Let {ψn, Sn} be a smooth partition of unity, where the length of Sn is δ, the support of ψn is in 2Sn, k ¯∂ψnk ≤ C/δ, P ψn = 1, and ∪∞n=1Sn = C with Int(Sn) ∩ Int(Sm) = ∅. For a function f ∈ A(K), f = Tψn f. ∞ Xn=1 For a fixed n, let T = 2Sn, if γ(B(cT ,√2dT ) \ K) < ǫN√2dT , then, by Lemma 6, hn = Tψn f ∈ R(K). If γ(B(cT ,√2dT ) \ K) ≥ ǫN√2dT , then γ(Int(2T ) \ K) ≥ ǫN√2dT . Since Z (z − cT )f ¯∂ψndA(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Z f ¯∂ψndA(cid:12)(cid:12)(cid:12)(cid:12) ≤ C1dT ω(f, δ), ≤ C1d2 T ω(f, δ), where (cid:12)(cid:12)(cid:12)(cid:12) ω(f, δ) = sup z,w∈B(cT ,√2δ) f (z) − f (w), , β = R (z − cT )f ¯∂ψndA T ω(f, δ) C1d2 . α = R f ¯∂ψndA C1dT ω(f, δ) we set then Using Lemma 1, we can find a function g ∈ R(C∞\(Int(2T )\K)) ⊂ R(K) satisfying (1) to (5) of Lemma 1. Now let hn = C1ω(f,δ) g, then hn ∈ R(K), hn is analytic off 2T, khnk ≤ Cω(f, δ), and hn − Tψn f has triple zeros at ∞. So P∞n=1 hn goes to f uniformly when δ tends to zero. This completes the proof of Proof of Theorem 3: Let M be the largest power of ¯z among all terms of q1(z, ¯z), q2(z, ¯z), ..., qN (z, ¯z), the lemma. π A(K) ⊂ HP (q1, ..., qN , K) ⊂ HP ((¯z, ¯z2, ..., ¯zM ), K). By Lemma 7, we conclude A(K) = R(K). Can Theorem 3 work for more general functions? For N = 1 and a smooth function g with ¯∂g 6= 0, it is proved in Yang (1994) that A(K) ⊂ HP (g, K) implies A(K) = R(K). We think this may still hold for N > 1. However, in this section, we provide an example for a single function as stated in Proposition 1. Let us construct a compact subset K0 of the closed unit square with center at zero and sides parallel to coordinate axes such that P (K0) = C(K0) and Area(K0) = a, 0 < a < 1. In fact, we can construct a planar Cantor set K0 as the following. Given a sequence {λn} with 0 < λn < 1 2 , let Q0 = [0, 1] × [0, 1]. At the first step we take four closed squares inside Q0, with side length λ1, with sides parallel to the coordinate axes, and so that each square contains a vertex of Q0. At the second step we apply the preceding procedure to each of the four squares obtained in the first step, but now using the proportion factor λ2. In this way, we get 16 squares of side length σ2 = λ1λ2. Proceeding inductively, at each step we obtain 4n squares Qn j , j = 1, 2, ...4n with side length σn = λ1λ2...λn. Now let Ln = ∪4n j=1Qn j , K0 = ∩∞n=1Ln, and λn = 1 2n+1 . a 1 2 14 Then Area(K0) = lim n→∞ 4nσ2 n = a, By construction, C \ K0 is connected, so P (K0) = C(K0). (3-1) σn σn+1 = 2, we can choose n0 so that for n ≥ n0, we have 2σn+1 ≤ σn ≤ 2.1σn+1. Now let T Since limn→∞ be a square with dT ≤ σn0 , then we can find an integer n1 > n0 such that 2σn1+1 ≤ σn1 ≤ dT ≤ σn1−1. Suppose T ∩ K0 6= ∅, there exists Qn1+1 Area((2T ) ∩ K0) ≥ Area(Qn1+1 6= ∅. Therefore, Qn1+1 n1+1+n ≥ aσ2 with T ∩ Qn1+1 4nσ2 ∩ K0) = lim n→∞ (2(2.1)2)2 Area((2T ). ⊂ 2T and a n1+1 ≥ j1 j1 j1 j1 In this case, γ((2T ) ∩ K0)) ≥ a (4(2.1)2)√π (= c0)dT . (3-2) Now we will construct a sequence of disjoint small open disks {Bk(zk, rk)}∞k=1 within G = B(0, 1) \ K0 satisfying the following conditions: (1) Each point in K0 is a limit of a subsequence of the disks; (2) No point in G is a limit of a subsequence of the disks; (3) P∞k=1 rk < ∞. Let {xk} ⊂ K0 be a subset that is dense in K0. We begin with a point y11 ∈ G, choose 0 < r11 < 1 with B(y11, r11) ⊂ G. Let d11 = dist(y11, K0) − r11. We finish the level 1 construction. For level 2, choose y21 ∈ G so that dist(y21, x1) < d11 2 and 0 < r21 < min( d11 22 ) with B(y21, r21) ⊂ G. Define d21 = dist(y21, K0) − r21. Choose y22 ∈ G so that dist(y22, x2) < d21 22 ) with B(y22, r22) ⊂ G. Define d22 = dist(y22, K0) − r22. We continue this process and get {yij}, {rij}, and {dij} satisfying: 4 , 1 2 and 0 < r22 < min( d21 4 , 1 2 and dist(yij, K0) = dij + rij , X rij < ∞; di,1 < , dij < di−1,i−1 2 di,j−1 , 2 where 1 < j ≤ i. Let {zk} = {yij} and {rk} = {rij}. Clearly the conditions (1)-(3) are met. Set K = ¯B(0, 1) \ ∞ [k=1 B(zk, rk)! . (3.3) It is easy to verify that the inner boundary ∂iK equals K0. For this K, we can prove Proposition 1. Proof of Proposition 1: Let φ = C(dAK0 ), then φ ∈ A(K). Define µ = dz∂B(0,1) − ∞ Xk=1 dz∂B(zk ,rk), then µ is a finite Borel measure, µ ⊥ R(K), and Z φdµ =ZK0Z 1 λ − z dµ(z)dA(λ) = −2πiArea(K0) 6= 0. Hence, R(K) 6= A(K). Since for a polynomial p, C(pdAK0 ) − pφ =ZK0 p(w) − p(z) w − z dA(w) ∈ R(K), we see that C(pdAK0 ) ∈ HP (φ, K). Using (3-1), we conclude that C(χEdA) ∈ HP (φ, K), E ⊂ K0, (3-3) where χE is the characteristic function of E. Let T and Qn1+1 squares j1 i=1Qn1+2 ∪4 ki ⊂ Qn1+1 j1 . be squares as above. There are four 15 Choose two of them, say Qn1+2 Qn1+2 Then from (3-3), we have fj ∈ HP (φ, K) and ∩ K0 and E2 = Qn1+2 and Qn1+2 k2 k1 k1 , with the same y coordinates of the centers. Set E1 = ∩ K0. Let φ1(z) = −χE1 and φ2(z) = −χE2 . Set fj = C(φjdA) for j = 1, 2. k2 since Area(E1) = Area(E2), and f′j (∞) = Area(Ej)(= AE) β(fj , cT ) =Z (z − cEj )χE(z)dA + (cEj − cT )Area(Ej) = (cEj − cT )AE. g1 = (cE2 − cT )f1 − (cE1 − cT )f2 (cE2 − cE1 )AE dT , We Set and g2 = d2 T . f2 − f1 (cE2 − cE1 )AE Notice that cE2 −cT < 2dT , cE1 −cT < 2dT , and cE2 −cE1 is comparable with dT . Using the arguments before (3-2), we see that kgjk ≤ C ( absolute constant), gj ∈ HP (φ, K), and g′1(∞) = dT , g′2(∞) = 0, β(g1, cT ) = 0, β(g2, cT ) = d2 T . Now we use standard Vitushkin approximation scheme (see Gamelin (1969) for example). Let {ψn, Sn} be a smooth partition of unity, where the length of Sn is δ, the support of ψn is in 2Sn, k ¯∂ψnk ≤ C/δ, σn0 P ψn = 1, and ∪∞n=1Sn = C with Int(Sn) ∩ Int(Sm) = ∅. We assume δ is less than 2 . For a function f ∈ A(K), f = Tψn f. ∞ Xn=1 For a fixed n, if (2Sn) ∩ K0 = ∅, then hn = Tψnf ∈ R(K). If (2Sn) ∩ K0 6= ∅, then let T = 2Sn and hn = R f ¯∂ψndA πdT g1 + R (z − cT )f ¯∂ψndA πd2 T g2, then hn ∈ HP (φ, K), hn is analytic off 2T, khnk ≤ Cω(f, δ), and hn − Tψn f has triple zeros at ∞. So P∞n=1 hn goes to f uniformly when δ tends to zero. This completes the proof of the proposition. References A. Aleman, S. Richter, and C. Sundberg. Nontangential limits in P t(µ)-spaces and the index of invariant subspaces. Ann. of Math., 169(2):449 -- 490, 2009. A. Aleman, S. Richter, and C. Sundberg. A quantitative estimate for bounded point evaluations in P t(µ)-spaces. Topics in operator theory. Operators, matrices and analytic functions, Oper. Theory Adv. Appl., 202, Birkhuser Verlag, Basel, 1:1 -- 10, 2010. A.D. Baranova, J.J. Carmonab, and K.Yu. Fedorovskiyc. Density of certain polynomial modules. Journal of Approximation Theory, 206:1 -- 16, 2016. J. E. Brennan. Thomson's theorem on mean-square polynomial approximation, algebra i analiz 17 no.2 (2005), 1-32. Russian. St. Petersburg Math. J., 17(2):217 -- 238, 2006. English. J. E. Brennan. The structure of certain spaces of analytic functions. Comput. Methods Funct. theory, 8 (2):625 -- 640, 2008. J. E. Brennan and E. R. Militzer. Lp-bounded point evaluations for polynomials and uniform rational approximation. St. Petersburg Math. J., 22(1):41 -- 53, 2011. J. B. Conway. The theory of subnormal operators. Mathematical Survey and Monographs 36, 1991. J. B. Conway and N. Elias. Analytic bounded point evaluations for spaces of rational functions. J. Functional Analysis, 117:1 -- 24, 1993. 16 J. Deny. Systemes totauc de fonctions harmoniques. Ann. Inst. Fourier, pages 103 -- 113, 1949. T. W. Gamelin. Uniform algebras. American Mathematical Society, Rhode Island, 1969. M. V. Keldysh. On the solubility and stability of the dirichlet problem. Amer. Math. Soc. Efransl, 51: 1?73, 1966. M.Ya. Mazalov. Uniform approximations by bianalytic functions on arbitrary compact subsets of . Sb. Math., 195(5):687?709, 2004. J. E. Thomson. Approximation in the mean by polynomials. Ann. of Math., 133(3):477 -- 507, 1991. J. E. Thomson. Uniform approximation by rational functions. Indiana Univ. Math. J, 42:167 -- 177, 1993. X. Tolsa. Painleves problem and the semiadditivity of analytic capacity. Acta Math., 190(1):105 -- 149, 2003. X. Tolsa. The semiadditivity of continuous analytic capacity and the inner boundary conjecture. Amer. J. Math., 126:523 -- 567, 2004. T. Trent and J. L. Wang. Uniform approimation by rational modules on nowhere dense sets. Proc. Amer. Math. Soc., 81:62?64, 1981. J. Verdera. On the uniform approximation problem for the square of the cauchy-riemann operator. Pacific J. Math., 159(2):379 -- 392, 1993. A. G. Vitushkin. Analytic capacity of sets in problem of approximation theory. Russian Math. Surveys, 22:139?200, 1967. A. G. Vitushkin and M. S. Melnikov. Analytic capacity and rational approximation. Linear and complex analysis, Problem book, Lecture Notes in Math. 1403, Springer-Verlag, Berlin, 1984. L. Yang. On uniform approximation problems and t -invariant algebras. Indiana Univ. Math. J, 43: 639 -- 650, 1994. L. Yang. Uniform rational approximation. Proc. Amer. Math. Soc., 123(1):201 -- 206, 1995. L. Yang. A note oo Lp-bounded point evaluations for polynomials. Proc. Amer. Math. Soc., 144(11): 4943 -- 4948, 2016. L. Yang. Bounded point evaluations for rationally multicyclic subnormal operators. Journal of Mathe- matical Analysis and Applications, 2018. 17
1312.2304
1
1312
2013-12-09T04:10:13
Isomorphisms of $AC(\sigma)$ spaces
[ "math.FA", "math.MG" ]
Analogues of the classical Banach-Stone theorem for spaces of continuous functions are studied in the context of the spaces of absolutely continuous functions introduced by Ashton and Doust. We show that if $AC(\sigma_1)$ is algebra isomorphic to $AC(\sigma_2)$ then $\sigma_1$ is homeomorphic to $\sigma_2$. The converse however is false. In a positive direction we show that the converse implication does hold if the sets $\sigma_1$ and $\sigma_2$ are confined to a restricted collection of compact sets, such as the set of all simple polygons.
math.FA
math
ISOMORPHISMS OF AC(σ) SPACES IAN DOUST AND MICHAEL LEINERT Abstract. Analogues of the classical Banach-Stone theorem for spaces of con- tinuous functions are studied in the context of the spaces of absolutely contin- uous functions introduced by Ashton and Doust. We show that if AC(σ1) is algebra isomorphic to AC(σ2) then σ1 is homeomorphic to σ2. The converse however is false. In a positive direction we show that the converse implication does hold if the sets σ1 and σ2 are confined to a restricted collection of compact sets, such as the set of all simple polygons. 1. Introduction In [3] Ashton and Doust defined the Banach algebra AC(σ) consisting of 'abso- lutely continuous' functions with domain an arbitrary nonempty compact subset σ of C (or equivalently of R2). The motivation for their definition was to extend the spectral theory of well-bounded operators to cover operators whose spectra need not be contained in the real line. This led to the definition of an AC(σ) op- erator being a bounded operator on a Banach space X which admits a bounded functional calculus Ψ : AC(σ) → B(X). Under some additional assumptions, the image of this map Ψ is an algebra of operators which is isomorphic to AC(σ). Ac- cordingly, one can recover certain aspects of the theories of normal operators and of scalar-type spectral operators, replacing algebras of continuous functions C(Ω) with algebras of absolutely continuous functions. Quite naturally then, underly- ing many of the open problems in this area are questions which ask for analogues of the classical topological results about C(Ω) spaces. (Details of the theory of AC(σ) operators can be found in [5].) One of the most classical of these topological results is the Banach-Stone theo- rem which says that two compact Hausdorff spaces Ω1 and Ω2 are homeomorphic if and only if the function algebras C(Ω1) and C(Ω2) are linearly isometric. There have been many generalizations and extensions of this result (see [11]). Work of Amir [1] shows that one may still deduce that Ω1 and Ω2 are homeomorphic if one only assumes that C(Ω1) and C(Ω2) are (linearly) isomorphic with Banach-Mazur distance less than 2. Cohen [7] has shown that the value 2 is sharp. In a different direction, one might require that the spaces C(Ω1) and C(Ω2) be isomorphic as algebras. In this case one may argue using the maximal ideal spaces Date: 28 November 2013. 2010 Mathematics Subject Classification. Primary: 46J10. Secondary: 46J35,47B40,26B30. 1 2 IAN DOUST AND MICHAEL LEINERT to get the same conclusion. In particular, as the next result shows, if C(Ω1) and C(Ω2) are algebra isomorphic, then they are isometrically isomorphic. This result was originally proved in [12]; a modern treatment is given in [11]. Theorem 1.1 (Gelfand and Kolmogoroff 1939). Let Ω1 and Ω2 be compact Haus- dorff spaces. Then C(Ω1) and C(Ω2) are isomorphic as algebras if and only if Ω1 and Ω2 are homeomorphic. Moreover, every algebra isomorphism j : C(Ω1) → C(Ω2) is of the form j(f ) = f ◦ h where h : Ω1 → Ω2 is a homeomorphism. The main issue that we shall address in this paper is the corresponding relation- ship between the topological structure of the set σ and the algebraic structure of AC(σ). In Section 2 we shall recall the definition and main properties of AC(σ) and then give a simple proof that if AC(σ1) and AC(σ2) are algebra isomorphic, then σ1 and σ2 are homeomorphic. The converse of this is false however. In Sec- tion 3 we show that the algebra of absolutely continuous functions over the closed unit disk is not isomorphic to the algebra of absolutely continuous functions over a square. If, however, one restricts the class of sets in which σ may lie, one can recover some sort of analogue of the Banach-Stone Theorem. In Theorem 6.3 we show that if P1 and P2 are simple polygons, then AC(P1) is algebra isomorphic to AC(P2). In Section 7 we extend this result to cover more general sets based on polygons. We note that different applications have led to quite a number of different concepts of absolute continuity for functions of two or more variables. The reader is cautioned that these concepts are generally distinct, and often, as is the case here, impose particular conditions on the domains of the functions considered. The definition of absolute continuity that is studied here was developed to have specific properties which are appropriate for the intended application to spectral theory, namely: (1) it should apply to functions defined on the spectrum of a bounded operator, that is, an arbitrary nonempty compact subset σ of the plane, (2) it should agree with the usual definition if σ is an interval in R; (3) AC(σ) should contain all sufficiently well-behaved functions; (4) if α, β ∈ C with α 6= 0, then the space AC(ασ + β) should be isometrically isomorphic to AC(σ). The interested reader may consult [4], [9] and [6] for a sample of what is known about the relationships between some of these definitions. Notation. Suppose that A and B are Banach algebras. We shall write A ≃ B to mean that A and B are isomorphic (as Banach algebras). Throughout, we shall use the term polygon to refer to a simple polygon including its interior. In particular, every such polygon is homeomorphic to the closed unit disk. ISOMORPHISMS OF AC(σ) SPACES 3 2. Preliminaries In this section we shall briefly outline the definition of the spaces AC(σ). Here we follow [10] rather than the original definitions given in [3]. Throughout, σ, σ1 and σ2 will denote nonempty compact subsets of the plane. Although the original motivation for these definitions came from considering functions defined on subsets of the complex plane, for this paper it will be notationally easier consider the domains of the functions to be subsets of R2. We shall work throughout with algebras of complex-valued functions. Suppose that f : σ → C. Let S = (cid:2)x0, x1, . . . , xn(cid:3) be a finite ordered list of elements of σ, where, for the moment, we shall assume that n ≥ 1. Note that the elements of such a list do not need to be distinct. We define the curve variation of f on the set S to be (2.1) cvar(f, S) = n Xi=1 f (xi) − f (xi−1) . We shall also need to measure the 'variation factor' of the list S. Loosely speaking, this is the greatest number of times that γS crosses any line in the plane, where γS denotes the piecewise linear curve joining the points of S in order. The following definition makes precise just what is meant by a crossing. Definition 2.1. Suppose that ℓ is a line in the plane. We say that xi xi+1, the line segment joining xi to xi+1, is a crossing segment of S = (cid:2)x0, x1, . . . , xn(cid:3) on ℓ if any one of the following holds: (i) xi and xi+1 lie on (strictly) opposite sides of ℓ. (ii) i = 0 and xi ∈ ℓ. (iii) i > 0, xi ∈ ℓ and xi−1 6∈ ℓ. (iv) i = n − 1, xi 6∈ ℓ and xi+1 ∈ ℓ. In this case we shall write xi xi+1 ∈ X(S, ℓ). Definition 2.2. Let vf(S, ℓ) denote the number of crossing segments of S on ℓ. The variation factor of S is defined to be vf(S) = max ℓ vf(S, ℓ). Clearly 1 ≤ vf(S) ≤ n. For completeness, in the case that S = (cid:2)x0(cid:3) we set cvar(f,(cid:2)x0(cid:3)) = 0 and let vf((cid:2)x0(cid:3), ℓ) = 1 whenever x0 ∈ ℓ. Example 2.3. Consider the line ℓ and the list S = [xi]8 i=0 as shown in Figure 1. Let si = xi xi+1. Then the crossing segments for S on ℓ are s0 (rule (ii)), s2 (rule (i)), s4 (rule (iii)) and s7 (rule (iv)). All the other segments are not crossing segments of S on ℓ. Thus vf(S, ℓ) = 4. 4 IAN DOUST AND MICHAEL LEINERT x2 x7 x0 x1 x4 x5 x6 x8 ℓ x3 Figure 1. Examples of crossing segments. The two-dimensional variation of a function f : σ → C is defined to be (2.2) var(f, σ) = sup S cvar(f, S) vf(S) , where the supremum is taken over all finite ordered lists of elements of σ. The variation norm is and the set of functions of bounded variation on σ is kf kBV (σ) = kf k∞ + var(f, σ) BV (σ) = {f : σ → C : kf kBV (σ) < ∞}. The space BV (σ) is a Banach algebra under pointwise operations [3, Theorem 3.8]. If σ = [0, 1] then the above definition is equivalent to the more classical one. Let P2 denote the space of polynomials in two real variables of the form p(x, y) = Pn,m cnmxnym, and let P2(σ) denote the restrictions of elements on P2 to σ. The algebra P2(σ) is always a subalgebra of BV (σ) [3, Corollary 3.14]. Definition 2.4. The set of absolutely continuous functions on σ, denoted AC(σ), is the closure of P2(σ) in BV (σ). The set AC(σ) forms a closed subalgebra of BV (σ) and hence is a Banach algebra. We shall say that f ∈ C 1(σ) if there exists an open neighbourhood U of σ and an extension F of f to U such that the partial derivatives of F (of order one) are continuous on U. The space CTPP (σ) consists of those functions f for which there is a triangulation of a neighbourhood U of σ and an extension of f to U which is continuous and piecewise planar on this triangulation. It was shown in [10] that both C 1(σ) and CTPP (σ) are dense subsets of AC(σ). Our first step is to show that if AC(σ1) and AC(σ2) are isomorphic as algebras, then σ1 and σ2 must be homeomorphic. We note that one does not need to assume that the isomorphism is continuous. Lemma 2.5. Suppose that f ∈ AC(σ). Then the spectrum of f is σ(f ) = f (σ) and hence the spectral radius of f is r(f ) = kf k∞. Proof. This is more or less immediate from [3, Corollary 3.9]. (cid:3) ISOMORPHISMS OF AC(σ) SPACES 5 Theorem 2.6. Suppose that j : AC(σ1) → AC(σ2) is an algebra isomorphism. Then (1) kf k∞ = kj(f )k∞ for all f ∈ AC(σ1). (2) there exists a homeomorphism h : σ1 → σ2. (3) j(f ) = f ◦ h−1 for all f ∈ AC(σ1). (4) j is continuous. Proof. Since j preserves the identity element, it also preserves the spectrum of elements. Thus, using Lemma 2.5, kf k∞ = r(f ) = r(j(f )) = kj(f )k∞ for all f ∈ AC(σ). Since AC(σ1) is dense in C(σ1), this implies that j extends to an isometric isomorphism  : C(σ1) → C(σ2) and hence, by the Banach-Stone Theorem, σ1 is homeomorphic to σ2. Indeed there exists a homeomorphism h : σ1 → σ2 such that (f ) = f ◦ h−1 for all f ∈ C(σ1). Restricting this to AC(σ1) gives part 3. Suppose that fn → f in AC(σ1). Then certainly fn → f uniformly and hence pointwise. Suppose that j(fn) → g in AC(σ2) (and hence also pointwise). Then for all x ∈ σ2, g(x) = lim n j(fn)(x) = lim n fn(h−1(x)) = f (h−1(x)) = j(f )(x) and hence j(f ) = g. Thus, by the Closed Graph Theorem, j is continuous. (cid:3) It is easy to find homeomorphic sets σ1 and σ2 for which AC(σ1) and AC(σ2) are algebra isomorphic, but not isometrically. If the isomorphism preserves norms, then part 1 of the above theorem implies that it also preserves variation. Corollary 2.7. Suppose that j : AC(σ1) → AC(σ2) is an isometric Banach algebra isomorphism. Then var(f, σ1) = var(j(f ), σ2) for all f ∈ AC(σ1). Example 2.8. Let σ1 = {0, 1, 2} and σ2 = {0, 1, i}. Since σ1 ⊆ R, kf kBV (σ1) = kf k∞ + f (1) − f (0) + f (2) − f (1), f ∈ AC(σ1). On the other hand, any function defined on σ2 can clearly be written in the form f (x + iy) = g(y − x) for some function g of one real variable. Lemma 3.12 and Proposition 3.10 of [3] then imply that the norm for f ∈ AC(σ2) is given by kf kBV (σ2) = kf k∞ + max(cid:0)f (1) − f (0), f (i) − f (0), f (i) − f (1)(cid:1). Any isomorphism must map idempotents to idempotents. However it is easy to see that while all idempotents in AC(σ2) have variation at most 1, the algebra AC(σ1) contains the idempotent with f (0) = f (2) = 0 and f (1) = 1 which has variation 2. Thus these algebras can not be isometrically isomorphic. 6 IAN DOUST AND MICHAEL LEINERT In the other direction, suppose that α : R2 → R2 is an invertible affine It is clear from the definition of variation that kf kBV (σ) = transformation. kf ◦ α−1kBV (α(σ)). Since affine transformations preserve polynomials, it is clear that AC(σ) is isometrically isomorphic to AC(α(σ)). (This is a very small ex- tension of [3, Theorem 4.1].) In Section 4 we shall extend this to slightly more general transformations of the plane, at the expense of the algebra isomorphism no longer being isometric. 3. The disk and the square Let Q = [0, 1] × [0, 1] ⊆ R2 denote the closed unit square and let D = {x ∈ R2 : kxk ≤ 1} denote the closed unit disk. These sets are clearly homeomorphic. The aim of this section is to show that AC(Q) 6≃ AC(D). Theorem 3.1. AC(Q) and AC(D) are not isomorphic as algebras. Proof. Suppose that j : AC(Q) → AC(D) is an algebra isomorphism. By The- orem 2.6, the map j is continuous and hence kj(f )kAC(D) ≤ kjk kf kAC(Q) for all f ∈ AC(Q). Let h : Q → D be the homeomorphism associated with j. Then h([0, 1] × {0}) is a closed arc on the unit circle ∂D. Let n ∈ N be even. For 0 ≤ k ≤ n, let pk = h( k n , 0). Now choose ǫn > 0 small enough so that, for every odd k, the ǫn-disc with centre pk does not meet the line segment pk−1 pk+1. Now let δn > 0 be chosen (using the uniform continuity of h) so that if x, x′ ∈ Q with kx − x′k ≤ δn then kh(x) − h(x′)k < ǫn. Without loss we may assume that δn ≤ 1. Q h (cid:0) k n , δn(cid:1) k−1 n k n k+1 n D pk+1 pk pk pk−1 Figure 2. The construction of pk in the proof of Theorem 3.1. For each odd k, let pk = h( k n , δn) ∈ B(pk, ǫn). Let S = [p0, p1, p2, p3, . . . , pn−1, pn] ⊆ D. It is easy to see that the points of S form the vertices of a convex subset of D and so, in particular, vf(Sn) = 2. Consider the map fn : Q → R defined by fn(x, y) = min(y/δn, 1). Clearly fn ∈ AC(Q) with kfnkAC(Q) = 2. Define gn : D → R by gn = fn ◦ h−1 = j(fn). ISOMORPHISMS OF AC(σ) SPACES 7 Then gn(pk) = 0 for k even, and gn(pk) = 1 for k odd. Thus kgnkAC(D) ≥ var(gn, D) ≥ cvar(gn, Sn) vf(Sn) = n 2 . But for all n, kgnkAC(D) ≤ kjk kfnkAC(Q) ≤ 2 kjk and hence we have a contradic- tion. (cid:3) As we shall now show, there are severe restrictions on the behaviour of any algebra isomorphism which is associated with a C 2 homeomorphism from Q to another compact subset of the plane. Lemma 3.2. Suppose that Ω ⊆ R2 is compact. Then a set ℓ ⊆ Ω is a closed line segment if and only if it is closed, convex and can be disconnected by the removal of a single point. Proof. The forward implication is obvious. Suppose now that the second condition holds. Let x ∈ ℓ denote a point whose removal splits ℓ \ {x} into disjoint sets ℓ1 and ℓ2. Choose points y1 ∈ ℓ1 and y2 ∈ ℓ2. Then y1 y2 lies inside ℓ. and must pass through x as ℓ1 ∪ ℓ2 is not connected. Since y2 was an arbitrary element of ℓ2, the line through y1 and x contains every element of ℓ2 -- and similarly every element of ℓ1 must lie on the same line. Thus ℓ is a closed convex subset of a line, or in other words, a line segment. (cid:3) Recall that a set U is mid-point convex if 1 2(x + y) ∈ U for all x, y ∈ U. Lemma 3.3. Suppose that Ω ⊆ R2 is compact and that ℓ = {x + λv : 0 ≤ λ ≤ 1} ⊆ Ω is a closed line segment. Let h : Ω → σ ⊆ R2 be a homeomorphism. Then h(ℓ) is a line segment if and only if it is mid-point convex. Proof. Again the forward implication is clear. Now suppose that h(ℓ) is mid-point convex. Since h is a homeomorphism, h(ℓ) is closed and can be disconnected by a point. Since h(ℓ) is closed and mid-point convex, it is convex, and so the result follows from the previous lemma. (cid:3) Lemma 3.4. Suppose that σ ⊆ R2 and that h : Q → σ is a homeomorphism. For y ∈ [0, 1] let ℓy = [0, 1] × {y}. If h(ℓ0) is not a line segment, then there exists δ > 0 such that h(ℓy) is not a line segment for any y ∈ [0, δ]. Proof. As h(ℓ0) is not a line segment we may choose x, x′ ∈ [0, 1] such that v = 1 2(h(x, 0)+h(x′, 0)) is not an element of h(ℓ0). Let ǫ = d(v, h(ℓ0)) > 0. Now choose δ > 0 small enough so that if u, u′ ∈ Q with ku − u′k ≤ δ then kh(u) − h(u′)k < ǫ/3. Suppose that 0 ≤ y ≤ δ and that h(ℓy) is a line segment. Since h(ℓy) is mid- 2 (h(x, y) + h(x′, y)). But point convex, there exists t ∈ [0, 1] such that h(t, y) = 1 8 then IAN DOUST AND MICHAEL LEINERT kh(t, 0) − vk ≤ (cid:13)(cid:13)(cid:13) h(t, 0) − h(t, y)(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13) + 0 + < ǫ ǫ 3 ≤ ǫ 3 contradicting that d(v, h(ℓ0)) = ǫ. h(x, y) + h(x′, y) h(t, y) − h(x, y) + h(x′, y) − 2 2 h(x, 0) + h(x′, 0) (cid:13)(cid:13)(cid:13) 2 (cid:13)(cid:13)(cid:13) (cid:3) A consequence of this result is that if h : Q → σ is a homeomorphism and there exists a line segment ℓ ∈ Q such that h(ℓ) is not a line segment, then we may assume that both ℓ and h(ℓ) lie in the interiors of their respective sets. Theorem 3.5. Suppose that ∅ 6= σ ⊆ R2 is compact and that j : AC(Q) → AC(σ) is an algebra isomorphism with associated homeomorphism h : Q → σ. If h is C 2 then h maps line segments to line segments. Proof. Suppose that there exist σ, j and h as above such that for some line segment ℓ ⊆ Q, h(ℓ) is not a line segment. We shall show that this leads to a contradiction. In order to streamline the proof, a number of simplifications can be made. By the above remark we can assume that ℓ lies in the interior of Q. Since h is C 2, h(ℓ) has a tangent at each point and the curve can at least locally be considered as the graph of a C 2 function of a parametrization of this tangent line. Since h(ℓ) is not a line segment, one can therefore choose an invertible affine map β : R2 → R2 such that there is a subsegment ℓ0 of ℓ such that (1) (β ◦ h)(ℓ0) = {(s, t(s)) : 0 ≤ s ≤ 1}, and (2) t′′(s) > 0 for 0 < s < 1. Now choose an invertible affine map α : R2 → R2 such that (1) α(ℓ0) = [0, 1] × {0} ⊂ [0, 1] × [−1, 1] ⊆ int(α(Q)), and (2) (β ◦ h ◦ α−1)([0, 1] × [0, 1]) lies above (β ◦ h)(ℓ0). We shall write h1 = β ◦ h ◦ α−1 for the homeomorphism from α(Q) to β(σ), and C for the curve (β ◦ h)(ℓ0). (See Figure 3.) The proof now mimics that of Theorem 3.1. Let n ∈ N be even. For 0 ≤ k ≤ n )) and let xk ∈ [0, 1] denote the (unique) number such that n, t( k n, let pk = ( k h1(xk, 0) = pk. Choose ǫn small enough so that for all odd k, the ball B(pk, ǫn) lies beneath the line segment pk−1 pk+1. (This is of course possible by the convexity of the function t.) Now choose 0 < δn < 1 so that if x, y ∈ α(Q) with kx − x′k ≤ δn then kh1(x) − h1(x′)k < ǫn. ISOMORPHISMS OF AC(σ) SPACES 9 α(Q) h1 α(ℓ0) β(σ) C 1 Figure 3. The homeomorphism h1 : α(Q) → β(σ) in the proof of Theorem 3.5. For each odd k, let pk = h1(xk, δ) ∈ B(pk, ǫ). Then pk lies above the curve C but below the chord pk−1 pk+1. Let Sn = [p0, p1, p2, p3, . . . , pn−1, pn], so that as in the proof of Theorem 3.1, vf(Sn) = 2. Consider the map fn : α(Q) → R defined by fn(x, y) = y/δn, if y ≥ δn, if 0 ≤ y < δn, if y < 0. 1, 0,   Clearly fn ∈ AC(α(Q)) with kfnkAC(α(Q)) = 2. Again define gn : β(σ) → R by gn = fn ◦ h−1 to produce a function with kgnkAC(β(σ)) ≥ n/2. But, noting the 1 remarks at the end of Section 2 about the invariance of variation norms under affine transformations, we have that for all n, kgnkAC(β(σ)) ≤ kjk kfnkAC(α(Q)) ≤ 2 kjk which is the required contradiction. (cid:3) Functions mapping line segments to line segments have been studied by various authors. We refer the reader to [8] for further details. It is worth noting that such maps need not be affine. For example h(x, y) = ((x + 1)/(y + 1), 2y/(y + 1)), which maps the unit square to the trapezoid with vertices (1, 0), (2, 0), (1, 1) and ( 1 2, 1), is a nonaffine line-segment preserving function. 4. Half-plane-affine maps Despite the results of the previous section there are some positive statements that can be made about when AC(σ1) and AC(σ2) are isomorphic. As we noted earlier, this is certainly the case if σ2 is the image of σ1 under an affine home- omorphism α. In this section we weaken this condition on the homeomorphism mapping σ1 to σ2. Definition 4.1. A half-plane splitting of R2 is a pair of closed half-planes {H1, H2} whose union is R2 and which only intersect along their shared boundary line. 10 IAN DOUST AND MICHAEL LEINERT Definition 4.2. An invertible map α : R2 → R2 is said to be a half-plane-affine map if there exists a half-plane splitting {H1, H2} of R2 and two affine maps α1, α2 : R2 → R2 such that α(x) = αj(x) whenever x ∈ Hj. We shall write α = {α1, α2}H1,H2. Any half-plane-affine map is clearly continuous. The assumption of invertibility ensures that {α(H1), α(H2)} is a half-plane splitting of R2. We also have the following easy fact. Lemma 4.3. The inverse of a half-plane-affine map is a half-plane-affine map. Suppose for the remainder of this section that α = {α1, α2}H1,H2 is a half-plane- affine map. We shall show below that for such maps BV (σ) ≃ BV (α(σ)). The main point in proving this is showing that given any finite ordered list S of elements of σ, vf(S) is comparable to vf(α(S)). Heuristically, if the number of times that the curve γS crosses a line ℓ is k, then γα(S) should cross either α1(ℓ) or α2(ℓ) at least k 2 times. Proving this is a little delicate however because in the case that a segment si = xi xi+1 has at least one of its endpoints on the line ℓ, it is possible that si is a crossing segment on ℓ but that α(xi) α(xi+1) is not a crossing segment on either α1(ℓ) or α2(ℓ). Lemma 4.4. Let α : R2 → R2 be a half-plane affine map. Suppose that S = [x0, . . . , xn] is a finite ordered list of points in R2 and that S = [v0, . . . , vn] is the list of the images of these points under α. Then 1 2 vf(S) ≤ vf( S) ≤ 2 vf(S). Proof. By the previous lemma it suffices to just prove the left-hand inequality. Let k = vf(S) and fix a line ℓ such that vf(S, ℓ) = k. For 0 ≤ i ≤ n − 1 let si = xi xi+1 and let si = vi vi+1. Our aim is to find a correspondence between crossing segments of S on ℓ and crossing segments of S on either α1(ℓ) or α2(ℓ). If ℓ = α(ℓ) is also a line then si ∈ X(S, ℓ) if and only if si ∈ X( S, ℓ) and so vf( S) ≥ vf(S) which certainly gives the required inequality. This occurs in particular if ℓ is parallel to the boundary line between H1 and H2. If ℓ is not a line, then α1(ℓ) and α2(ℓ) do not coincide, and there must exist a unique point w ∈ ℓ that lies on the shared boundary of H1 and H2. Suppose that si ∈ X(S, ℓ). If xi and xi+1 lie strictly on opposite sides of ℓ, then vi and vi+1 lie strictly on opposite sides of ℓ, and so si is a crossing segment for S for at least one of α1(ℓ) or α2(ℓ). The more difficult situation is if one of the endpoints of si lies on ℓ. Referring to Definition 2.1 we have the following possibilities: (i) i = 0 and xi ∈ ℓ. Then v0 ∈ ℓ and hence s0 is a crossing segment of S on at least one of α1(ℓ) or α2(ℓ). ISOMORPHISMS OF AC(σ) SPACES 11 (ii) i > 0, xi ∈ ℓ and xi−1 6∈ ℓ. Note that in this case si−1 6∈ X(S, ℓ). Without loss of generality we may label the half-planes so that xi ∈ H1. Now si is a crossing segment of S on α1(ℓ) except in the case that vi−1 ∈ α1(ℓ) If vi−1 ∈ α1(ℓ) then, as α1(ℓ) and α2(ℓ) do not coincide, (see Figure 4). xi−1 ∈ H2 and vi−1 6∈ α2(ℓ). We must now distinguish the case when xi lies in the interior of H1 and the case when xi lies in the boundary of H1. If xi = w then vi ∈ α2(ℓ) so since vi−1 6∈ α2(ℓ) we have si ∈ X( S, α2(ℓ)). If xi 6= w, then (as in Figure 4) vi−1 and vi lie on opposite sides of α2(ℓ) and hence si−1 ∈ X( S, α2(ℓ)). α xi+1 xi xi−1 ℓ α1(ℓ) vi+1 vi vi−1 α2(ℓ) H1 H2 α(H1) α(H2) Figure 4. Mapping of crossing segments in the proof of Lemma 4.4. Thus, while only one of si−1 and si are crossing segments of S on ℓ, at least one of si−1 and si is a crossing segment of S on either α1(ℓ) or α2(ℓ). (iii) i = n − 1, xi 6∈ ℓ and xi+1 ∈ ℓ. Again we may assume that xn ∈ H1 and hence that vn ∈ α1(ℓ). If vn−1 6∈ α1(ℓ) then clearly sn−1 ∈ X( S, α1(ℓ)). On the other hand, if vn−1 ∈ α1(ℓ) then one may argue as in (ii) that sn−1 ∈ X( S, α2(ℓ)). As before then, sn−1 is a crossing segment of S on at least one of α1(ℓ) or α2(ℓ). Let k1 and k2 be the number of crossing segments of S on α1(ℓ) and α2(ℓ) respec- tively. Then the above discussion shows that k1 + k2 ≥ k. It follows therefore that vf( S) ≥ k (cid:3) 2 as required. Theorem 4.5. Let α : R2 → R2 be a half-plane affine map. Suppose that σ1 is a nonempty compact subset of R2 and that σ2 = α(σ1). Then BV (σ1) ≃ BV (σ2) and AC(σ1) ≃ AC(σ2). Proof. Suppose that α = {α1, α2}H1,H2 and that ℓ0 is the boundary line between H1 and H2. For f ∈ BV (σ1) let f : σ2 → R be defined by f (α(x)) = f (x), x ∈ σ1. The first step is to show that f ∈ BV (σ2). 12 IAN DOUST AND MICHAEL LEINERT By the previous lemma cvar( f , S) vf( S) ≤ 2 cvar(f, S) vf(S) ≤ 2 var(f, σ1). Taking the supremum over all such finite lists S shows that k fkBV (σ2) ≤ 2 kf kBV (σ1) and in particular that f ∈ BV (σ2). Let j : BV (σ1) → BV (σ2) be defined by j(f ) = f . It is clear that j is a continuous algebra homomorphism. Lemma 4.3 can now be used to deduce that j is also onto and hence that j is a Banach algebra isomorphism. Suppose that g ∈ CTPP (σ1). Then j(g) will also be planar on polygonal regions (on a neighbourhood) of σ2. Indeed j is a bijection from CT P P (σ1) to CT P P (σ2) and hence j is also a bijection between the closures of these sets, AC(σ1) and AC(σ2). (cid:3) As the example below shows, the factor of 2 in the above proof is necessary. Example 4.6. Suppose that σ1 is the nonconvex quadrilateral with vertices at (1, 0), (0, 4), (−1, 0) and (0, 2). Note that σ1 is the image of the closed unit square under a half-plane-affine map. Theorem 4.5 then implies that AC(Q) ≃ AC(σ1). Define f (x, y) = max(1 − y, 0), (x, y) ∈ σ1. Then, as f only varies in the y direction, it is clear that kf k∞ = 1, that var(f, σ1) = 1 and hence that kf kBV (σ1) = 2. Now write f = f1 + f2 where f1(x, y) = f (x, y)χ[−1,0](x) and f2(x, y) = f (x, y)χ[0,1](x). Note that both f1 and f2 are in CTPP (σ1) ⊆ AC(σ1) and that these functions have disjoint supports. Suppose now that j : AC(σ1) → AC(Q) is a Banach algebra isomorphism, with associated homeomorphism h : σ1 → Q. Let g1 = j(f1), g2 = j(f2) and g = j(f ). Using Theorem 2.6(1), we can choose points z1, z2 ∈ Q such that gk(zℓ) = δkℓ. Let γ denote the line segment joining z1 and z2. Then h−1(γ) is a continuous curve in σ1 which necessarily passes though some point (0, y) ∈ σ1. Let w = h(0, y) be the corresponding point on γ. Then g(w) = g1(w) + g2(w) = f1(0, y) + f2(0, y) = 0. But this implies that cvar(g, γ) ≥ 2 and hence var(g, σ2) ≥ 2. By Corollary 2.7 we see that j can not be isometric. In particular, this example shows that factor of 2 that appears in the proof of Theorem 4.5 is necessary. Remark 4.7. A simple adjustment to the above example, replacing the vertex (0, 2) with the point (0, 0), shows that if T is a closed triangular region in R2, then AC(T ) ≃ AC(Q). Thus isomorphism class does not distinguish between the number of vertices in polygonal regions. We shall come back to this issue later in the paper. ISOMORPHISMS OF AC(σ) SPACES 13 5. Locally piecewise affine maps The results of the previous section can be extended to cover homeomorphisms of the plane made up from more than two affine maps. Let α : R2 → R2 be an invertible affine map, and let C be a convex n-gon. Then α(C) is also a convex n-gon. Denote the sides of C by s1, . . . , sn. Suppose that x0 ∈ int(C). The point x0 determines a triangulation T1, . . . , Tn of C, where Tj is the (closed) triangle with side sj and vertex x0. A point y0 ∈ int(α(C)) determines a similar triangularization T1, . . . , Tn of α(C), where the numbering is such that α(sj) ⊆ Tj. The following fact is then clear. Lemma 5.1. With the notation as above, there is a unique map h : R2 → R2 such that (1) h(x) = α(x) for x 6∈ int(C), (2) h maps Tj onto Tj, for 1 ≤ j ≤ n. (3) αj = hTj is affine, for 1 ≤ j ≤ n. (4) h(x0) = y0. We shall say that h is the locally piecewise affine map determined by (C, α, x0, y0). h α(C) C x0 y0 Figure 5. The locally piecewise affine map h determined by (C, α, x0, y0). It is clear that h is necessarily continuous and invertible. Indeed the following result is straightforward. Lemma 5.2. Let h be the locally piecewise affine map determined by (C, α, x0, y0). Then h−1 is the locally piecewise affine map determined by (h(C), α−1, y0, x0). In the last section we shall repeatedly use the following special case of Lemma 5.1 applied with α = id, the identity mapping on R2. Lemma 5.3. Suppose that T and T are two triangles in R2 with vertices a, b, c and a, b, c respectively. Suppose that Q is a convex quadrilateral in R2 which contains T and T and which has bc as one side. Then the locally piecewise affine map determined by (Q, id, a, a) maps T onto T and fixes all points outside of Q. 14 IAN DOUST AND MICHAEL LEINERT h c T b Q a a c T b Q a a Figure 6. A locally piecewise affine map moving T to T . Our first aim is to show that for any locally piecewise affine map h, AC(σ) ≃ AC(h(σ)). Lemma 5.4. Suppose that h is a locally piecewise affine map determined by (α, C, x0, y0) where C is a convex n-gon. Let S = [w0, w1, . . . , wm] be a list of elements in R2 and let S = [h(w0), h(w1), . . . , h(wm)]. Then 1 cn vf(S) ≤ vf( S) ≤ cn vf(S) for some positive constant cn which is independent of S. Proof. By the previous lemma it suffices to prove either one of the inequalities. For notational simplicity, we shall write si = wi wi+1, for 0 ≤ i ≤ m − 1 and write vi = h(wi) for 0 ≤ i ≤ m. Let T1, . . . , Tn denote the subsets of the plane defined at the start of this section, and let T0 = R2 \ int(C). Suppose then that vf(S) = k and that ℓ is a line such that vf(S, ℓ) = vf(S). At least one of the regions T0, . . . , Tn has at least k1 = ⌈ k n+1⌉ crossing segments of S on ℓ with at least one of their endpoints in that region. Suppose that this region is Tr. Let K denote the set of indices i such that si ∈ X(S, ℓ) and si has at least one endpoint in Tr. Our aim is to show that each of these crossing segments corresponds to a crossing segment of S on one of a finite number of lines. This will require a careful consideration of cases. Let δ = min{d(vi, h(Tr)) : 0 ≤ i ≤ n and vi 6∈ Tr}. We take the minimum of the empty set to be zero. Suppose first that 1 ≤ r ≤ n. Then h(Tr) is a triangle. Choose a triangle T with sides parallel to those of h(Tr), which contains h(Tr) in its interior, and such that if v ∈ T then d(v, h(Tr)) < δ 2 . Let ℓ1,ℓ2 and ℓ3 denote the three lines forming the sides of T (see Figure 7). From this construction, every segment si with i ∈ K either lies entirely inside h(Tr), or else it is a crossing segment for S on at least one of the lines ℓ1, ℓ2 or ℓ3. ISOMORPHISMS OF AC(σ) SPACES 15 Let ℓ0 denote the line which is the image of ℓ under αr (considered as extended to the whole plane). T ℓ0 ℓ2 ℓ1 h(Tr) ℓ3 Figure 7. The construction of ℓ0, ℓ1 ,ℓ2 and ℓ3. Suppose then that i ∈ K, and that both wi and wi+1 lie in Tr. Referring to Definition 2.1 there are four possibilities: (i) wi and wi+1 lie on strictly opposite sides of ℓ. In this case vi and vi+1 lie on strictly opposite sides of ℓ0 and so si ∈ X( S, ℓ0). (ii) i = 0 and wi ∈ ℓ. Clearly then vi ∈ ℓ0 and so si ∈ X( S, ℓ0). (iii) i > 0, wi ∈ ℓ and wi−1 6∈ ℓ. In this case si−1 6∈ X(S, ℓ) and either: (a) wi−1 ∈ Tr. In this case vi ∈ ℓ0 and vi−1 6∈ ℓ0 and hence si ∈ X( S, ℓ0). (b) wi−1 6∈ Tr. In this case si need not be in X( S, ℓ0) since vi−1 might lie on ℓ0. However, si−1 must be a crossing segment for S on one of the boundary lines ℓ1, ℓ2 or ℓ3. (iv) i = m − 1, wi 6∈ ℓ and wi+1 ∈ ℓ. Again si ∈ X( S, ℓ0). Suppose next that i ∈ K and that one of wi and wi+1 does not lie in Tr. As noted above, in this case si ∈ X( S, ℓj) for some j = 1, 2, 3. At this stage we have shown that if 1 ≤ r ≤ n, then there are at least k1 segments of S which are crossing segments for at least one of the lines ℓj with 0 ≤ j ≤ 4. Thus, for at least one of these values of j, vf( S, ℓj) ≥ ⌈ k1 4 ⌉. The remaining case is where r = 0 and so Tr is not a triangle. The proof in this case is almost identical except that now one must work with (1) lines ℓ1, . . . , ℓn chosen close to the boundary of h(T0) so that all the end- points vi which are not in T0 lie inside the smaller n-gon determined by these lines, and (2) the line ℓ0 = α(ℓ) (see Figure 8). Every segment si = vi vi+1 with only one endpoint in T0 lies in X( S, ℓj) for at least one j with 1 ≤ j ≤ n. 16 IAN DOUST AND MICHAEL LEINERT Following the proof above one can then show that there are (at least) k1 seg- ments of S which are crossing segments for at least one of the lines ℓj where 0 ≤ j ≤ n, and hence vf( S, ℓj) ≥ ⌈ k1 n+1⌉ for at least one value of j in this range. ℓ3 ℓ4 h(T0) ℓ0 = α(ℓ) ℓ1 ℓ5 ℓ2 Figure 8. Choosing ℓ0, . . . , ℓn when r = 0 (and n = 5). In either case then vf( S) ≥ (cid:24) k1 n + 1(cid:25) = (cid:24)⌈ vf(S)/(n + 1) ⌉ n + 1 (cid:25) ≥ vf(S) (n + 1)2 . (cid:3) The above proof of course shows that cn ≤ (n + 1)2. Heuristically we expect that cn = n + 1 but we are unable to prove this. Theorem 5.5. Suppose that σ is a nonempty compact subset of the plane, and that h is a locally piecewise affine map. Then BV (σ) ≃ BV (h(σ)) and AC(σ) ≃ AC(h(σ)) Proof. For f ∈ BV (σ), let f : h(σ) → C be defined by f (h(x)) = f (x). Suppose that f ∈ BV (σ) and that S = [v0, . . . , vm] is a list of points in h(σ). Let S = [w0, . . . , wm] ⊂ σ denote the list of preimages of the points in S. Then, using the notation of Lemma 5.4, cvar( f , S) vf( S) = cvar(f, S) vf( S) ≤ Cn cvar(f, S) vf(S) ≤ Cn var(f, σ). Thus, f is of bounded variation with k f kBV (h(σ)) ≤ (1 + Cn) kf kBV (σ). It follows, using Lemma 5.2 that the map j : f 7→ f is a bounded isomorphism from BV (σ) onto BV (h(σ)). It is clear that j maps CTPP (σ) onto CTPP (h(σ)) and hence that j provides (cid:3) an isomorphism from AC(σ) onto AC(h(σ)). ISOMORPHISMS OF AC(σ) SPACES 17 6. Polygons and ears The main result from this section is that given any two simple polygons P1 and P2 we have that AC(P1) ≃ AC(P2). The proof requires a nice fact from computational geometry called the 'Two Ears Theorem' which was proven by Meisters [13]. Given a, b ∈ R2 we shall let a b ◦ denote the 'open' line segment between a and b, that is ◦ a b = {λa + (1 − λ)b : 0 < λ < 1}. Let v be a vertex of a polygon P and suppose that a, b are the neighbouring lies entirely in the interior of P . vertices to v. We say that v is an ear of P if a b ◦ Theorem 6.1 (Two Ears Theorem). Every simple polygon with more than 3 vertices has at least 2 ears. A simple consequence of the Two Ears Theorem is that it is possible to trian- gulate any polygon. That is, given any polygon P , one may construct a finite family of 'disjoint' triangles {Tn} whose vertices are all vertices of P and whose union equals P . u Q v a P T w b Figure 9. Lemma 6.2. Lemma 6.2. Suppose that v is an ear of a polygon P . Let T = △avb denote the triangle formed by the two sides of P which meet at v and the corresponding diagonal a b. Then there exists a convex quadrilateral Q with vertices a, u, b and w such that (1) T \ {a, b} ⊆ int(Q), (2) a u◦ and b u (3) a w◦ and b w ◦ lie in the complement of P , and lie in the interior of P . ◦ 18 IAN DOUST AND MICHAEL LEINERT Proof. We begin by triangulating P using the standard algorithm of removing one ear at a time. We may clearly start by dealing with the ear at v and so T is one of the triangles in our triangulation. The diagonal a b must form an edge of two of the triangles, namely T , and another which we shall denote by T1. One can choose w to be any interior point of T1 and this will clearly have property 3. Let m denote the midpoint of a b and let ℓ denote the median of T that passes through v and m. As we shall see below, if we demand that w also lies on ℓ then this will ensure that the quadrilateral Q is convex. (See Figure 10.) v ℓ T a m b w T1 Figure 10. Construction of w. Finding a suitable point u is slightly more delicate. For small t > 0, u(t) = (1 + t)v − tw lies in the complement of P . If a u(t) does not lie in the complement of P then it must be the case that some vertices of P lie in the interior of the quadrilateral avbu(t) or on one of the boundary lines a u(t) or b u(t). Since there can be only finitely many such vertices, by choosing t0 > 0 sufficiently small we can ensure that u = u(t0) satisfies condition (2). or b u(t) ◦ ◦ Property (1) is clear so it remains to check convexity. By the construction, the and m ∈ u w◦. But by two diagonals of Q will meet at m and clearly m ∈ a b Theorem 6.7.9 of [15] a quadrilateral is convex if and only if the diagonals meet at a point in the interior of these diagonals, and hence Q is convex. (cid:3) ◦ Theorem 6.3. Suppose that P1 and P2 are simple polygons. Then AC(P1) ≃ AC(P2). Proof. We shall use induction to prove that the statement S(n) : if P is any simple n-gon and T is a triangle, then AC(P ) ≃ AC(T ) holds for all n ≥ 3. ISOMORPHISMS OF AC(σ) SPACES 19 The statement is true for n = 3 since one can find an affine map between any two triangles. Suppose then that n > 3 and that the S(m) is true for all m with 3 ≤ m < n. Let P be an n-gon with vertices v1, . . . , vn. By the Two Ears Theorem, there exists an ear vj. Let Tvj be the triangle with vertices at vj−1, vj and vj+1. Using Lemma 6.2 fix a convex quadrilateral Q with vertices at vj−1 and vj+1 and two additional points u, w chosen so that Tvj \ {vj−1, vj+1} lies in the interior of Q and so that vj−1 w◦ and vj+1 w◦ lie in the interior of P . As Q is convex, the point vj and the midpoint m of vj−1 vj+1 are both interior points of Q. Q u vj vj+1 h(u) h(Q) h(vj+1) vj−1 m P h h(vj) h(m) w h(vj−1) h(w) h(P ) Figure 11. The action of h. Let α denote the identity map on R2, and let h denote the unique locally piecewise affine map determined by (Q, α, vj, m). This map sends vj−1 vj to vj−1 m and vj vj+1 to m vj+1 (see Figure 11). All the other edges of P lie in the complement of Q and are therefore fixed. It follows therefore that the image of P under h is an m-gon for some m < n. By Theorem 5.5, AC(P ) ≃ AC(h(P )). But by the induction hypothesis, AC(h(P )) ≃ AC(T ) for any triangle T and so the proof is complete. (cid:3) 7. Polygonal regions with holes The results of the last section have a natural extension to a wider class of regions. Let P be a simple polygon in the plane. A set W is a window in P if it is the interior of a polygon P ′ where P ′ lies in the interior of P . We shall say that a compact set σ is a polygonal region of genus n if there exists a simple polygon P with n nonoverlapping windows W1, . . . , Wn such that σ = P \ (W1 ∪ · · · ∪ Wn) and write G(σ) = n for the genus of σ. 20 IAN DOUST AND MICHAEL LEINERT Figure 12. A polygonal region of genus 3. If σ1 and σ two are polygonal regions of differing genus, then these sets are not homeomorphic and hence AC(σ1) 6≃ AC(σ2). In this section we shall show that show within this class of sets, the isomorphism class of the the corresponding function algebras is completely determined by their genus. This is achieved by showing that there is always a finite sequence of locally piecewise affine maps whose composition sends σ1 to σ, and then applying Theorem 5.5. One of the main tools in doing this is to show that via such maps, one may 'move' triangular windows anywhere within any rectangle that contains no other other windows. Lemma 7.1. Suppose that R is a rectangle and that T = △abc and T ′ = △a′b′c′ are two triangles in the interior of R. Then there is a continuous bijection h : R2 → R2 such that (1) h can be written as a composition of finitely many locally piecewise affine maps, (2) h(x) = x for all x 6∈ R, (3) h(R) = R, and (4) h(T ) = T ′. Proof. Choose ǫ > 0 such that no point of T or T ′ lies within distance 2ǫ of the boundary of R. We shall call the four interior points of R which lie at distance ǫ along the diagonals from the vertices of R, the ǫ-corner points of R. Fix any three of these ǫ-corner points and let T0 denote the triangle with these points as vertices. We shall show that there is a function h satisfying (1), (2) and (3) and such that h(T ) = T0. The same proof of course would construct a corresponding map sending T ′ to T0. Since the inverse of a locally piecewise affine map is also locally piecewise affine, this produces a finite sequence of locally piecewise affine maps which has properties (1) -- (4). The line through b and c splits R into two convex polygons. Let P denote the polygon which contains a. At least one of the vertices of P , say v1, is a vertex of R not lying on the line through b and c. ISOMORPHISMS OF AC(σ) SPACES 21 Let a1 be the ǫ-corner point of R near v1. Using the triangulations of P generated by a and by a1, Lemma 5.1 produces a locally piecewise affine map h1 which is the identity outside of P , and which maps a to a1. Indeed, as T lies entirely in a region on which h1 is affine, h1 maps T to the triangle △a1bc (see Figure 13). v1 P a c b v1 h1 a1 c b Figure 13. Moving the first vertex in Lemma 7.1. Consider now the quadrilateral Q with vertices at a1 and the three vertices of R other than v1. Note that the position of a1 ensures that Q is convex, and hence that the line through a1 and c splits Q into two convex polygons. Let P1 denote the polygon containing b. One of the vertices of Q adjacent to a1 (which is therefore also a vertex of R) must lie in P1. Denote this vertex by v2 and let b1 be the ǫ-corner point of R near v2. (Note that b1 must lie in P1.) Applying Lemma 5.1 again we produce a locally piecewise affine map h2 which is the identity outside of Q and which maps △a1bc onto △a1b1c. Finally, consider the convex quadrilateral Q1 with vertices a1, b1 and the two remaining vertices of R. Let c1 be the ǫ-corner point of R near one of these remaining vertices of R. Noting that c and c1 are both in the interior of Q1 we can find a locally piecewise affine map h3 which is the identity outside of Q1 and which maps △a1b1c onto △a1b1c1. The vertices of this final triangle are all ǫ-corner points. With one or two further applications of locally piecewise affine maps we can arrange that the image of T under this composition of maps is T0. (cid:3) Theorem 7.2. Suppose that σ1 and σ2 are polygonal regions of genus n1 and n2. Then AC(σ1) ≃ AC(σ2) if and only if n1 = n2. Proof. As noted above it only remains to show the 'if' part of the theorem. Fix a genus n. Let τ denote the polygonal region of genus n τ = T \ (T1 ∪ · · · ∪ Tn) 22 IAN DOUST AND MICHAEL LEINERT where T is the triangle with vertices at (0, −1), (1, 0) and (0, 1) and, for k = 1, . . . , n, the window Tn is the triangle with vertices at ( 3k−2 3n , 0) and ( 2k−1 3n ). We shall proceed by showing that if σ is any polygonal regions of genus n, then AC(σ) ≃ AC(τ ). 3n , 0), ( 3k−1 2n , 1 Q vj V h h(Q) h(vj) h(V ) Figure 14. Reducing the number of edges in a window. Suppose then that σ = P \ (W1 ∪ · · · ∪ Wn). The image of σ under any locally piecewise affine map is also a polygonal region of genus n and as before, the isomorphism class of the corresponding AC function space is preserved under such maps. By applying a finite sequence of locally piecewise affine maps as in the proof of Theorem 6.3 we may reduce the number of vertices in P to 3. Note that the effect of these maps might be to increase the number of vertices in some of the windows. By applying a suitable affine map then, we see that AC(σ) ≃ AC(σ′) where σ′ = T \ (V1 ∪ · · · ∪ Vn), and where V1, . . . , Vn are windows in T . The same algorithm can now be used to reduce the windows V1, . . . , Vn to triangles. Specifically, suppose that V is a window in T with at least 4 vertices v1, . . . , vk. By the Two Ears Theorem we can choose an ear vj in V . Since σ′ can be triangulated, the proof of Theorem 6.2 allows us to choose a convex quadrilateral Q containing the triangular region vj−1 vjvj+1 but not intersecting any of the other windows of σ′. Applying a suitable locally piecewise affine map h which fixes the complement of Q and maps vj to the midpoint of vj−1 vj+1 we reduce the number of vertices in V while leaving all the other windows unchanged. (See Figure 14.) ISOMORPHISMS OF AC(σ) SPACES 23 It just remains to prove that if σ′ = T \ (V1 ∪ · · · ∪ Vn) where each window is a triangle, then we can apply a finite sequence of locally piecewise affine maps to move the triangles {Vk} to the corresponding triangles {Tk} in the description of our standard set τ . Our main tool is Lemma 7.1 which allows us to move a triangle anywhere within the interior of rectangular region while leaving everything outside the rectangle undisturbed. Although in concrete examples it is easy to efficiently move the triangles to their final position, for completeness we shall now give a general algorithm shows that this is always possible. Note that it follows from Lemma 5.3 that one may always move a vertex of a triangle to any point in the interior of that triangle, or, by applying two such moves, shrink any triangle towards one of its vertices. We shall use the lexicographical ordering of points in the plane to choose the smallest vertex (xk, yk) for each of the triangles Vk. The steps in the algorithm are as follows. (1) Label the triangles so that x1 ≤ x2 ≤ · · · ≤ xn. (2) Starting from the right, shrink as many triangles (toward one vertex say) as is necessary to ensure that the x-coordinates of the smallest vertex of each of the triangles are distinct. (3) Starting from the left, shrink each triangle towards its smallest vertex. If the triangles are shrunk to a sufficiently small size then the projections of these triangles onto the x-axis will form disjoint intervals [ak, bk]. Indeed, after sufficient shrinking the triangles will sit within the interiors of disjoint rectangles Rk as in Figure 15. (4) Using Lemma 7.1 use a sequence of locally piecewise affine maps to move the kth triangle (within rectangle Rk to one with vertices at (ak, 0), (bk, 0) and (ak, ek) where ek is chosen small enough so that this triangle sits in Rk (5) It remains to move the triangles to the correct positions to form the stan- 3n ). Starting from the left, move , 0), dard configuration τ . Choose δ < min(a1, 1 each triangle in turn (using Lemma 7.1) so that it has vertices ( δ(3k−1) ( δ(3k−2) 3n 3n , 0) and ( δ(2k−1) 2n , 1 3n ). (6) Now starting from the right, one can move the k-th triangle to the standard triangle Tk. Since we have only applied a finite sequence of locally piecewise affine maps, AC(σ′) ≃ AC(τ ), and this completes the proof. (cid:3) 24 IAN DOUST AND MICHAEL LEINERT R3 R1 R2 Figure 15. Steps in the algorithm to map σ′ to τ : (2) making the smallest vertices distinct; (3) shrinking the triangles so they have disjoint projections on the x-axis. Acknowledgements. The authors would like to thank Michael Cowling and Han- ning Zhang for some helpful discussions regarding this work. References [1] D. Amir, On isomorphisms of continuous function spaces, Israel J. Math. 3 (1965), 205 -- 210. [2] Takao Asano, Tetsuo Asano and R. Y. Pinter, Polygon triangulation: efficiency and mini- mality, J. Algorithms 7 (1986), 221 -- 231 [3] B. Ashton and I. Doust, Functions of bounded variation on compact subsets of the plane, Studia Math. 169 (2005), 163 -- 188. [4] B. Ashton and I. Doust, A comparison of algebras of functions of bounded variation, Proc. Edinb. Math. Soc. (2) 49 (2006), 575 -- 591. [5] B. Ashton and I. Doust, AC(σ) operators, J. Operator Theory 65 (2011), 255 -- 279. [6] D. Bongiorno, Absolutely continuous functions in Rn, J. Math. Anal. Appl. 303 (2005) 119 -- 134. [7] H. B. Cohen, A bound-two isomorphism between C(X) Banach spaces, Proc. Amer. Math. Soc. 50 (1975), 215 -- 217. [8] A. Cap, M. G. Cowling, F. de Mari, M. Eastwood, and R. McCallum, The Heisenberg group, SL(3, R), and rigidity, in Harmonic analysis, group representations, automorphic forms and invariant theory, Lect. Notes Ser. Inst. Math. Sci. Natl. Univ. Singap., 12, 41 -- 52, World Sci. Publ., Hackensack, NJ, 2007. [9] M. Csornyei, Absolutely continuous functions of Rado, Reichelderfer, and Mal´y, J. Math. Anal. Appl. 252 (2000) 147 -- 166. [10] I. Doust and M. Leinert, Approximation in AC(σ), arXiv:1312.1806v1, 2013. [11] M. I. Garrido and J. A. Jaramillo, Variations on the Banach-Stone theorem, IV Curso Espacios de Banach y Operadores (Laredo, 2001), Extracta Math. 17 (2002), 351 -- 383. [12] I. Gelfand, and A. Kolmogoroff, On rings of continuous functions on topological spaces, Dokl. Akad. Nauk. SSSR 22 (1939), 11 -- 15. [13] G. H. Meisters, Polygons have ears, Amer. Math. Monthly 82 (1975), 648 -- 651. [14] M. H. Stone, Applications of the theory of Boolean rings to general topology, Trans. Amer. Math. Soc. 41 (1937), 375 -- 481. ISOMORPHISMS OF AC(σ) SPACES 25 [15] G. A. Venema, The Foundations of Geometry, Prentice Hall, Upper Saddle River, NJ, 2005. Ian Doust, School of Mathematics and Statistics, University of New South Wales, UNSW Sydney 2052 Australia E-mail address: [email protected] Michael Leinert, Institut fur Angewandte Mathematik, Universitat Heidel- berg, Im Neuenheimer Feld 294, D-69120 Heidelberg Germany E-mail address: [email protected]
1309.0201
1
1309
2013-09-01T10:50:19
Split functions, Fourier transforms and multipliers
[ "math.FA" ]
We study the effect of a splitting operator S_t on the L^p norm of the Fourier transform of a function f and on the operator norm of a Fourier multiplier m. Most of our results assume p is an even integer, and are often stronger when f or m has compact support.
math.FA
math
Split functions, Fourier transforms and multipliers Laura De Carli and Steve Hudson Abstract We study the effect of a splitting operator St on the Lp norm of the Fourier transform of a function f and on the operator norm of a Fourier multiplier m. Most of our results assume p is an even integer, and are often stronger when f or m has compact support. Mathematics Subject Classification: 42B15, 42B10 1. Introduction We study the effect of a splitting operator St on the Fourier transform of a function f . For f ∈ L2(R, C) ∩ L1(R, C), let τtf (x) = f (x − t). Let f+(x) = f (x)χ{x>0} and f−(x) = f (x)χ{x<0}, where χ denotes the characteristic function; so f = f+ + f− almost everywhere. For t > 0, let Stf (x) = τtf+ + τ−tf− We call St a splitting operator on L2 ∩ L1(R, C). For functions of x = (x′, xn) ∈ Rn−1 × R = Rn, we will often treat x′ as a parameter and view f a function of the single variable xn. Most of our translations and convolutions will be in this variable. For f ∈ L2(Rn, C) ∩ L1(Rn, C), let τtf (x) = f (x′, xn − t), f+(x) = f (x)χ{xn≥0}, f−(x) = f (x)χ{xn≤0} and Stf (x) = τtf+ +τ−tf−. The Fourier transform of Stf is e−2πixyStf (x)dx = e−2πityn f+(y) + e2πityn f−(y). dStf (y) =ZRn We study how Ntf = dStfp = (RR dStf (x)p dx)1/p depends on t. The results are often trivial when p = 2, since by the Plancherel theorem, Ntf is independent of t. Throughout the paper, 1 < p < ∞ is fixed. In Sections 2 and 3, we will also assume that p ≥ 2 and is an even integer. Our main result is for real-valued functions in the class (1.1) S = {f ∈ L2 ∩ L1(R) : f+ ∗ f−(x) decreases for x ≥ 0} (1.2) where by Lp(Rn) we mean Lp(Rn, R). We will show in Section 3 that many standard functions satisfy (1.2). For example, non-negative even 1 2 functions that decrease for x > 0, (also called radially-decreasing func- tions or simply bump functions) satisfy (1.2). Our main result is the following. Theorem 1.1. Assume f ∈ S is even and non-negative, and that p is even. Then, Ntf decreases with t. In particular Ntf ≤ N0f . We prove this in Section 3 and show that the hypothesis f ∈ S cannot be removed. It is unclear whether the other assumptions on f are necessary. It is also unclear whether this theorem generalizes to functions in Rn, except in some simple cases (See Corollary 3.4). In Section 2, we prove that Ntf is eventually constant when f has compact support. Theorem 1.2. Suppose h = f + ig ∈ L2(R, C) is supported on [−A, A] and p is even. Then, Nth = Nt0h whenever (p − 2)A (1.3) . t ≥ t0 = 4 The lower bound in (1.3) cannot be improved in any obvious way. For example, let f (x) = χ(−1,1)(x) and p = 4. Then cos(2πxy)dx = sin(2π(1 + t)y) − sin(2πty) πy (1.4) dStf (y) = 2Z 1+t t and 1 2, but not for t < 1 2. (Ntf )4 = Stf4 24(cid:0)6 + (1 − 2t)3 + 1 − 2t3)(cid:1) 4 = which is constant for t ≥ t0 = 1 We offer numerical evidence (but no rigorous proof) that Theorem 1.2 can fail when p = 3. Let f (x) = χ(−1,1)(x). If Theorem 1.2 were true for p = 3, then Ntf would be constant for t ≥ 1 4. Numerical f )3 ∼ 2.6247, (N1f )3 ∼ 2.6124, (N5f )3 ∼ 2.6116 integration gives (N 1 and (N12f )3 ∼ 2.6121. If this data is accurate, it also shows that Ntf is not decreasing, so that Theorem 1.1 also fails when p = 3. 4 In Section 4 we study the effect of the splitting operator St on the norm of a Fourier multiplier. Let m ∈ L∞(Rn, C). Suppose the oper- ator Tmf (x) =RRn e2πixym(y) f (y)dy, initially defined for f ∈ C∞0 (Rn), extends to a bounded operator on Lp(Rn). Then we say that m is a (p, p) Fourier multiplier and we write its operator norm (or multi- plier norm) as TmR If we apply Tm to complex-valued functions, it has a possibly larger operator norm, de- noted mp,p. p,p. We re- late Stmp,p to m+p,p and the norm of the half line multiplier p,p ≤ mp,p ≤ 2mR p,p or simply by mR p,p. Indeed, mR s = χ(0, ∞) (see Theorems 4.1, 4.3 and Proposition 4.5). We also es- timate the (p, p) norm of m+, with exact values in some cases (See Corollary 4.4). 3 p R[0,1]n(cid:12)(cid:12)(cid:12)Pj∈Zn∩N W e2πij·x(cid:12)(cid:12)(cid:12) One of the authors became interested in splitting operators while investigating Lp Lebesgue constants, i.e. the functions L(N, W, p) = dx, where W is a convex set in Rn, p ≥ 1 and N > 0. This problem has a long history (see [L], [NP], [Y]; see also [TB] and the references cited there). These constants depend on N in an explicit way, but their dependence on W is not as clear. It would be interesting to estimate the Lp Lebesgue constants when W = Pk is a regular polygon in R2 with 2k sides, especially with values of p for which these constants are not uniformly bounded in terms of k (see [AD]). To do this, it might be useful to compare the Lp norms of St(Pk) and that of Pk when St divides χPk into two congruent parts. The comparison of the (p, p) norms of χPk and St(χPk) is also very interesting. When 4 3 < p < 4, asymptotic estimates of the (p, p) norm of χPk in terms of k are in [Co1], (see also [Co2], [CS], and see [F] for the multiplier problem for the disk). But we know of no such estimates for other p, or when Pk is not convex - for example, when Pk is the disjoint union of two congruent polygons. See Theorem 4.1 and Proposition 4.5 for results in this direction. Acknowledgement: The authors wish to thank Marshall Ash for pointing out related problems which inspired this work. 2. Proof of Theorem 1.2 and related results Let h be as in Theorem 1.2. We plan to show that (Nth)p = C + P Hj(t), where C is a constant and each Hj is a convolution of p functions with compact support (see formula (2.3)). This implies, using Lemma 2.1 below, that each Hj has compact support, and that Nth = C for large t, as desired. Lemma 2.1. Let φ1 and φ2 be integrable functions supported in [a, b] a φ1(y)φ2(x − y)dy is supported in and [c, d], resp. Then φ1 ∗ φ2(x) =R b [a + c, b + d]. Proof. Suppose x > b + d. If y ≤ b, then x − y > d and φ2(x − y) = 0, so φ1 ∗ φ2(x) = 0. The case x < a + c is similar. (cid:3) 4 Proof of Theorem 1.2. Let h = f + ig ∈ L2(R, C) with support in [−A, A] and let t ≥ t0. Denote the even part of f by fe(x) = 1 2(f (x) + f (−x)) and the odd part by fo = f − fe. Note that if φ is real, then Re φ = bφe. With φ = Stf , we get RedStf (y) = \(Stf )e(y) = \St(fe)(y). Likewise, Im φ = −i φo and Re \St(ig)(y) = −Im cStg(y) = i\St(go)(y). By linearity, Sth(x) = St(fe)(x) + St(fo)(x) + i(St(ge)(x) + St(go)(x)). So, By (1.1), RedSth(y) = \St(fe)(y) + i\St(go)(y). RedSth = e−2πity(cid:16) fe+ + igo+(cid:17) + e2πity(cid:16) fe− + igo−(cid:17) . ImdSth can be written as a similar sum, but with terms containing fo± and ge±. Let h1 through h4 be the functions fe−, fo−, ige−, igo−, which are supported on [−A, 0]. Let h5 through h8 be the functions fe+, fo+, ige+, igo+, which are supported on [0, A]. Squaring both sides of (2.1) and the similar formula for Im dSth, and adding them, (2.1) (2.2) dSth2 =X e2πityCα 8Yj=1 haj j where the sum is over certain α = (a1, . . . a8), such that each ai is a non-negative integer, and P ai = 2. We do not assert that all such α appear exactly once in this sum. Here, Cα = P4 Raising both sides of (2.2) to the power p j=1 aj −P8 2, we get a sum of the form k=5 ak. dSthp =Xβ kβe2πityPβ 8Yj=1 hbj j where β = (b1, . . . b8), each bj is a non-negative integer, P bj = p, and Pβ =P4 When we integrate this, the exponential factors produce inverse k=5 bk, which we will write as p+ − p−. j=1 bj −P8 Fourier transforms, which convert products to convolutions. So, (Nth)p =ZR dSth(y)pdy =Xβ kβHβ(Pβt) (2.3) where Hβ is a convolution of p functions (the ones appearing in Since Pβ = p − 2p−, it is even. Whenever Pβ = 0, the summand in 8Yj=1 hbj j ). 5 (2.3) is constant in t. If Pβ ≥ 2, then 2p− ≤ p − 2 and Pβt ≥ Pβt0 ≥ A(p − 2)/2 ≥ Ap−. By Lemma 2.1, Hβ is supported on [−Ap+, Ap−], so Hβ(Pβt) = 0. For summands with Pβ ≤ −2, we have p − 2 ≥ 2p+, leading to the same conclusion. This proves that (Nth)p is constant for t ≥ t0. (cid:3) Theorem 1.2 has an analogue for Fourier series, with a similar proof, which we omit. Let f = {ck}A k=−A be a finite sequence of complex num- bers. Let f (x) =P+A k=−A cke−2πikx be the corresponding trigonometric polynomial. Let p and t be positive integers, with p even. Define Stf = {bk}k≤A+t, where bk−t = ck for k < 0, bk+t = ck for k > 0, b−t = bt = c0/2 and bk = 0 otherwise. Note that +AXk=1 cke−2πi(k+t)x + −1Xk=−A cke−2πi(k−t)x. (2.4) dStf (x) = c0 cos(2πtx) + Theorem 2.2. For t ≥ t0 = (p−2)A of t. 4 , Ntf = dStfLp[0,1] is independent The split f = f− + f+ defined in the Introduction is a bit arbitrary and can be generalized. Suppose f1 is supported in [−A, b] and h2 is supported in [−b, A], with b ≤ A (so, the supports overlap if b > 0). Let f = f1 + f2 and let Stf (x) = f2(x − t) + f1(x + t) for t ∈ R. Let Ntf = dStfp. Corollary 2.3. With f , Stf and Nt as above, and p even, let t ≥ t0 = p−2 4 (A + b) + b. Then Ntf = Nt0f . Proof. Let t > p−2 4 (A + b) + b. Note that g = Sbf is supported in [−A − b, A + b] and Stg = St+bf . By Theorem 1.2, Ntg is constant if t > p−2 4 (A + b) + b as required. (cid:3) 4 (A + b). So, Ntf = Nt−bg is constant if t > p−2 One can generalize this result a bit more, and similarly Theorem 2.2, with essentially the same proof. Suppose f1 is supported in [−A, b1] and 4 (cid:16)A + b1+b2 2 (cid:17) f2 is supported in [b2, A], with b1,b2 ≤ A. If t≥ t0 = p−2 + p−2 8 (b1 − b2), then Ntf = Nt0f . Theorem 1.2 also has an n-dimensional analogue. 6 4 Corollary 2.4. Assume f (x) ∈ L1 ∩ L2(Rn, C) is supported where xn < A. If p is even, and t ≥ t0 = (p−2)A , then Ntf is independent of t. Proof. Note that bf (y) = Fn−1F1f (y), where Fn−1 denotes the Fourier transform of f (x′, xn) with respect to x′, with xn fixed, and F1 denotes the Fourier transform with respect to xn, with x′ fixed. We will some- times write g as F1g for functions g defined on R. Fix y′ ∈ Rn−1 and t ≥ t0. Define g(xn) = g(y′, xn) = Fn−1f (y′, xn), and g+, g−, Stg as usual. Note that St commutes with Fn−1 because they act on different variables, so F1Stg(yn) = F1Fn−1Stf (y′, yn) =dStf (y). So, ZR dStf (y)p dyn =ZR F1(Stg)(yn)pdyn = (Ntg)p, which is constant for t ≥ t0, by Theorem 1.2. Since St is an isometry on L1 and L2, the Hausdorff-Young inequality implies that dStf ∈ L2 ∩ L∞ ⊂ Lp(Rn). Integrating (2.5) over y′ concludes the proof. (cid:3) In the corollary, we split Rn around the hyperplane H = {x : xn = 0} only for convenience. Since all the norms are invariant under rotations and translations, we could use any other hyperplane in Rn instead. After adjusting the definition of Stf in an obvious manner, an analogue of Corollary 2.4 holds for any such H. (2.5) The next result will be applied to the study of multipliers in Section 4, and involves functions that depend on x and t. Let F (x, t) : Rn × [0, ∞) → C such that F (·, t) ∈ L1 ∩ L2(Rn, C) for every t ≥ 0. We let F+(x, t) = F+(x′, xn, t) = F (x, t)χ{xn>0}(x) and similarly for F−. We let StF (x, t) = StF (x′, ., t) = F+(x′, xn − t, t) + F−(x′, xn + t, t). That is, St acts only on the xn variable. Lemma 2.5. Suppose that F (x, t) as above is supported where xn < A and, for every t ≥ 0, that dStF (y, t) is real-valued. If t ≥ t0 = (p−2)A then ZRn dStF (y, t)pdy =(cid:18)p 2(cid:19)ZRn(cid:16)bF+(y, t)(cid:17) p 2 (cid:16)bF−(y, t)(cid:17) p (2.6) p 4 2 dy. Proof. Fix y′ ∈ Rn−1 and t > t0. Let f (xn) = Fn−1F (y′, xn, t). Observe that St commutes with Fn−1, and so Stf (xn) = Fn−1StF (y′, xn, t). Thus, F1Stf (yn) = dStF (y′, yn, t). Similarly, F1f±(yn) = bF±(y′, yn, t). We argue as in Theorem 1.2. By (1.1) and the binomial theorem, ZR dStF (y′, yn, t)pdyn =ZR F1Stf (yn)pdyn =ZR (e−2πitynF1f+(yn) + e2πitynF1f−(yn))pdyn k(cid:19)ZR (cid:18)p pXk=0 = e2πi(p−2k)tyn (F1f+(yn))k (F1f−(yn))p−kdyn 7 (2.7) p p which is an analog of (2.3). Reasoning as in the proof of Theorem 1.2, the integral on the right hand side of (2.7) simplifies to (f+)∗k ∗ (f−)∗(p−k) ((p− 2k)t) = 0, unless k = p 2 . The only nonzero term in (2.7) is (cid:0) p 2(cid:1)RR (F1f+(yn)) 2 (F1f−(yn)) ZR F1(StF )(y′, yn, t)pdyn =(cid:18)p 2(cid:19)ZR =(cid:18)p 2 (bF−(y′, yn, t)) (bF+(y′, yn, t)) 2 (F1f−(yn)) (F1f+(yn)) 2(cid:19)ZR p Integrating this with respect to y′, (2.6) follows. (cid:3) p 2 dyn. p 2 dyn, and so p p p 2 dyn p 3. On the decreasing function Nt We prove Theorem 1.1 and examine the class S in some detail to show that it is rather large. We show by example that the assumption f ∈ S cannot be removed from the theorem. 3.1. Proof of Theorem 1.1. In this section we will assume that f is even and f ∈ S unless specified otherwise. We need a few lemmas about convolution. Lemma 3.1. τtf+ ∗ τ−tf−(x) is even. Proof. Since f is even, τtf+(−x) = τ−tf−(x), and we have τtf+ ∗ τ−tf−(−x) =ZR =ZR =ZR τ−tf−(−y)τtf+(x + y)dy τ−tf−(y′)τtf+(x − y′)dy′ = τtf+ ∗ τ−tf−(x). (cid:3) τtf+(y)τ−tf−(−x − y)dy 8 Lemma 3.2. Let g, h and F ∈ L2(R). Let U = h ∗ F and V = g ∗ F . Suppose F = f1 ∗ f2 ∗ . . . ∗ fn, where every fi is either f+ or f−. Then (3.1) U ∗ V = h ∗ g ∗ (f+ ∗ f−)∗n where g(x) = g(−x). Proof. Note that f+(x) = f−(x) and V = g ∗ F . Similarly, each fi ∗ fi = f+ ∗ f−, so F ∗ F = (f+ ∗ f−)∗n and (3.1) follows. (cid:3) Lemma 3.3. Assume that u ≥ 0 is supported on (−∞, 0] and that v is decreasing on [0,∞). Then u ∗ v is decreasing on on [0, +∞). Proof. If 0 ≤ x1 < x2, then u ∗ v(x2) − u ∗ v(x1) =Z 0 u(y) (v(x2 − y) − v(x1 − y)) dy ≤ 0. (cid:3) −∞ Proof of Theorem 1.1: Assume that f ∈ S and p = 2m, with m a positive integer. By the Plancherel theorem, (Ntf )2m = dStf2m 2m = (dStf )m2 = (τtf+ + τ−tf−)∗m2 2. (3.2) By a variation of the binomial theorem, (τtf+ + τ−tf−)∗m(x) is a linear combination (with positive coefficients) of terms of the form 2 = (Stf )∗m2 2 + ∗ f∗(m−i) (τtf+)∗i ∗ (τ−tf−)∗(m−i)(x) = f∗i − where 0 ≤ i ≤ m. Let Gi = f∗i weighted sum of integrals of the following form: + ∗ f∗(m−i) (x − (2i − m)t) − . So, (Ntf )p = N 2m t (f ) is a Nijt = ZR = ZR Gi(x − (2i − m)t)Gj(x − (2j − m)t)dx Gi(x − 2(i − j)t)Gj(x)dx = Gj ∗ eGi(2(i − j)t). + ∗ f∗(m−i) We will show that each Nijt decreases with t > 0, and may assume i ≥ j. Let F (x) = f∗j , which is supported on [0,∞). So, g ∗ F (x) = Gi(x). Likewise, Gj(x) = h ∗ F (x) where h = f∗(i−j) is supported on (−∞, 0]. Applying Lemma 3.2 with n = m + j − i, Gj ∗ Gi = h ∗eg ∗ [f+ ∗ f−]∗(m+j−i) = h ∗eg ∗ H. By (1.2) and Lemma 3.1, f+ ∗ f−(x) is radially decreasing, so H(x) is too. By Lemma 3.3 (with u = h and v = H ), h ∗ H(x) decreases and g = f∗(i−j) − − + 9 for x ≥ 0. Applying the lemma again (with u = g), we see that Nijt = g ∗ (h ∗ H)(2(i − j)t) decreases for t ≥ 0. (cid:3). Example. Theorem 1.1 can fail without the assumption that f ∈ S. Let p = 4. The function f (x) = χ{x<1} + χ{10<x<11} is even and non-negative, but f+ ∗ f−(x) is not decreasing, so f 6∈ S. The Fourier transform of Stf (x) = χ{t<x<1+t}(x) + χ{10+t<x<11+t}(x) can be ex- plicitly evaluated. By the residue theorem (or Mathematica software), when 4 < t < 5 we have (Ntf )4 = 8 3 (4t3 − 48t2 + 192t − 247), which is increasing. If p = 6, the same f leads to similar conclusions. The following corollary is a variation of Theorem 1.1 for dimension n > 1. Corollary 3.4. Let f (x) = g(x′)h(xn), where g ∈ L1 ∩ L2(Rn−1) and h ∈ S. If p is even, Ntf decreases with t. Proof. Since Stf (x) = g(x′)Sth(xn), we have Ntf = dStfp = Fn−1gLp(Rn−1)F1(Sth)Lp(R). By Theorem 1.1, F1(Sth)Lp(R) decreases with t. (cid:3) 3.2. On the class S. Proposition 3.5 below shows that S contains the radially decreasing integrable functions, sometimes called bump functions, as well as even functions with two bumps. Roughly speaking, Lemma 3.6 shows that S is closed in L∞. Also, S is closed under the shift operator St, for t > 0, but we leave the proof to the reader. It is not closed under addition. For example, let h(x) = χ{x<1} and let g(x) = χ{10<x<11}. Proposition 3.5 shows that h, g ∈ S, but the previous example shows that f = h + g 6∈ S. Proposition 3.5. Let 0 ≤ r < ∞. Suppose that f ∈ L1 ∩ L∞(R) is even and non-negative. Suppose that f+ is increasing on (0, r], and decreasing on [r,∞). Then f ∈ S. If r = 0, for example, our hypotheses reduce to the single assumption that f+ is decreasing on [0,∞). Any function f as in Proposition 3.5 can be uniformly approximated by an increasing sequence of step functions, {gj}, such as appear in Lemma 3.8, with every gj1 ≤ f1. Then Lemma 3.6 proves Proposition 3.5. 10 Lemma 3.6. Suppose {gj}∞j=1 ∈ S ∩ L∞(R) and {gj} converges uni- formly to some f ∈ L1(R). Suppose that every gj1 ≤ C. Then f ∈ S. Proof. Since f must be bounded, it is also in L2(R). Let 0 ≤ x1 < x2. It is enough to show that f+ ∗ f−(x2) ≤ f+ ∗ f−(x1) + 2ǫ(C + f1) (3.3) for every ǫ > 0. Fix ǫ and choose j such that gj − f∞ < ǫ. Since gj ∈ S, gj+∗gj−(x2) ≤ gj+∗gj−(x1). In general, v∗u∞ ≤ v1u∞, so gj+ ∗ gj− − f+ ∗ f−∞ ≤ gj+ ∗ [gj− − f−]∞ + [gj+ − f+] ∗ f−∞ This proves (3.3). (cid:3) ≤ ǫ(C + f1). Lemma 3.7. Suppose I1 = [b, c] ⊂ [a, d] = I2 and f1(x) = χI1(x) and f2(x) = χI2(−x). Then f1 ∗ f2(x) decreases on [0,∞). The same result holds when I2 ⊂ I1 instead. Proof. f1 ∗ f2(x) =RR f1(y)f2(x − y) dy =R c b χI2(y − x) dy, which is the measure of [b, c] ∩ x + [a, d]. This is c − b for 0 ≤ x ≤ b − a, then linear (and decreasing) in x until x = c − a, when it becomes zero. (cid:3) Lemma 3.8. Suppose Ak are nested intervals with r ∈ A1 ⊂ A2 ⊂ . . . Ak ⊂ [0,∞). Suppose g+(x) = Pk j=1 cjχAj (x), where all cj > 0, and g−(−x) = g+(x). Then g+ ∗ g− decreases on [0,∞). Proof. The convolution splits into k2 pairs, each of which decreases by Lemma 3.7. (cid:3) 4. Split Fourier multipliers This section studies effects of the splitting operator St on the norm of Fourier multipliers (see the Introduction for definitions and nota- tion; for the basic properties of multipliers, see [S]). When p = 2, Plancherel's theorem implies that every bounded function is a Fourier multiplier on L2(Rn, C) and m2,2 = m∞. Since Stm∞ = m∞, we see that Stm2,2 = m2,2. 11 In general, mp,p may be larger than mR p,p. We will write Tm ∈ R to mean Tm maps real-valued functions into real-valued functions. This occurs, for example, when m is even and real-valued. To see this, let f = fe+fo, where fe and fo are the even and odd components. Then Tmf = Tmfe + Tmfo is real. For such m, we have mR p,p = mp,p (see [MZ]). Explicit formulas for multiplier norms are known only in very few cases. The characteristic function σ of any segment [a, b] ⊂ R1 (or the segment multiplier) has the same multiplier norm as the Hilbert transform. It is: np = σp,p = σR p,p = max(cid:26)tan( (4.1) p 2p p 2p ), cot( )(cid:27) 2 maxnsec( p . (see [DL1]) and [P]). The (p, p) norm of the characteristic function of p,p = 1 the half line s(x) = χ(0,∞)(x) is cR (see [E] and also [V]). Also, cp = sp,p = 1 2 = 1√2 (see [HV]). Note that cR 2p ), csc( p 2p) = csc( π p ) p < cp and that cR p = sR 2p ) csc( p 2p )o 2 sec( p 2,2 = 1√2 2,2 ≥ 1. 2,2 6= sR Though the operator norm of a Fourier multiplier on Lp(R, C) is translation invariant, we note that τ−tsR for t > 0. To see this, let I = (−t, t) and let f = χI. Note that f is real-valued and \Tτ−tsf = f . By Plancherel, Tτ−tsf2 = f2, which shows that τ−tsR Even when the norm of a multiplier m is known, the norm of the split multiplier Stm cannot usually be explicitly evaluated. However, the main theorem in this section shows that norm of Stm compares naturally with that of m+ and m−. Theorem 4.1. Let p be even and let m : Rn → C be a (p, p) Fourier multiplier supported in Rn−1 × [−A, A]. Suppose that, for every t > 0, TStm ∈ R. Then, for every t ≥ t0 = (p−2)A , 4 Stmp,p ≤(cid:18)p 2(cid:19) 1 p p (m+p,pm−p,p) 1 2 . (4.2) Remark. If m is real-valued, it is not too difficult to prove that Tm ∈ R if and only if x′ → m(x′, xn) and xn → m(x′, xn) are both even. For such m's, m+p,p = m−p,p and (4.2) reduces to StmR 2(cid:19) 1 p,p = Stmp,p ≤(cid:18)p p p m+p,p. (4.3) 12 2(cid:1) 1 k(cid:1)(cid:1) 1 k=0(cid:0)p p >(cid:0) p Since 2 =(cid:0)Pp p , (4.3) is an improvement over the trivial estimate Stmp,p ≤ 2τtm+p,p = 2m+p,p. Also, observe that mp,p = mp′,p′, and so (4.2) is equivalent to p 2(cid:19) 1 Stmp′,p′ ≤(cid:18)p p p (m+p′,p′m−p′,p′) 1 2 . (4.4) Proof. Let TStm = Tt = T+ + T−, where T+ = Tτtm+ and T− = Tτ−tm−. Since TStm ∈ R, Stmp,p = StmR p,p = supTtfp, where fp = 1. By density, we can fix such an f ∈ C∞0 (Rn), and estimate Ttfp. Then T+f (y) = ZRn = ZRn = e2πitynZRn m+(x′, xn − t) f (x)e2πixydx m+(x)bf (x′, xn + t)e2πi((xn+t)yn+x′y′)dx m+(x)F (f e−2πityn)(x)e2πixydx where F is the Fourier transform on Rn. Define ψ = ψ+ + ψ− by ψ+(x) = m+(x)F (f e−2πityn)(x), and ψ−(x) = m−(x)F (f e2πityn)(x). By (4.5), T+f (y) = e2πityncψ+(−y). A similar calculation shows that T−f (y) = e−2πitynZRn m−(x)F (f e2πityn)(x)e2πixydx (4.5) (4.6) = e−2πityncψ−(−y). By (1.1), Ttf (y) = e2πityncψ+(−y) + e−2πityncψ−(−y) = dStψ(−y). By hypothesis Ttf , and hence dStψ, are real-valued. Since m is supported in Rn−1 × [−A, A], so is ψ(x). Thus, we can apply Lemma 2.5 with F = ψ and by (2.6) and Holder's inequality, p 2(cid:19) 1 p(cid:18)ZRn(cid:16)bψ+(y, t)(cid:17) p dStψp ≡ (cid:18)p 2(cid:19) 1 ≤ (cid:18)p bψ+ p bψ− p . 1 2 1 2 p p 2 (cid:16)bψ−(y, t)(cid:17) p 2 p dy(cid:19) 1 (4.7) But bψ+p = T+fp ≤ τtm+p,p = m+p,p and likewise bψ−p ≤ m−p,p, providing the required estimate on Ttfp = dStψp, and proving (4.2). (cid:3) 13 Our next propositions estimate m+p,p in terms of mp,p and cp. Similar results hold for m−p,p with similar proofs. Proposition 4.2. Let m : Rn → C be a (p, p) Fourier multiplier. Then, (4.8) m+p,p ≤ cpmp,p. If Tm ∈ R we also have If p = 2 and m is real-valued and even with respect to xn, then m+R p,p ≤ cR pmR p,p. (4.9) equality holds in (4.9). Proof. Let H(x) = χ{xn>0}(x) be the half-plane multiplier so that m+ = mH. But mHp,p ≤ mp,pHp,p and (4.8) follows if Hp,p ≤ sp,p = cp. To see this, note that s(xn) f (xn, x′)e2πiyxdx THf (y) =ZRn =ZR s(xn)e2πiynxnF1f (y′, yn)dxn = Tsf (y′,·). p,pf (y′,·)p So,RR THfpdyn ≤ sp Lp(R), and integration over y′ proves that Hp,p ≤ cp. Equality can be proved by setting f = g(x′)h(xn), so that THf (y) = g(y′)Tsh(yn). The proof of (4.9) is similar. To finish the proof, we must show that mR 2,2. The standard proof of the identity m2,2 = m∞ can be easily modified to show that mR 2,2 = m∞ = supTmf2, where the sup is taken over the f ∈ L2(Rn) that are even in xn, with f2 = 1. Fixing such an f , and using the Plancherel theorem, 2,2 ≤ √2m+R Tmf2 = √2Tm+f2 ≤ √2m+R 2 . (cid:3) cp = lim p→∞ cp. Since lim p→1 Our next theorem provides a lower bound for Stmp,p in terms of cp = ∞, inequality (4.10) shows that Stmp,p cannot be bounded above by a constant independent of p. Here, we m(x) exist and that ℓ = assume ℓ+ = lim x→0+ max{ℓ+,ℓ−} > 0. Theorem 4.3. Let m ∈ L∞(R, C) be a (p, p) Fourier multiplier. Then, for every t > 0 and every 1 < p < ∞, m(x) and ℓ− = lim x→0− Stmp,p ≥ ℓcp. (4.10) 14 Remark: Suppose m is supported in [−A, A], TStm ∈ R for every t > 0, p is even and t ≥ t0 = (p−2)A . Then by Theorem 4.3 and Theorem 4.1, 4 2(cid:19) 1 ℓ cp ≤ Stmp,p ≤ cp(cid:18)p p p mp,p. (4.11) Proof. Fix m, t and p. Without loss of generality, ℓ = ℓ+ = lim x→0+ 0. The norm of a multiplier is invariant by translation and dilation, so the (p, p) norm of Stm is the same as that of k(x) = τ−tStm(x) = m−(x + 2t) + m+(x). For λ > 0, let kλ(x) = k(x/λ). Note that kλ(x) = ℓs(x). Fix f ∈ C∞0 (R, C). By the Lebesgue dominated lim λ→∞ convergence theorem, m(x) > Tkλf (y) = lim f (x)kλ(x)e2πiyxdx lim λ→∞ f (x)s(x)e2πiyxdx = ℓTsf (y). λ→∞ZR = ℓZR By Fatou's Lemma λ→+∞ ZR Tkλf (y)pdy ≥ ZR lim inf lim inf λ→+∞ Tkλf (y)pdy = ℓpZR Tsf (y)pdy = ℓp Tsfp p. We obtain ℓcp = ℓ sp,p ≤ kλp,p = Stmp,p. (cid:3) (4.12) The proof of Theorem 4.3 also provides a lower bound (and in some cases an exact value) for the (p, p) norm of m+. Corollary 4.4. Let m : R → C be a (p, p) multiplier, continuous at x = 0, with ℓ = m(0) 6= 0. Then, cpℓ ≤ m+p,p ≤ cpmp,p. Suppose also that m is real-valued and even. Then pmp,p. cR pℓ ≤ m+R p,p ≤ cR (4.14) The four inequalities in (4.13) and (4.14) are equalities if m ∈ L1(R) and m ≥ 0. Remarks. The same bounds hold for m−p,p and m−R p,p. Equality also holds in (4.13) if ℓ = m∞ and p = 2 because m∞ = m2,2. It also holds in (4.14), since Tm ∈ R implies mR 2,2 = m2,2. (4.13) 15 Proof of Corollary 4.4. The upper bounds on the norms of m+ are in Proposition 4.2. The proof of (4.10) establishes the lower bounds. To prove that equality holds in (4.13), assume that m ∈ L1(R) and m ≥ 0 so that m1 =RR m(x)dx = m(0) = ℓ. For every f ∈ C∞0 (Rn), Tmf (x) = f ∗ m(−x), so by Young's inequality, Tmfp ≤ fp m1 = fpℓ. Thus mp,p ≤ ℓ, which implies equality in (4.13); the proof for (4.14) is similar. (cid:3) If m is increasing on R with sup m(x) = 1 and inf m(x) = 0, then mp,p = cp (see [DL2]). Such a multiplier cannot have compact support, but the one in the example below does, again with m+p,p = cp. Example. Let m(x) = (1 − x)χ(−1,1)(x). m(y) = sin2(πy) m+p,p = cp and m+R It is easy to verify that π2y2 ≥ 0, and m∞ = ℓ = 1. Corollary 4.4 shows that p , and mp,p = 1. p,p = cR As a geometric application, we estimate the multiplier norms of "split polygons". Let P be a polygon with diameter 2A in R2, and assume that χP (x1, x2) even in both x1 and x2. It is well known that χP is a (p, p) multiplier for every 1 < p < ∞, with bounds on its multiplier norm that depend on p and the number of sides (see [Co1]). Let P+ = P ∩ {x2 > 0}. Assume that p is even, and np is defined by (4.1). Proposition 4.5. If the intersection of P and the line {x2 = 0} is a segment I, and t ≥ t0 = (p−2)A , then 4 2(cid:19) 1 npcp ≤ St(χP )p,p ≤(cid:18)p p p χP+p,p. (4.15) Proof. Since P is symmetric, St(χP ) ∈ R and the second inequality follows from Theorem 4.1 (see also the Remark following it). For the first, we can assume by dilation that I = (−1, 1). The (p, p) norm of St(χP ) is the same as that of τ−tSt(χP )(x), where the translation acts only on the x2 variable. Let k(x) = τ−tSt(χP )(x). For λ > 0, let kλ(x) = k(x1, λ−1x2). Note that kλ(x) ≡ 0 if −2tλ < x2 < 0, and kλ(x1, 0) ≡ 1 whenever −1 < x1 < 1, and is ≡ 0 otherwise. Thus lim λ→∞ kλ(x1, x2) = s(x2) · σ(x1) (4.16) 16 where σ(x1) = χ(−1,1)(x1). By Lebesgue's Theorem and Fatou's Lemma (as in the proof of Theorem 4.3), we get npcp = s · σp,p ≤ lim inf λ→+∞ kλp,p = kp,p = St(χP )p,p. (cid:3) Example. Let Q be the square with corners at the points (±A, 0) and (0,±A). Assume p is even and t ≥ t0 = (p−2)A . Since χQ+p,p ≤ c3 (see [De]), Proposition 4.5 implies npcp ≤ St(χQ)p,p ≤(cid:18)p 2(cid:19) 1 (4.17) p 4 p c3 p. p References [AD] Ash, M. J. and De Carli, L., Growth of Lp Lebesgue constants for convex polyhedra and other regions. Trans. Amer. Math. Soc. 361, no. 8 (2009) 4215 -- 4232. [Co1] Cordoba, A., The Kakeya maximal function and the spherical summation multiplier. Am. J. Mathematics 99, no. 1, (1974) 1 -- 22. [Co2] Cordoba, A., The multiplier problem for the polygon. Ann. Math. 105, no. [CS] [De] [Do] 3 (1977) 581 -- 588. Carleson, L. and Sjolin, S., Oscillatory integral and a multiplier problem from the disk. Studia Math. 44 (1972), 287 -- 299 De Carli, L., On Fourier multipliers over tube domains. Recent Advances in Harmonic Analysis and Applications (In Honor of Konstantin Oskolkov), Springer Proceedings in Mathematics (2012) 79 -- 92. Dowson, H. R., Spectral Theory of Linear Operators. London Math. Soc. Monogr. vol. 12, Academic Press (1978). [DL1] De Carli, L. and Laeng, E., Truncations of weak- Lp functions and sharp Lp bounds for the segment multiplier. Collect. Math. 51, no. 3 (2000) 309 -- 326. [DL2] De Carli, L. and Laeng, E., On the (p,p) norm of monotonic Fourier mul- [E] [F] [HV] [L] tipliers. C. R. Acad. Sci. Paris Sr. I Math. 330, no. 8 (2000) 657 -- 662. Ess´en M., A superharmonic proof of the M. Riesz conjugate function theo- rem. Ark. Mat. 22 no. 2 (1984) 241 -- 249. Fefferman, C., The multiplier problem for the ball. Ann. of Math. 94, no. 2 (1971) 330-336. Hollenbeck, B., and Verbitsky, I. E., Best Constants for the Riesz Projec- tion. Journal of Functional Analysis 175 (2000) 370 -- 392. Liflyand, E. R.,Lebesgue Constants of multiple Fourier series. Online Jour- nal of Analytic Combinatorics 1, no. 5 (2006) 1 -- 112. [MZ] Marcinkiewicz J., Zygmund A., Quelques in´egalit´es pour les op´erations lin´eaires. J. Marcinkiewicz Collected Papers, edited by A. Zygmund, War- saw (1964) 541 -- 546. 17 [NP] [P] [S] [TB] [V] [Y] Nazarov, F. L., and Podkorytov, A. N., On the behavior of the Lebesgue constants for two-dimensional Fourier sums over polygons. St. Petersburg Math. J. 7 (1996), 663 -- 680. Pichorides, S.K., On the best values of the constants in the theorems of M. Riesz, Zygmund and Kolmogorov. Studia Math. 44 (1972) 165 -- 179. Stein, E. M., Singular Integrals and Differentiability Properties of Func- tions. Princeton Univ. Press, 1970. Trigub, R. M., and E. S. Belinsky,Fourier Analysis and Approximation of Function. Kluwer Academic Publishers, Dordrecht, 2004. Verbitsky I.E., An estimate of the norm of a function in a Hardy space in terms of the norms of its real and imaginary parts. A.M.S. Transl.(2), (1984) 11 -- 15. (Translation of Mat. Issled. Vyp. 54 (1980) 16 -- 20). Yudin, V. A., Behavior of Lebesgue constants. Translated from Matem- aticheskie Zametki, 17 (1975) 401 -- 405. (Translation is in Mathematical Notes no. 17, 233 -- 235.)
1806.09366
2
1806
2019-04-23T06:31:34
The Bishop-Phelps-Bollob\'as property and absolute sums
[ "math.FA" ]
In this paper we study conditions assuring that the Bishop-Phelps-Bollob\'as property (BPBp, for short) is inherited by absolute summands of the range space or of the domain space. Concretely, given a pair (X, Y) of Banach spaces having the BPBp, (a) if Y1 is an absolute summand of Y, then (X, Y1) has the BPBp; (b) if X1 is an absolute summand of X of type 1 or \infty, then (X1, Y) has the BPBp. Besides, analogous results for the BPBp for compact operators and for the density of norm attaining operators are also given. We also show that the Bishop-Phelps-Bollob\'as property for numerical radius is inherited by absolute summands of type 1 or \infty. Moreover, we provide analogous results for numerical radius attaining operators and for the BPBp for numerical radius for compact operators.
math.FA
math
THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS YUN SUNG CHOI, SHELDON DANTAS, MINGU JUNG, AND MIGUEL MART´IN Abstract. In this paper we study conditions assuring that the Bishop-Phelps-Bollob´as property (BPBp, for short) is inherited by absolute summands of the range space or of the domain space. Concretely, given a pair (X, Y ) of Banach spaces having the BPBp, (a) if Y1 is an absolute summand of Y , then (X, Y1) has the BPBp; (b) if X1 is an absolute summand of X of type 1 or ∞, then (X1, Y ) has the BPBp. Besides, analogous results for the BPBp for compact operators and for the density of norm attaining operators are also given. We also show that the Bishop-Phelps-Bollob´as property for numerical radius is inherited by absolute summands of type 1 or ∞. Moreover, we provide analogous results for numerical radius attaining operators and for the BPBp for numerical radius for compact operators. 1. Introduction & Preliminaries 0(x0) = (cid:107)x∗ 0 ∈ X∗ such that x∗ 0(cid:107) for some x0 ∈ SX and (cid:107)x∗ Let X be a Banach space. We denote by BX and SX the unit ball and the unit sphere of X, respectively. We consider the topological dual space of X and we denote it by X∗. We say that x∗ ∈ X∗ attains its norm if there is x0 ∈ SX such that x∗(x0) = (cid:107)x∗(cid:107) = supx∈SX x∗(x). The famous Bishop-Phelps theorem [7] says that given ε > 0 and x∗ ∈ X∗, there exists x∗ 0 − x∗(cid:107) < ε. It is natural to ask if it is true also for bounded linear operators. Given two Banach spaces X and Y , we denote by L(X, Y ) the set of all continuous linear operators. When Y = X, we denote it simply by L(X). We say that T ∈ L(X, Y ) attains its norm when there exists x0 ∈ SX such that (cid:107)T x0(cid:107) = (cid:107)T(cid:107) = supx∈SX (cid:107)T (x)(cid:107). We denote by NA(X, Y ) the set of all norm attaining operators from X to Y . Then, the Bishop-Phelps theorem states that NA(X, K) is dense in X∗ for every Banach space X (where K denotes the base field (= R or C)). Trying to extend the Bishop-Phelps theorem for bounded linear operators, J. Lindenstrauss [27] showed that there are operators which can not be approximated by norm attaining ones. Therefore, in general, there is no version of the Bishop-Phelps theorem for operators. On the other hand, if X is reflexive, then NA(X, Y ) is dense in L(X, Y ) for every Banach space Y (actually, this holds for Banach spaces X with the Radon-Nikod´ym property by a result of J. Bourgain [11]); if Y is a closed subspace of (cid:96)∞ containing the canonical copy of c0, then NA(X, Y ) is dense in L(X, Y ) for every Banach space X. We refer the reader to the survey [1] for a detailed account on norm attaining operators. In 1970, Bollob´as [8] proved a quantitative version of the Bishop-Phelps theorem which turned out to be very useful in numerical range theory. Nowadays, this result is known as the Bishop-Phelps-Bollob´as theorem. It can be enunciated as follows. Let X be a Banach space, let 0 < ε < 2 and suppose that x ∈ BX and x∗ ∈ BX∗ satisfy Re x∗(x) > 1 − ε2 2 . Then, there are y ∈ SX and y∗ ∈ SX∗ such that y∗(y) = 1, (cid:107)y − x(cid:107) < ε, and (cid:107)y∗ − x∗(cid:107) < ε (see [13] for this slightly improved version of the original one [8]). This result motivated M. Acosta, R. Aron, D. Garc´ıa and M. Maestre [2] to introduce in 2008 the following property. Date: June 19th, 2018. 2010 Mathematics Subject Classification. Primary: 46B04; Secondary: 46B20, 46E40, 47A12. Key words and phrases. Bishop-Phelps theorem, Bishop-Phelps-Bollob´as property, norm attaining operators, absolute sums. The first author was supported by Basic Science Research Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Education (NRF-2015R1D1A1A09059788). The second author was supported by the project OPVVV CAAS CZ.02.1.01/0.0/0.0/16 019/0000778, Centrum pokrocil´ych aplikovan´ych pr´ırodn´ıch ved (Center for Advanced Applied Sci- ence) and by Pohang Mathematics Institute (PMI), POSTECH, Korea and Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (NRF-2015R1D1A1A09059788). Last author partially supported by Spanish MINECO/FEDER grant MTM2015-65020-P. 1 9 1 0 2 r p A 3 2 ] . A F h t a m [ 2 v 6 6 3 9 0 . 6 0 8 1 : v i X r a 2 CHOI, DANTAS, JUNG, AND MART´IN Definition 1.1 ([2, Definition 1.1]). A pair (X, Y ) of Banach spaces has the Bishop-Phelps-Bollob´as property (BPBp, for short) if given ε > 0, there exists η(ε) > 0 such that whenever T ∈ L(X, Y ) with (cid:107)T(cid:107) = 1 and x0 ∈ SX are such that (cid:107)T x0(cid:107) > 1 − η(ε), and (cid:107)S − T(cid:107) < ε. there are S ∈ L(X, Y ) with (cid:107)S(cid:107) = 1 and x1 ∈ SX such that (cid:107)x1 − x0(cid:107) < ε, (cid:107)Sx1(cid:107) = 1, In this case, we say that the pair (X, Y ) has the BPBp with the function ε (cid:55)−→ η(ε). If we restrict the operators T and S to be compact in the above definition, then the corresponding property is called the Bishop-Phelps-Bollob´as property for compact operators (BPBp for compact operators, for short) (see [16]). The aim of the authors of [2] was to study the conditions that X and Y must satisfy to get a Bishop-Phelps- Bollob´as type theorem for bounded linear operators. They characterized when the pair ((cid:96)1, Y ) has the BPBp via a geometric property of the Banach space Y which is satisfied by many Banach spaces as C(K), L1(µ), but not for all Banach spaces. They also proved that (X, Y ) has the BPBp when X and Y are finite-dimensional, or when X is arbitrary and Y is a closed subspace of (cid:96)∞ containing the canonical copy of c0. There is a vast literature about this topic and we invite the reader to take a look at the papers cited here and the references therein, as the already cited [2] and [4, 5, 15, 16, 18, 17, 25]. In this paper we are interested in the behavior of the Bishop-Phelps-Bollob´as property and other related properties with respect to absolute sums. Our main motivation is the similar study done in [5] for c0-, (cid:96)1- and (cid:96)∞-sums, which was a very useful technique to produce some important results and examples. Before we continue, let us give the proper terminology and notation. An absolute norm is a norm · a in R2 such that (1, 0)a = (0, 1)a = 1 and (s, t)a = (s,t)a for every s, t ∈ R. Given two Banach spaces W and Z and an absolute norm · a, the absolute sum of W and Z with respect to · a, denoted by W ⊕a Z, is the Banach space W × Z endowed with the norm (cid:107)(w, z)(cid:107)a = ((cid:107)w(cid:107),(cid:107)z(cid:107))a (w ∈ W, z ∈ Z). max{(cid:107)w(cid:107),(cid:107)z(cid:107)} (cid:54) (cid:107)(w, z)(cid:107)a (cid:54) (cid:107)w(cid:107) + (cid:107)z(cid:107) It is immediate to see that (cid:107)(w, 0)(cid:107)a = (cid:107)w(cid:107) for all w ∈ W , so W is isometric to the subspace {(w, 0) : w ∈ W} of W ⊕a Z. It is also easy to show that (1) for every (w, z) ∈ W ⊕a Z and every absolute norm (cid:107) · (cid:107)a. We will say that the Banach space W is an absolute summand of the Banach space X, if there are another Banach space Z and an absolute norm ·a in R2 such that X = W ⊕a Z. This terminology extends the well-known concepts of L-summand and M -summand (see [22]): W is a L-summand of X if there is another Banach space Z such that X = W ⊕1 Z; analogously, if X = W ⊕∞ Z for some Banach space Z, then we say that W is a M -summand of X. For background on absolute norms and absolute sums, we refer the reader to [10, 28, 29, 30, 31, 32]. For a more recent reference, we suggest [21], where the author studies the stability of some geometrical properties of Banach spaces by absolute sums. Examples of absolute sums are the (cid:96)p-sums for 1 (cid:54) p (cid:54) ∞ associated to the (cid:96)p-norms in R2. In his doctoral dissertation [31], R. Pay´a proposed an intuitive classification of absolute norms defined through its behavior at the unit vector (1, 0) of R2 (see also [28, p. 38]). Some of our results depend on this classification, so we include it here. Let us first recall some necessary definitions. For x ∈ X, let D(X, x) be the set of all x∗ ∈ SX∗ such that x∗(x) = (cid:107)x(cid:107), which is convex and nonempty by the Hahn-Banach theorem. We say that x ∈ SX is a vertex of BX if D(X, x) separates the points of X and we say that x is a smooth point of BX if D(X, x) is a singleton subset of X∗. A vertex of the unit ball is an extreme point (see, for example, the remark after [9, Theorem 4.6]). Definition 1.2. Let · a be an absolute norm in R2. We say that · a is of (i) type 1 if the vector (1, 0) is an vertex of B(R2,(cid:107)·(cid:107)a); (ii) type 2 if the vector (1, 0) is a smooth and extreme point of B(R2,(cid:107)·(cid:107)a); (iii) type ∞ if the vector (1, 0) is not extreme point of B(R2,(cid:107)·(cid:107)a). THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 3 The (cid:96)p-norm is of type 1 for p = 1, of type ∞ for p = ∞, and of type 2 for 1 < p < ∞. In subsection 1.1, at the end of this introduction, we will give an account on the results of absolute sums that we will need in this paper. In section 2 we show that if Y1 is an absolute summand of Y and a pair (X, Y ) has the BPBp (resp. BPBp for compact operators), then so does (X, Y1). The analogous result for the density of norm attaining operators and norm attaining compact operators also hold. For domain spaces, we show in section 3 that analogous results hold for type 1 and type ∞ absolute norms: if X1 is an absolute summand of type 1 or ∞ of a Banach space X and a pair (X, Y ) has the BPBp (resp. BPBp for compact operators), then so does (X1, Y ). The corresponding results for the density of norm attaining operators and norm attaining compact operators also hold for type 1 absolute norms. The last section of the paper (§4) is devoted to the study of the Bishop-Phelps-Bollob´as property for numerical radius. Let us recall the relevant notation and terminology about this. Let X be a Banach space and consider the set The numerical radius of an operator T ∈ L(X) is defined by Π(X) := {(x, x∗) ∈ SX × SX∗ : x∗(x) = 1}. v(T ) := sup{x∗(T x) : (x, x∗) ∈ Π(X)}. It is clear that v(T ) (cid:54) (cid:107)T(cid:107) for all T ∈ L(X) and that v(·) is a seminorm in L(X). We say that T ∈ L(X) 0) ∈ Π(X) such that attains its numerical radius (or it is a numerical radius attaining operator) if there is (x0, x∗ x∗ 0(T x0) = v(T ). We denote by NRA(X) the set of all numerical radius attaining operators on X. We refer the reader to the classical books [9, 10] for background on numerical radius of operators and to [1, 12, 32] and the references therein for background on the study of the density of the set of numerical radius attaining operators. Let us give the definition of two properties related to the Bishop-Phelps-Bollob´as property. We take the definitions from [23] although they had appeared earlier for concrete Banach spaces (see [20]). We refer to [3, 6, 20, 23, 26] and references therein for background. Definition 1.3. Let X be a Banach space. We say that (a) X has the BPBp for numerical radius (BPBp-nu, for short) if given ε > 0, there is η(ε) > 0 such that whenever T ∈ L(X) with v(T ) = 1 and (x, x∗) ∈ Π(X) satisfy x∗(T x) > 1 − η(ε), there are S ∈ L(X) with v(S) = 1 and (x0, x∗ 0 − x∗(cid:107) < ε, x∗ 0(Sx0) = 1, (cid:107)x∗ 0) ∈ Π(X) such that (cid:107)x0 − x(cid:107) < ε, and (cid:107)S − T(cid:107) < ε. (b) X has the weak BPBp for numerical radius (weak BPBp-nu, for short) if given ε > 0, there is η(ε) > 0 such that whenever T ∈ L(X) with v(T ) = 1 and (x, x∗) ∈ Π(X) satisfy x∗(T x) > 1 − η(ε), there are S ∈ L(X) and (x0, x∗ x∗ 0(Sx0) = v(S), 0) ∈ Π(X) such that (cid:107)x∗ 0 − x∗(cid:107) < ε, (cid:107)x0 − x(cid:107) < ε, and (cid:107)S − T(cid:107) < ε. Observe that the only difference between the BPBp-nu and the weak BPBp-nu is the normalization of the numerical radius of the operator S given in the first definition. Both properties imply the density of the set of numerical radius attaining operators (see Lemma 4.6). When v(·) is a norm, equivalent to the operator norm (and actually in more situations, see [23, 24]), both properties are equivalent. As far as we know, it is not known whether both properties are always equivalent. Let us also say that both properties have their corresponding versions for compact operators, defined in the obvious way. In section 4 we will show that if X is a Banach space with the BPBp-nu and W is an absolute summand of type 1 or ∞ of X, then W has the BPBp-nu. The analogous result for the weak BPBp-nu, for the BPBp-nu for compact operators, and for the weak BPBp-nu for compact operators also hold. Furthermore, we show that if X is a Banach space such that NRA(X) is dense in L(X) and W is an absolute summand of type 1 or ∞ of X, then NRA(W ) is dense in L(W ). The same result holds for compact operators. 4 CHOI, DANTAS, JUNG, AND MART´IN Let us finally say that some of our results were previously known for the particular case of L-summands and/or M -summands, but other ones are new even in this context. We will highlight in the main part of the paper of which kind is each result. Let us also mention that for L-summands of the domain and for M - summands of the range, there is a formula for the norm of the operators involving the norms of the restrictions or projections (see the proof of [33, Lemma 2], for instance) which makes things easier. This is no longer true for arbitrary absolute summands. 1.1. Some background on absolute sums. Let us recall some known facts on absolute sums which will be relevant in our discussion. Let W , Z be Banach spaces and let (cid:107) · (cid:107)a be an absolute norm. There exists an isometric isomorphism between [W ⊕a Z]∗ and W ∗ ⊕a∗ Z∗, where · a∗ is the dual norm associated to · a, which is also absolute. The action of a functional (w∗, z∗) ∈ W ∗ ⊕a∗ Z∗ at a point (w, z) ∈ W ⊕a Z is given by (cid:104)(w, z), (w∗, z∗)(cid:105) = w∗(w) + z∗(z). We will profusely use the following useful results which were proved in [31]. Lemma 1.4 ([31, Propositions 5.3, 5.5, and 5.6]). Let · a be an absolute norm in R2. Then, (a) · a is of type 1 if and only if there exists K > 0 such that x + Ky (cid:54) (x, y)a for every x, y ∈ R. (b) · a is of type ∞ if and only if there exists b0 > 0 such that (1, b0)a = 1 (so, (1, b)a = 1, ∀b (cid:54) b0). (c) · a is of type 1 if and only · a∗ is of type ∞. (d) · a is of type ∞ if and only · a∗ is of type 1. Finally, we state the following easy result (for its proof see, for example, [19, Lemma 2.2]). Lemma 1.5. Let W and Z be Banach spaces and ⊕a be any absolute sum in R2. If (w, z) ∈ SW⊕aZ and (w∗, z∗) ∈ SW ∗⊕a∗ Z∗ are such that (cid:104)(w, z), (w∗, z∗)(cid:105) = 1, then w∗(w) = (cid:107)w∗(cid:107)(cid:107)w(cid:107) and z∗(z) = (cid:107)z∗(cid:107)(cid:107)z(cid:107). 2. Results on Range Spaces We start this section by showing that the BPBp passes from (X, Y ) to (X, Y1), when Y1 is an absolute summand of Y . This result extends [5, Propositions 2.3 and 2.7], where the results were done for L- and M -summands, and [19, Theorem 2.3], where it was done for the particular case of X = (cid:96)1. Theorem 2.1. Let X, Y be Banach spaces and let Y1 be an absolute summand of Y . If the pair (X, Y ) has the BPBp, then so does (X, Y1). Proof. Given ε ∈ (0, 1), consider η(ε) > 0 to be the BPBp function for the pair (X, Y ) and let Y2 be such that Y = Y1 ⊕a Y2. Let T1 ∈ L(X, Y1) with (cid:107)T1(cid:107) = 1 and x0 ∈ SX be such that x1 ∈ SX such that that (cid:107)T1x0(cid:107) > 1 − η(ε). Define (cid:101)T ∈ L(X, Y ) by (cid:101)T (x) = (T1x, 0) for all x ∈ X. Then (cid:107)(cid:101)T(cid:107) = (cid:107)T(cid:107) = 1 and (cid:107)(cid:101)T x0(cid:107)a = (cid:107)(T1x0, 0)(cid:107)a = (cid:107)T1x0(cid:107) > 1 − η(ε). Since (cid:107)(cid:101)T(cid:107) = 1, x0 ∈ SX and the pair (X, Y ) has the BPBp with η, there are (cid:101)S ∈ L(X, Y ) with (cid:107)(cid:101)S(cid:107) = 1 and (cid:107)x1 − x0(cid:107) < ε and (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. (cid:107)(cid:101)Sx1(cid:107)a = 1, Write (cid:101)S = ((cid:101)S1,(cid:101)S2), where (cid:101)Sj ∈ L(X, Yj) for j = 1, 2. By using (1), for all x ∈ BX , we have Then, (cid:107)(cid:101)S1 − T1(cid:107) < ε and (cid:107)(cid:101)S2(cid:107) < ε. Now we consider y∗ = (y∗ (cid:107)((cid:101)S1x − T1x,(cid:101)S2x)(cid:107)∞ (cid:54) (cid:107)((cid:101)S1x − T1x,(cid:101)S2x)(cid:107)a (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. 1 ⊕a∗ Y ∗ 2) ∈ Y ∗ 1((cid:101)S1x1) + y∗ 2((cid:101)S2x1). 1 = (cid:107)(cid:101)Sx1(cid:107) = y∗((cid:101)Sx1) = y∗ 1, y∗ 2 with (cid:107)y∗(cid:107)a∗ = 1 to be such THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 5 Lemma 1.5 gives that y∗ (cid:107)y∗ we have that y∗ (2) Then, for every x ∈ BX , we have that (cid:107)S1x(cid:107) (cid:54) (cid:107)y∗ S1(x) := (cid:107)y∗ 1((cid:101)S1x1) = (cid:107)y∗ 2((cid:101)S2x1) = (cid:107)y∗ 1(cid:107)(cid:107)(cid:101)S1x1(cid:107) = y∗ 1(cid:107)(cid:107)(cid:101)S1x1(cid:107) and y∗ 2(cid:107)(cid:107)(cid:101)S2x1(cid:107). Since 2((cid:101)S2x1) (cid:62) 1 − (cid:107)(cid:101)S2(cid:107) > 0, 1((cid:101)S1x1) = 1 − y∗ 1 (cid:54)= 0 and (cid:107)(cid:101)S1x1(cid:107) (cid:54)= 0. Define S ∈ L(X, Y1) by (cid:101)S1x1 2((cid:101)S2x) 1(cid:107)(cid:101)S1x + y∗ (cid:107)(cid:101)S1x1(cid:107) (x ∈ X). 2(cid:107)(cid:107)(cid:101)S2x(cid:107) 1(cid:107)(cid:107)(cid:101)S1x(cid:107) + (cid:107)y∗ = (cid:104)((cid:107)(cid:101)S1x(cid:107),(cid:107)(cid:101)S2x(cid:107)), ((cid:107)y∗ (cid:54) ((cid:107)(cid:101)S1x(cid:107),(cid:107)(cid:101)S2x(cid:107))a((cid:107)y∗ 2(cid:107))(cid:105) 1(cid:107),(cid:107)y∗ = (cid:107)((cid:101)S1x,(cid:101)S2x)(cid:107)a(cid:107)(y∗ 2(cid:107))a∗ 1(cid:107),(cid:107)y∗ 2)(cid:107)a∗ (cid:54) (cid:107)(cid:101)S(cid:107)(cid:107)y∗(cid:107)a∗ = 1. = (cid:107)(cid:101)Sx(cid:107)a(cid:107)(y∗ 2)(cid:107)a∗ 1, y∗ 1, y∗ (cid:33) (cid:101)S1x1 1((cid:101)S1x1) + y∗ 2((cid:101)S2x1) 1(cid:107)(cid:101)S1x1 + y∗ (cid:107)(cid:101)S1x1(cid:107) 2((cid:101)S2x1) (cid:54) (cid:107)(cid:101)S2(cid:107) < ε, 1((cid:101)S1x1) = y∗ 1(cid:107)(cid:101)S1x −(cid:101)S1x(cid:107) + (cid:107)(cid:101)S1 − T1(cid:107) + (cid:107)S2(cid:107) < 3ε (cid:107)S1x − T1x(cid:107) (cid:54) (cid:107)(cid:107)y∗ 1 − (cid:107)y∗ 1(cid:107) (cid:54) 1 − y∗ So, (cid:107)S1(cid:107) (cid:54) 1. On the other hand, (cid:107)S1x1(cid:107) (cid:62) y∗ 1(cid:107)y∗ 1(cid:107) (cid:32) (cid:107)y∗ we have for all x ∈ BX that This shows that (cid:107)S1(cid:107) = (cid:107)S1x1(cid:107) = 1. It remains to prove that (cid:107)S1 − T1(cid:107) < ε. Indeed, since = y∗ 2((cid:101)S2x1) = 1. Therefore, (cid:107)S1 − T1(cid:107) < 3ε. Since we already have (cid:107)x1 − x0(cid:107) < ε, we conclude that the pair (X, Y1) has the (cid:3) BPBp as desired. There is a property related to the BPBp for which we may also give an analogous result. A pair (X, Y ) of Banach spaces has the pointwise BPB property (see [18, Definition 1.2] or [17, Definition 1.1]) if given ε > 0, there is η(ε) > 0 such that whenever T ∈ L(X, Y ) with (cid:107)T(cid:107) = 1 and x0 ∈ SX satisfy (cid:107)T x0(cid:107) > 1 − η(ε), there is S ∈ L(X, Y ) such that (cid:107)S(cid:107) = (cid:107)Sx0(cid:107) = 1 and (cid:107)S − T(cid:107) < ε. It is clear that this property is stronger than the BPBp. It was proved in [18] that if (X, Y ) has the pointwise BPB property for some Y , then the space X must be uniformly smooth. The proof of Theorem 2.1 can be obviously adapted to the case of the pointwise BPB property. Therefore, we can state the following result. Proposition 2.2. Let X, Y be Banach spaces and let Y1 be an absolute summand of Y . If the pair (X, Y ) has the pointwise BPB property, then so does (X, Y1). We would like to notice also that, in Theorem 2.1, if one starts with a compact operator T1 : X −→ Y1 and assume that the pair (X, Y ) has the BPBp for compact operators, then the operator (cid:101)T defined in the proof is compact, so we can continue the proof getting a compact operator (cid:101)S and, therefore, the operator S1 : X −→ Y1 defined in (2) is also compact. Thus, we have the following analogous result for this class of operators. This generalizes [16, Lemma 2.6.ii], where the result was enunciated for L- and M -summands. Proposition 2.3. Let X, Y be Banach spaces and let Y1 be an absolute summand of Y . If the pair (X, Y ) has the BPBp for compact operators, then so does (X, Y1). Let D be a bounded closed convex subset of a Banach space X and let Y be another Banach space. Define (cid:107)T(cid:107)D := sup{(cid:107)T x(cid:107) : x ∈ D} 6 CHOI, DANTAS, JUNG, AND MART´IN for every T ∈ L(X, Y ). In [15] the following version of the BPBp was defined: the pair (X, Y ) has the BPBp on D if for every ε > 0, there is ηD(ε) > 0 such that whenever T ∈ L(X, Y ) with (cid:107)T(cid:107)D = 1 and x ∈ D satisfy that (cid:107)T x(cid:107) > 1 − ηD(ε), there are S ∈ L(X, Y ) with (cid:107)S(cid:107)D = 1 and z ∈ D such that (cid:107)Sz(cid:107) = 1, (cid:107)z − x(cid:107) < ε and (cid:107)S − T(cid:107) < ε (the distance (cid:107)S − T(cid:107) being calculated in the usual operator norm). Note that in the proof of Theorem 2.1, we can work with any bounded closed convex subset D of X such that D ⊂ BX instead of BX . So, we also obtain the following result which is an extension of [15, Propositions 4.3 and 4.5], where the result was done for L- and M -summands. Proposition 2.4. Let X, Y be Banach spaces, let D be a bounded closed convex subset of X such that D ⊂ BX and let Y1 be an absolute summand of Y . If the pair (X, Y ) has the BPBp on D, then so does (X, Y1). We next consider analogous results for norm attaining operators. Proposition 2.5. Let X, Y be Banach spaces and let Y1 be an absolute summand of Y . If NA(X, Y ) is dense in L(X, Y ), then NA(X, Y1) is dense in L(X, Y1). Proof. Let ε ∈ (0, 1) and T1 ∈ L(X, Y1) with (cid:107)T1(cid:107) = 1 be given. Let Y2 be such that Y = Y1 ⊕a Y2. Define (cid:101)T ∈ L(X, Y1 ⊕a Y2) by (cid:101)T (x) := (T1x, 0) for all x ∈ X. Then, (cid:107)(cid:101)T(cid:107) = (cid:107)T(cid:107) = 1. Since NA(X, Y1 ⊕a Y2) is dense in L(X, Y1 ⊕a Y2), there are x0 ∈ SX and (cid:101)S ∈ L(X, Y1 ⊕a Y2) with (cid:107)(cid:101)S(cid:107) = 1 such that (cid:107)(cid:101)Sx0(cid:107) = 1 and (cid:107)(cid:101)S −(cid:101)T(cid:107) < ε. Write (cid:101)S = ((cid:101)S1,(cid:101)S2), where (cid:101)Sj : X −→ Yj for j = 1, 2. By (1), we have that (cid:107)(cid:101)S1 − (cid:101)T1(cid:107) < ε and (cid:107)(cid:101)S2(cid:107) < ε. Now, 2 with (cid:107)y∗(cid:107)a∗ = 1 to be such that 1 = (cid:107)(cid:101)Sx0(cid:107) = y∗((cid:101)Sx0). Then, by Lemma 1.5, 2((cid:101)S2x0) = (cid:107)(cid:101)S2x0(cid:107). Since 1((cid:101)S1x0) = (cid:107)(cid:101)S1x0(cid:107) and y∗ 2) ∈ Y ∗ 1 ⊕a∗ Y ∗ 1(cid:107)(cid:107)(cid:101)S1x0(cid:107) = y∗ take y∗ = (y∗ y∗ 1, y∗ (cid:107)y∗ 1((cid:101)S1x0) = 1 − y∗ 1(cid:107)(cid:101)S1x + y∗ 2((cid:101)S2x0) (cid:62) 1 − (cid:107)(cid:101)S2(cid:107) > 1 − ε > 0, 2((cid:101)S2x) (cid:101)S1x1 (cid:107)(cid:101)S1x1(cid:107) (x ∈ X). S1(x) := (cid:107)y∗ we may defined S1 ∈ L(X, Y1) by This operator attains its norm at x0 and it its close to T1 (see the end of the proof of Theorem 2.1). (cid:3) The above result was known for L-summands [5, Proposition 2.9] and for M -summands [33, Lemma 2]. Notice that the proof of Proposition 2.5 also works for compact operators, providing the following result. Proposition 2.6. Let X, Y be Banach spaces and let Y1 be an absolute summand of Y . If NA(X, Y )∩K(X, Y ) is dense in K(X, Y ), then NA(X, Y1) ∩ K(X, Y1) is dense in K(X, Y1). We finish the section with an small discussion about the validity of somehow reciprocal results. It is shown in [5, Proposition 2.4] that if X, Y1, Y2 are Banach spaces and the pairs (X, Y1) and (X, Y2) have the BPBp, then so does the pair (X, Y1 ⊕∞ Y2); the same result holds for the BPBp for compact operators [16, Lemma 3.16] and for the density of norm attaining operators [33, Lemma 2]. We do not know whether any of these reciprocal results is also true for arbitrary absolute sums, even for the case of (cid:96)1-sum. 3. Results on Domain Spaces We first prove that if X1 is an absolute summand of type 1 or ∞, and the pair (X, Y ) has the BPBp, so does (X1, Y ). This extends [5, Proposition 2.6], where the result was shown for L- and M -summands. Theorem 3.1. Let X, Y be Banach spaces and let X1 be an absolute summand of X of type 1 or ∞. If the pair (X, Y ) has the BPBp, then so does (X1, Y ). Proof. Let ε ∈ (0, 1) be given and suppose that the pair (X, Y ) has the BPBp with some function ε (cid:55)−→ η(ε). Let X2 be the Banach space such that X = X1 ⊕a X2. Pick any T ∈ L(X1, Y ) with (cid:107)T(cid:107) = 1 and x1 ∈ SX1 such that (cid:107)T x1(cid:107) > 1 − η(ε). 7 2) ∈ SX such that THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS Define (cid:101)T ∈ L(X, Y ) by (cid:101)T (z1, z2) := T z1 for (z1, z2) ∈ X. Then (cid:107)(cid:101)T(cid:107) = 1. Since (x1, 0) ∈ SX , and the pair (X, Y ) has the BPBp with η, there are (cid:101)S ∈ L(X, Y ) with (cid:107)(cid:101)S(cid:107) = 1 and (x(cid:48) 2) − (x0, 0)(cid:107)a < ε and (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. (cid:107)(cid:101)T (x1, 0)(cid:107) = (cid:107)T x1(cid:107) > 1 − η(ε) 1, x(cid:48) (cid:107)(cid:101)S(x(cid:48) 2)(cid:107) = 1, (cid:107)(x(cid:48) 1, x(cid:48) 1 − x0(cid:107) < ε and (cid:107)x(cid:48) Using (1), we get that (cid:107)x(cid:48) 1, x(cid:48) S(z1) := (cid:101)S(z1, 0) 2(cid:107) < ε. Define S ∈ L(X1, Y ) by (z1 ∈ X1). Then (cid:107)S(cid:107) (cid:54) (cid:107)(cid:101)S(cid:107) = 1 and (cid:107)S − T(cid:107) (cid:54) (cid:107)(cid:101)S −(cid:101)T(cid:107) < ε. To finish the proof, we will prove that S attains its norm at 1 since we already have (cid:107)x(cid:48) x(cid:48) Case 1: Suppose that ⊕a is an absolute norm of type 1. By Lemma 1.4.a, there exists K > 0 such that 1 − x0(cid:107) < ε and (cid:107)S − T(cid:107) < ε. To do so, we divide the proof in two cases. We prove that x(cid:48) (cid:107)x(cid:48) 1, x(cid:48) 1(cid:107),(cid:107)x(cid:48) 2 = 0. Note that for all z2 ∈ BX2, we have 2(cid:107) (cid:54) ((cid:107)x(cid:48) 2(cid:107))a = (cid:107)(x(cid:48) 1(cid:107) + K(cid:107)x(cid:48) 2)(cid:107)a = 1. (cid:107)(cid:101)S(0, z2)(cid:107) = (cid:107)(cid:101)S(0, z2) − (cid:101)T (0, z2)(cid:107) (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. (cid:13)(cid:13)(cid:13)(cid:13)(cid:101)S (cid:18) 2 (cid:54)= 0, we get for all ε ∈ (0, K) that (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) + (cid:107)x(cid:48) 2(cid:107) 0, Therefore, if we assume that x(cid:48) (cid:18) x(cid:48) 1 = (cid:107)(cid:101)S(x(cid:48) 1, x(cid:48) (cid:13)(cid:13)(cid:13)(cid:13)(cid:101)S 1(cid:107) 2)(cid:107) = (cid:107)x(cid:48) 1(cid:107)x(cid:48) 1(cid:107) , 0 2(cid:107) 1(cid:107) + ε(cid:107)x(cid:48) (cid:54) (cid:107)x(cid:48) 1 + K(cid:107)x(cid:48) 2(cid:107) (cid:54) 1 < (cid:107)x(cid:48) 1)(cid:107) = (cid:107)(cid:101)S(x(cid:48) which is a contradiction. Then, (cid:107)S(cid:107) = (cid:107)S(x(cid:48) ε > 0 and consider the vector (x(cid:48) 1(cid:107),(cid:107)ρx(cid:48) 2)(cid:107)a (cid:54) (1, b0)a = 1 and then (x(cid:48) 1, ρx(cid:48) 1, 0)(cid:107) = 1. Case 2: Now assume that ⊕a is an absolute norm of type ∞. By Lemma 1.4.b, there is b0 > 0 such that (1, b0)a = 1. Set ρ = b0 2(cid:107) = 2(cid:107) < b0, we have by the ρ(cid:107)x(cid:48) definition of b0 that (cid:107)(x(cid:48) 2(cid:107) < ρε = b0. Therefore, since (cid:107)(x(cid:48) 2(cid:107) < ε, (cid:107)ρx(cid:48) 1, ρx(cid:48) (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) x(cid:48) 2(cid:107)x(cid:48) 2(cid:107) (x(cid:48) 1, x(cid:48) (x(cid:48) 2) = 1, 0) + 1 − ε b0 2)(cid:107)a = ((cid:107)x(cid:48) (cid:19) (cid:18) (cid:18) (cid:19) 1, 0)(cid:107)a + (cid:19) (cid:18) (cid:107)(cid:101)S(x(cid:48) 1, 0)(cid:107) + 2)(cid:107) (cid:54) 1 − ε 1(cid:107) = (cid:107)(cid:101)S(x(cid:48) b0 1, 0)(cid:107) = 1 = (cid:107)S(cid:107). 1 − ε b0 (cid:107)(x(cid:48) 2)(cid:107)a (cid:54) we get and Then, (cid:107)x(cid:48) 1(cid:107) = (cid:107)(x(cid:48) 1 = (cid:107)(x(cid:48) 1, x(cid:48) 1 = (cid:107)(cid:101)S(x(cid:48) 1, x(cid:48) 1, 0)(cid:107)a = 1 and (cid:107)Sx(cid:48) 2) ∈ X. Note that since (cid:107)x(cid:48) 1, ρx(cid:48) 1(cid:107) (cid:54) 1 and (cid:107)ρx(cid:48) 2(cid:107))a, (cid:107)x(cid:48) 2) ∈ BX . So, writting 1, ρx(cid:48) (x(cid:48) ε b0 1, ρx(cid:48) 2), (cid:107)(x(cid:48) 1, ρx(cid:48) (cid:107)(cid:101)S(x1, ρx(cid:48) ε b0 ε b0 2)(cid:107)a (cid:54) 1 2)(cid:107) (cid:54) 1. (cid:3) We would like to notice that there is a more general result than Theorem 3.1 for the pointwise BPB property (see [17, Proposition 2.1]), which says that if X1 is one-complemented in X and (X, Y ) has the pointwise BPB property, then so does (X1, Y ). We do not know if it is possible to get such a general result for the BPBp. On the other hand, we can easily adapt the proof of Theorem 3.1 in order to get an analogous result for the BPBp for compact operators. This extends [16, Lemma 2.6.i], where the result was enunciated for L- and M -summands. Proposition 3.2. Let X, Y be Banach spaces and let X1 be an absolute summand of X of type 1 or ∞. If the pair (X, Y ) has the BPBp for compact operators, then so does (X1, Y ). Let us present now the version of Theorem 3.1 for norm attaining operators, but in this case we may only deal with type 1 absolute norms. The result was previously known for L-summands (see [33, Lemma 2]). Proposition 3.3. Let X, Y be Banach spaces and let X1 be an absolute summand of X of type 1. If NA(X, Y ) is dense in L(X, Y ), then NA(X1, Y ) is dense in L(X1, Y ). 8 CHOI, DANTAS, JUNG, AND MART´IN Proof. Let ε ∈ (0, 1) and T ∈ L(X1, Y ) with (cid:107)T(cid:107) = 1 be given. Consider X2 to be a Banach space such that X = X1 ⊕a X2. Define (cid:101)T ∈ L(X, Y ) by (cid:101)T (z1, z2) := (T z1, 0) for all (z1, z2) ∈ X1 ⊕a X2. Then (cid:107)(cid:101)T(cid:107) = (cid:107)T(cid:107) = 1. Since NA(X, Y ) is dense in L(X, Y ), there are ((cid:101)x1,(cid:101)x2) ∈ SX1⊕aX2 and (cid:101)S ∈ L(X, Y ) with (cid:107)(cid:101)S(cid:107) = 1 such that (cid:107)(cid:101)S((cid:101)x1,(cid:101)x2)(cid:107) = 1 and (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. Define S ∈ L(X1, Y ) by S(z1) := (cid:101)S(z1, 0) for all z1 ∈ X1. Then, (cid:107)S(cid:107) (cid:54) 1 and for all z1 ∈ SX1, we have (cid:107)Sz1 − T z1(cid:107) = (cid:107)(cid:101)S(z1, 0) − (cid:101)T (z1, 0)(cid:107) (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. (cid:107)(cid:101)S(0, z2)(cid:107) = (cid:107)(cid:101)S(0, z2) − (cid:101)T (0, z2)(cid:107) (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. So, (cid:107)S − T(cid:107) < ε. It remains to prove that S attains its norm. Indeed, first notice that for all z2 ∈ BX2, we have This implies that(cid:101)x1 (cid:54)= 0, otherwise, we would have 1 = (cid:107)(cid:101)S((cid:101)x1,(cid:101)x2)(cid:107) = (cid:107)(cid:101)S(0,(cid:101)x2)(cid:107) < ε, which is a contradiction. Since ⊕a is of type 1, by Lemma 1.4.a, there is K > 0 such that (cid:107)(cid:101)x1(cid:107) + K(cid:107)(cid:101)x2(cid:107) (cid:54) ((cid:107)(cid:101)x1(cid:107),(cid:107)(cid:101)x2(cid:107))a = (cid:107)((cid:101)x1,(cid:101)x2)(cid:107)a = 1. If(cid:101)x2 (cid:54)= 0, we have that, for every ε ∈ (0, K), (cid:13)(cid:13)(cid:13)(cid:13)(cid:101)S (cid:18) 1 = (cid:107)(cid:101)S((cid:101)x1,(cid:101)x2)(cid:107) (cid:54) (cid:107)(cid:101)x1(cid:107) + (cid:107)(cid:101)x2(cid:107) which is a new contradiction. So, (cid:101)x2 = 0 and, then, (cid:107)S(cid:101)x1(cid:107) = (cid:107)(cid:101)S((cid:101)x1, 0)(cid:107) = 1 = (cid:107)(cid:101)x1(cid:107). Therefore, NA(X1, Y ) is (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) < (cid:107)(cid:101)x1(cid:107) + ε(cid:107)(cid:101)x2(cid:107) < (cid:107)(cid:101)x1(cid:107) + K(cid:107)(cid:101)x2(cid:107) (cid:54) 1, 0, (cid:101)x2(cid:107)(cid:101)x2(cid:107) (cid:3) dense in L(X1, Y ). With the same proof, when one restricts it to compact operators, we get the following result. Proposition 3.4. Let X, Y be Banach spaces and let X1 be an absolute summand of type 1 of X. If the set NA(X, Y ) ∩ K(X, Y ) is dense in K(X, Y ), then NA(X1, Y ) ∩ K(X1, Y ) is dense in K(X1, Y ). We do not know if the analogous result of Proposition 3.3 holds true also for absolute norms of type ∞. Actually, we do not know what happens even for M -summands. As in the previous case, we finish the section with an small discussion about the validity of reciprocal results. Let X1, X2, Y be Banach spaces. It is shown in [33, Lemma 2] that if NA(X1, Y ) and NA(X2, Y ) are dense in their respective spaces of operators, then NA(X1 ⊕1 X2, Y ) is dense in L(X1 ⊕1 X2, Y ). The validity of the analogous result for the BPBp is not true: the pair (R, Y ) has the BPBp for every Banach space Y (trivial), while there are Y 's such that (R ⊕1 R, Y ) does not have the BPBp (see [5, Corollary 3.3] for instance). As R ⊕1 R ≡ R ⊕∞ R, the same example shows that the reciprocal result is not true for the (cid:96)∞-sum. We do not know what is the situation for (cid:96)p-sums with 1 < p < ∞. 4. Results for Numerical Radius We would like now to tackle the analogous questions of the previous sections for numerical radius. Our first result in this line is the following one which extends [23, Lemma 19], where the result was proved for L- and M -summands. Theorem 4.1. Let X be a Banach space and let W be an absolute summand of type 1 or ∞ of X. If X has the BPBp-nu, so does W . We will profusely use in this section the following result which is a particular case of [14, Lemma 3.3]. Lemma 4.2. Let W , Z be Banach spaces and let · a be an absolute norm. Given T ∈ L(W ), we define (cid:101)T ∈ L(W ⊕a Z) by (cid:101)T (w, z) := (T w, 0) for every (w, z) ∈ W ⊕a Z. Then v((cid:101)T ) = v(T ) and (cid:107)(cid:101)T(cid:107) = (cid:107)T(cid:107). We are now able to provide the proof of Theorem 4.1. Proof of Theorem 4.1. Let ε ∈ (0, 1) be given and suppose that X has the BPBp-nu with some function η(ε) > 0. Consider Z to be a Banach space with X = W ⊕a Z. We will prove that W satisfies the BPBp-nu with η. Let T ∈ L(W ) with v(T ) = 1 and (w0, w∗ 0) ∈ Π(W ) be such that w∗ 0(T w0) > 1 − η(ε). 9 Since ((w0, 0), (w∗ THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS (cid:104)(cid:101)T (w0, 0), (w∗ 1 )) ∈ Π(W ⊕a Z) such that Consider (cid:101)T ∈ L(W ⊕a Z) to be defined by (cid:101)T (w, z) = (T w, 0) for every (w, z) ∈ W ⊕a Z. Lemma 4.2 gives v((cid:101)T ) = v(T ) = 1. Also, 0, 0)) ∈ Π(W ⊕a Z), v((cid:101)T ) = 1 and W ⊕a Z has the BPBp-nu with η, there are (cid:101)S ∈ L(W ⊕a Z) with v((cid:101)S) = 1 and ((w1, z1), (w∗ 1, z∗ (a) (cid:104)(cid:101)S(w1, z1), (w∗ 1 )(cid:105) = 1, 1, z∗ 1 ) − (w∗ 0, 0)(cid:107)a∗ < ε, (d) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. Write (cid:101)S = ((cid:101)S1,(cid:101)S2), where (cid:101)S1 : W ⊕a Z −→ W and (cid:101)S2 : W ⊕a Z −→ Z. Define S ∈ L(W ) by (b) (cid:107)(w∗ (c) (cid:107)(w1, z1) − (w0, 0)(cid:107)a < ε, and 0(T w0) > 1 − η(ε). 0, 0)(cid:105) = w∗ 1, z∗ S(w) := (cid:101)S1(w, 0) (w ∈ W ). Then, for every (w, w∗) ∈ Π(W ), we have w∗(Sw) = w∗((cid:101)S1(w, 0)) = (cid:104)(cid:101)S(w, 0), (w∗, 0)(cid:105) (cid:54) v((cid:101)S) = 1. (3) So, v(S) (cid:54) 1. Now, since (cid:107)(w∗ 1, z∗ 1 ) − (w∗ (cid:107)w∗ 0, 0)(cid:107)a∗ < ε and (cid:107)(w1, z1) − (w0, 0)(cid:107)a < ε, by using (1), we get that 1 − w∗ 0(cid:107) < ε and (cid:107)w1 − w0(cid:107) < ε. Moreover, for every w ∈ SW , we have (cid:107)Sw − T w(cid:107) = (cid:107)(cid:101)S1(w, 0) − T w(cid:107) (cid:54) max{(cid:107)(cid:101)S1(w, 0) − T w(cid:107),(cid:107)(cid:101)S2(w, 0)(cid:107)} = (cid:107)(cid:101)S(w, 0) − (cid:101)T (w, 0)(cid:107)∞ (cid:54) (cid:107)(cid:101)S(w, 0) − (cid:101)T (w, 0)(cid:107)a < ε. 1(w1) = 1 and w∗ So, (cid:107)S − T(cid:107) < ε. It remains to prove that w∗ cases. Case 1: Suppose that ⊕a is of type ∞. We will prove that z∗ Lemma 1.4.d, ⊕a∗ is of type 1. So, there is K > 0 such that 1(Sw1) = 1. To do so, we divide the proof in two 1 = 0. To do so, suppose that it is not true. By Notice that, since (cid:107)(w1, z1) − (w0, 0)(cid:107)a < ε, we have that (cid:107)z1(cid:107) < ε. So, if ε ∈ (0, K), then (cid:107)w∗ 1(cid:107) + K(cid:107)z∗ 1(cid:107) (cid:54) ((cid:107)w∗ 1(cid:107),(cid:107)z∗ 1(cid:107))a∗ = (cid:107)(w∗ 1, z∗ 1 )(cid:107)a∗ = 1. which is a contradiction. So, z∗ 1 = w∗ 1(w1) + z∗ 1 (z1) (cid:54) (cid:107)w∗ < (cid:107)w∗ 1(cid:107)(cid:107)w1(cid:107) + (cid:107)z∗ 1(cid:107) + ε(cid:107)z∗ 1 = 0. This implies that (w1, w∗ 1(cid:107)(cid:107)z1(cid:107) 1(cid:107) < (cid:107)w∗ 1) ∈ Π(W ). 1(cid:107) + K(cid:107)z∗ 1(cid:107) (cid:54) 1, Now, by Lemma 1.4.b, there is b0 > 0 such that (1, b0)a = 1. Put ρ = b0 ε > 0. Then, (cid:107)ρz1(cid:107) = ρ(cid:107)z1(cid:107) < ρε = which implies that (cid:107)(w1, ρz1)(cid:107)a = 1. So, ((w1, ρz1), (w∗ 1 = (cid:107)(w1, z1)(cid:107)a (cid:54) 1 − ε b0 (cid:107)(w1, ρz1)(cid:107)a (cid:54) 1, ε b0 (cid:107)(w1, 0)(cid:107)a + 1, 0)) ∈ Π(W ⊕a Z) and 1, 0)(cid:105) (cid:54) v((cid:101)S) = 1. (cid:104)(cid:101)S(w1, ρz1), (w∗ b0, so Writing we get that (cid:107)(w1, ρz1)(cid:107)a = ((cid:107)w1(cid:107),(cid:107)ρz1(cid:107))a (cid:54) (1, b0)a = 1. (w1, z1) = 1 − ε b0 (w1, 0) + ε b0 (w1, ρz1) (cid:18) (cid:18) (cid:19) (cid:19) 10 Therefore, CHOI, DANTAS, JUNG, AND MART´IN (cid:18)(cid:18) 1 = (cid:104)(cid:101)S(w1, z1), (w∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:19) 1, 0)(cid:105) (cid:18) (cid:104)(cid:101)S(w1, 0), (w∗ 1 − ε b0 (cid:19) (w1, 0) + (cid:54) = ε b0 (w1, ρz1) (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) (cid:104)(cid:101)S(w1, ρz1), (w∗ , (w∗ 1, 0) 1 − ε b0 1(Sw1) = (cid:104)(cid:101)S(w1, 0), (w∗ 1, 0)(cid:105) + 1, 0)(cid:105) (cid:54) 1. 1, 0)(cid:105) = 1. By (3), we get that v(S) = w∗ ε b0 This shows that w∗ Case 2: Now suppose that ⊕a is of type 1. Since (cid:107)(x∗ prove that z1 = 0. Suppose not. By Lemma 1.4.a, there is K > 0 such that 1(Sw1) = 1. 0) − (x, 0)(cid:107)a∗ < ε, we have that (cid:107)y∗ 0, y∗ 0(cid:107) < ε. We will (cid:107)w1(cid:107) + K(cid:107)z1(cid:107) (cid:54) ((cid:107)w1(cid:107),(cid:107)z1(cid:107))a = (cid:107)(w1, z1)(cid:107)a = 1. Therefore, if ε ∈ (0, K), then 1 = w∗ 1(w1) + z∗ 1 (z1) (cid:54) (cid:107)w∗ 1(cid:107)(cid:107)w1(cid:107) + (cid:107)z∗ 1(cid:107)(cid:107)z1(cid:107) < (cid:107)w1(cid:107) + ε(cid:107)z1(cid:107) < (cid:107)w1(cid:107) + K(cid:107)z1(cid:107) (cid:54) 1, 1) ∈ Π(W ). which is a contradiction. So, z1 = 0 and (w1, w∗ By Lemma 1.4.c, ⊕a∗ is of type ∞. So, there is b0 > 0 such that (1, b0)a∗ = 1. Put ρ = b0 1(cid:107) = ρ(cid:107)z∗ 1(cid:107) < ρε = b0. So, (cid:107)ρz∗ ε > 0. Then, (w∗ (cid:18) (cid:107)(w∗ 1, ρz∗ 1, ρz∗ 1 )(cid:107)a∗ = ((cid:107)w∗ 1(cid:107))a∗ (cid:54) (1, b0)a∗ = 1. 1(cid:107),(cid:107)ρz∗ (cid:19) 1 − ε 1, z∗ 1 ) = b0 1 )(cid:105) (cid:54) v((cid:101)S) = 1 (cid:104)(cid:101)S(w1, 0), (w∗ 1 )(cid:107)a∗ = 1 and then ((w1, 0), (w∗ 1 )) ∈ Π(W ⊕a Z). This implies that (cid:19) (cid:18) 1, 0)((cid:101)S(w1, 0)) + 1 = (cid:104)(cid:101)S(w1, 0), (w∗ 1, z∗ 1(Sw1) = (cid:104)(cid:101)S(w1, 0), (w∗ 1, 0)(cid:105) = 1 and this finishes the proof. 1 − ε b0 1, ρz∗ 1 ). 1 )(cid:105) (cid:54) 1, ρz∗ 1, ρz∗ 1, ρz∗ 1, 0) + (w∗ (w∗ (w∗ (w∗ ε b0 ε b0 Write So, (cid:107)(w∗ and So, w∗ 1 )((cid:101)S(w1, 0)) (cid:54) 1. (cid:3) We would like to point out that the same proof of Theorem 4.1 works also for compact operators and we have the analogous result for the BPBp-nu for compact operators. Proposition 4.3. Let X be a Banach space and let W be an absolute summand of type 1 or ∞ of X. If X has the BPBp-nu for compact operators, so does W . Arguing as in Theorem 4.1, we get the following result for the weak BPBp-nu. Recall that the only difference between the BPBp-nu and the weak BPBp-nu is the normalization of the numerical radius of the operator by its numerical radius (see Definition 1.3). We include an sketch of the proof for completeness. Proposition 4.4. Let X be a Banach space and let W be an absolute summand of type 1 or ∞ of X. If X has the weak BPBp-nu, so does W . Proof. Let ε ∈ (0, 1) be given and consider the weak BPBp-nu function η(ε) > 0 for the Banach space X = 0(T w0) > 1 − η(ε). Using the W ⊕a Z. Let T ∈ L(W ) with v(T ) = 1 and (w0, w∗ 1 )) ∈ 0(cid:107) < ε and (cid:107)w1 − w0(cid:107) < ε. Writing again (cid:101)S = ((cid:101)S1,(cid:101)S2), we define S ∈ L(W ) by 0, 0)(cid:107)a∗ < ε, (cid:107)(w1, z1) − (w0, 0)(cid:107)a < ε, and 1) ∈ Π(W ). ε > 0. Since (cid:107)z1(cid:107) < ε, we have that same notation of Theorem 4.1 and applying our hypothesis, there are (cid:101)S ∈ L(W ⊕a Z) and ((w1, z1), (w∗ Π(W ⊕a Z) such that (cid:104)(cid:101)S(w1, z1), (w∗ (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. Then, (cid:107)w∗ 1 ) − (w∗ S(w) := (cid:101)S1(w, 0) for every w ∈ W . Then, v(S) (cid:54) v((cid:101)S) and (cid:107)S − T(cid:107) < ε. Suppose that ⊕a is of type ∞. As in Theorem 4.1 (Case 1), we have that z∗ Since ⊕a is of type ∞, there is b0 > 0 such that (1, b0)a = 1. Put ρ = b0 1 )(cid:105) = v((cid:101)S), (cid:107)(w∗ 0) ∈ Π(W ) be such that w∗ 1 = 0 and then (w1, w∗ 1 − w∗ 1, z∗ 1, z∗ 1, z∗ 11 1, 0)(cid:105) (cid:54) (cid:107)(w1, ρz1)(cid:107)a (cid:54) (1, b0)a = 1. Then, ((w1, ρz1), (w∗ v((cid:101)S). Therefore, ε b0 ε b0 w∗ = (cid:54) = (w1, ρz1) , (w∗ 1, 0) (w1, 0) + (cid:19) (cid:19) (cid:18)(cid:18) 1 − ε b0 1 − ε b0 1 − ε b0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) (cid:18) THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 1, 0)(cid:105) + (cid:104)(cid:101)S(w1, ρz1), (w∗ ε b0 (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) (cid:104)(cid:101)S(w1, ρz1), (w∗ 1, 0)(cid:105) 1, 0)(cid:105) (cid:54) v((cid:101)S). 1, 0)) ∈ Π(W ⊕a Z). This implies that (cid:104)(cid:101)S(w1, ρz1), (w∗ v((cid:101)S) = (cid:104)(cid:101)S(w1, z1), (w∗ (cid:19) 1, 0)(cid:105) (cid:104)(cid:101)S(w1, 0), (w∗ 1(Sw1) + 1(Sw1) = v((cid:101)S) (cid:62) v(S). So, w∗ 1 )(cid:105) (cid:54) v((cid:101)S). Therefore, v((cid:101)S) = (cid:104)(cid:101)S(w1, 0), (w∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S(w1, 0), (cid:19) (cid:18) 1 )(cid:105) 1, z∗ (cid:18) (cid:19) 1 − ε (cid:104)(cid:101)S(w1, 0), (w∗ b0 (cid:19) (cid:18) 1(Sw1) + 1(Sw1) = v((cid:101)S) (cid:62) v(S). So, w∗ (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) 1, ρz∗ 1, 0) + 1 ) (cid:104)(cid:101)S(w1, 0), (w∗ 1, 0)(cid:105) + 1 )(cid:105) 1 )(cid:105) (cid:54) v((cid:101)S). (cid:104)(cid:101)S(w1, 0), (w∗ ε b0 1(Sw1) = v(S) and W has the weak BPBp-nu. 1 − ε b0 1 − ε b0 ε b0 ε b0 1, ρz∗ 1, ρz∗ = (cid:54) (w∗ (w∗ = w∗ This shows that w∗ 1(Sw1) = v(S) and W has the weak BPBp-nu. ⊕a∗ is of type ∞ (see Lemma 1.4.(c)), there is b0 > 0 such that (1, b0)a∗ = 1. Put ρ = b0 (cid:107)z∗ Now suppose that ⊕a is of type 1. As in Theorem 4.1 (Case 2), z1 = 0 and then (w1, w∗ that (cid:104)(cid:101)S(w1, 0), (w∗ 1(cid:107) < ε, we have that (cid:107)(w∗ 1 )(cid:107)a∗ (cid:54) (1, b0)a∗ = 1. Then ((w1, 0), (w∗ 1) ∈ Π(W ). Since ε > 0. Since 1 )) ∈ Π(W ⊕a Z). This implies 1, ρz∗ 1, ρz∗ 1, ρz∗ This shows that w∗ (cid:3) Again, the above proof can be adapted to compact operators to get the following result. Proposition 4.5. Let X be a Banach space and let W be an absolute summand of type 1 or ∞ of X. If X has the weak BPBp-nu for compact operators, so does W . Next, we have interest to investigate the density of the set of numerical radius attaining operators. To do so, we will prove the following easy lemma which says that, in order to prove the denseness of the set NRA(X) for a Banach space X, it is enough to consider operators with numerical radius one. Observe that T ∈ L(X) attains its numerical radius if and only if the operator λT does for every λ ∈ R. Lemma 4.6. Let X be a Banach space. The following statements are equivalents. (a) The set NRA(X) is dense in L(X). (b) For every T ∈ L(X) with v(T ) = 1, there is a sequence {Sn} ⊂ NRA(X) with v(Sn) = 1 for every (c) For every T ∈ L(X) with v(T ) = 1 and every ε > 0, there is S ∈ L(X) with v(S) = 1 such that n ∈ N and such that Sn −→ T in norm. (cid:107)S − T(cid:107) < ε. Proof. (a) ⇒ (b). Let T ∈ L(X) with v(T ) = 1. By hypothesis, there is a sequence {S(cid:48) n −→ T in norm. This implies that there is n0 ∈ N such that v(S(cid:48) n) − v(T ) (cid:54) (cid:107)S(cid:48) S(cid:48) Since v(T ) = 1, we have that v(S(cid:48) n) > 0 for all n (cid:62) n0. Consider then the sequence (Sn) defined by n} ⊂ NRA(X) such that n − T(cid:107) < 1 for all n (cid:62) n0. ∈ L(X) Then {Sn} ⊂ NRA(X), v(Sn) = 1 for all n ∈ N and, since v(S(cid:48) (b). (b) ⇔ (c) is immediate. 1 v(S(cid:48) Sn := S(cid:48) n+n0 n+n0 ) (n ∈ N). n) −→ v(T ) = 1, Sn −→ T in norm. This proves 12 CHOI, DANTAS, JUNG, AND MART´IN (b) ⇒ (a). Let T ∈ L(X) be given. If v(T ) = 0, then T attains its numerical radius and we are done. Otherwise, v(T ) (cid:54)= 0 and we may consider the operator T (cid:48) := T v(T ) which satisfies that v(T (cid:48)) = 1. By hypothesis, there is a sequence {Sn} ⊂ NRA(X) with v(Sn) = 1 for all n ∈ N and such that Sn −→ T (cid:48) in norm. This implies that v(T )Sn −→ v(T )T (cid:48) = T and we are done since v(T )Sn ∈ NRA(X) for every n ∈ N. (cid:3) Now we are ready to provide a result for the denseness of the operators which attain their numerical radius. Proposition 4.7. Let X be a Banach space and let W be an absolute summand of X of type 1 or ∞. If NRA(X) is dense in L(X), then NRA(W ) is dense in L(W ). Proof. Let Z be a Banach space such that X = W ⊕a Z. Let ε ∈ (0, 1) and T ∈ L(W ) be given. We may consider v(T ) = 1 by using Lemma 4.6. Define (cid:101)T ∈ L(W ⊕a Z) by (cid:101)T (w, z) := (T w, 0) for every (w, z) ∈ W ⊕a Z. By Lemma 4.2, v((cid:101)T ) = v(T ) = 1. Since NRA(W ⊕a Z) is dense in L(W ⊕a Z), there are (cid:101)S ∈ L(W ⊕a Z) with v((cid:101)S) = 1 and ((w0, z0), (w∗ Set (cid:101)S = ((cid:101)S1,(cid:101)S2), where (cid:101)S1 : W ⊕a Z −→ W and (cid:101)S2 : W ⊕a Z −→ Z. By Lemma 1.5, we have (cid:104)(cid:101)S(w0, z0), (w∗ 0 )) ∈ Π(W ⊕a Z) such that 0, z∗ 0(w0) = (cid:107)w∗ w∗ 0 )(cid:105) = v((cid:101)S) = 1 0(cid:107)(cid:107)w0(cid:107) and (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. 0(cid:107)(cid:107)z0(cid:107). 0 (z0) = (cid:107)z∗ and z∗ 0, z∗ (4) Now for all (w, z) ∈ BW⊕aZ, we have that max{(cid:107)(cid:101)S1(w, z) − T w(cid:107),(cid:107)(cid:101)S2(w, z)(cid:107)} = (cid:107)((cid:101)S1(w, z) − T w,(cid:101)S2(w, z))(cid:107)∞ = (cid:107)(((cid:101)S1(w, z),(cid:101)S2(w, z)) − (T w, 0))(cid:107)∞ = (cid:107)(cid:101)S(w, z) − (cid:101)T (w, z)(cid:107)∞ (cid:54) (cid:107)(cid:101)S(w, z) − (cid:101)T (w, z)(cid:107)a (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. (cid:107)(cid:101)S1(w, 0) − T w(cid:107) < ε and (cid:107)(cid:101)S2(cid:107) < ε. This implies that, for all w ∈ BW , (5) On the other hand, for all z ∈ BZ, we get that (6) In particular, (cid:107)(cid:101)S1(0, z)(cid:107) < ε and (cid:107)(cid:101)S2(0, z)(cid:107) < ε for all z ∈ BZ. (cid:107)(cid:101)S(0, z)(cid:107)a = (cid:107)(cid:101)S(0, z) − (cid:101)T (0, z)(cid:107)a (cid:54) (cid:107)(cid:101)S − (cid:101)T(cid:107) < ε. 0 )(cid:105) (cid:54) (cid:107)(cid:101)S(0, z0)(cid:107)a < ε, 1 = (cid:104)(cid:101)S(0, z0), (w∗ Claim: w0 (cid:54)= 0. Otherwise, by using (6), we would have 0, z∗ 0 ((cid:101)S2(w0, z0)) (cid:54) (cid:107)(cid:101)S2(cid:107) < ε, 1 = (cid:104)(cid:101)S(w0, z0), (0, z∗ 0 (cid:54)= 0. Otherwise, by using (5), we would have 0 )(cid:105) = z∗ (cid:19)(cid:19) (cid:19) (cid:18)(cid:18) w0(cid:107)w0(cid:107) , 0 which is a contradiction. Claim: w∗ 0 (cid:54)= 0, by using (4), we have that Therefore, since w0, w∗ which is a new contradiction. ∈ Π(W ⊕a Z). (7) Define the operator S ∈ L(W ) by (cid:18) w∗ 0(cid:107)w∗ 0(cid:107) , 0 S(w) := (cid:101)S1(w, 0) , (w ∈ W ). Note that for all (w, w∗) ∈ Π(W ), we have that ((w, 0), (w∗, 0)) ∈ Π(W ⊕a Z) and then, w∗(Sw) = w∗((cid:101)S1(w, 0)) = (cid:104)(cid:101)S(w, 0), (w∗, 0)(cid:105) (cid:54) v((cid:101)S) = 1. This shows that v(S) (cid:54) 1. Also, by using (5), note that for all w ∈ SW , (cid:107)Sw − T w(cid:107) = (cid:107)(cid:101)S1(w, 0) − T (w)(cid:107) < ε. THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 13 So, (cid:107)S − T(cid:107) < ε. It remains to prove that S attains its numerical radius, and we do this separating the proof in two cases. Case 1: Assume first that ⊕a is of type 1. We will prove that z0 = 0. Suppose not. Since ⊕a is of type 1, by Lemma 1.4.a there is K > 0 such that (cid:107)w0(cid:107) + K(cid:107)z0(cid:107) (cid:54) (cid:107)(w0, z0)(cid:107)a = 1. On the other hand, being ⊕a∗ of type ∞ (see Lemma 1.4.c), there is b0 > 0 such that (1, b0)a∗ = 1 and then , 0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) w∗ (cid:28)(cid:18) w0(cid:107)w0(cid:107) , 0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)a∗ 0(cid:107) , b0z∗ 0(cid:107)w∗ (cid:18) w∗ (cid:19) (cid:18)(cid:18) w0(cid:107)w0(cid:107) , 0 (cid:18) w∗ 0(cid:107) , b0z∗ 0(cid:107)w∗ (cid:19) (cid:19) and 0 , (cid:54) (1, b0)a∗ = 1. 0 = 1. (cid:19)(cid:29) 0(cid:107) , b0z∗ 0(cid:107)w∗ (cid:19)(cid:19) (cid:18) w∗ (cid:19) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:54) v((cid:101)S) = 1. 0(cid:107) , b0z∗ 0(cid:107)w∗ 0 , ∈ Π(W ⊕a Z). Using (4), we have that Then, (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) w∗ 0(cid:107) , b0z∗ 0(cid:107)w∗ 0 This implies that (8) Now, set (w∗ 0, z∗ So, we have that (cid:104)(cid:101)S(w0, 0), (w∗ 0, z∗ 0 )(cid:105) = (cid:54) By using (7), we get that (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)a∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S = 1 (cid:18) w0(cid:107)w0(cid:107) , 0 (cid:18) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S(w0, 0), (cid:18) 1 − 0 ) = 1 − 1 b0(cid:107)w∗ 0(cid:107) (cid:18) 1 − 1 b0(cid:107)w∗ 0(cid:107) (cid:19) (cid:104)(cid:101)S(w0, 0), (w∗ 0, 0)(cid:105) = (cid:54) Now, using (8), 0(cid:107)(cid:104)(cid:101)S(w0, 0), (w∗ 1 b0(cid:107)w∗ 0, b0(cid:107)w∗ 0(cid:107)z∗ 0 )(cid:105) = 1 b0(cid:107)w∗ 0(cid:107) (w∗ 0, 0) + 1 b0(cid:107)w∗ 0(cid:107) (w∗ 0, b0(cid:107)w∗ 0(cid:107)z∗ 0 ). 0(cid:107)z∗ 0 ) 0, b0(cid:107)w∗ (w∗ 1 − 1 − (cid:107)w∗ 0, b0(cid:107)w∗ 1 b0(cid:107)w∗ 0(cid:107) 1 b0(cid:107)w∗ 1 b0(cid:107)w∗ (cid:18) (cid:19) (cid:19) 0, 0) + (cid:104)(cid:101)S(w0, 0), (w∗ 0, 0)(cid:105) + (cid:19) (cid:18) (cid:18) (cid:19) (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) 0(cid:107) (w∗ 0(cid:107)(cid:104)(cid:101)S(w0, 0), (w∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) w0(cid:107)w0(cid:107) , 0 (cid:19) (cid:18) w∗ (cid:18) (cid:19) 0(cid:107)(cid:107)w0(cid:107) (cid:107)w0(cid:107)v((cid:101)S) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) w0(cid:107)w0(cid:107) , 0 (cid:18) w∗ (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) w0(cid:107)w0(cid:107) , 0 (cid:18) w∗ (cid:19) 0(cid:107)(cid:107)w∗ 0(cid:107)(cid:107)w∗ 0(cid:107) v((cid:101)S)(cid:107)w0(cid:107) = 1 b0(cid:107)w∗ 1 b0(cid:107)w∗ 1 b0(cid:107)w∗ 1 b0(cid:107)w∗ 0(cid:107) 1 b0(cid:107)w∗ 0(cid:107) 0(cid:107)(cid:107)w0(cid:107) 0(cid:107)(cid:107)w0(cid:107) 0(cid:107)(cid:107)w0(cid:107) 1 b0(cid:107)w∗ 0(cid:107) 1 b0(cid:107)w∗ 0(cid:107)w∗ (cid:107)w0(cid:107). 1 − 1 − (cid:54) = b0(cid:107)w∗ 0(cid:107)z∗ 0(cid:107)w∗ (cid:107)w∗ 0(cid:107) , 0(cid:107) 0(cid:107) , b0z∗ 0(cid:107)w∗ (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) 0(cid:107) , 0 0 0 , , , 0(cid:107)z∗ 0 )(cid:105) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18) Then, (cid:104)(cid:101)S(w0, 0), (w∗ 0 )(cid:105) (cid:54) 0, z∗ (cid:104)(cid:101)S(0, z0), (w∗ On the other hand, by using (6), we have that 0, z∗ (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) (cid:13)(cid:13)(cid:13)(cid:13)(cid:101)S (cid:18) 1 − 1 0(cid:107) b0(cid:107)w∗ (cid:107)w0(cid:107) + 1 b0(cid:107)w∗ 0(cid:107)(cid:107)w0(cid:107) = (cid:107)w0(cid:107). 0 )(cid:105) = (cid:107)z0(cid:107) (cid:54) (cid:107)z0(cid:107) (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) < ε(cid:107)z0(cid:107). 0, z∗ 0 ) , (w∗ 0, z0(cid:107)z0(cid:107) z0(cid:107)z0(cid:107) 0, 14 CHOI, DANTAS, JUNG, AND MART´IN Using these inequalities, for all ε ∈ (0, K), we get that 1 = (cid:104)(cid:101)S(w0, z0), (w∗ (cid:54) (cid:104)(cid:101)S(w0, 0), (w∗ 0, z∗ 0, z∗ 0 )(cid:105) + (cid:104)(cid:101)S(0, z0), (w∗ 0 )(cid:105) 0, z∗ 0 )(cid:105) < (cid:107)w0(cid:107) + ε(cid:107)z0(cid:107) < (cid:107)w0(cid:107) + K(cid:107)z0(cid:107) (cid:54) 1, which is a contradiction. So, z0 = 0. Being z0 = 0, we have that (w0, w∗ it is attained. Indeed, since (w0, w∗ Then, 1 = (cid:104)(cid:101)S(w0, 0), (w∗ 0, z∗ 0 )(cid:105) = 1. Let us show that v(S) = 1 and that 0, z∗ 0 )(cid:105) = (cid:54) 0 )(cid:105) (cid:54) v((cid:101)S) (cid:54) 1. 0) ∈ Π(W ) and (cid:104)(cid:101)S(w0, 0), (w∗ 0) ∈ Π(W ), we have that (cid:104)(cid:101)S(w0, 0), (w∗ 0, b0z∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:18) (cid:19) (cid:104)(cid:101)S(w0, 0), (w∗ (cid:18) (cid:19) 0, 0)(cid:105) + (cid:104)(cid:101)S(w0, 0), (w∗ (cid:18) (cid:19) 0, 0)(cid:105) + 0((cid:101)S1(w0, 0)) + (cid:19) (cid:18) 1 (cid:104)(cid:101)S(w0, 0), (w∗ b0 (cid:19) (cid:18) 0(Sw0) + v((cid:101)S) (cid:54) 1. 1 − 1 b0 1 − 1 b0 1 − 1 b0 1 − 1 b0 1 − 1 b0 v(S) + w∗ w∗ 1 b0 1 b0 = = (cid:12)(cid:12)(cid:12)(cid:12) 1 b0 1 b0 (cid:104)(cid:101)S(w0, 0), (w∗ 0 )(cid:105) 0, b0z∗ (cid:104)(cid:101)S(w0, 0), (w∗ 0 )(cid:105) 0, b0z∗ (cid:104)(cid:101)S(w0, 0), (w∗ 0 )(cid:105) 0, b0z∗ 0 )(cid:105) 0, b0z∗ 0 = 0. Suppose not. By Lemma 1.4.b, there is (cid:54) 0(Sw0) = v(S) = 1. By (7), we have that This implies that w∗ Case 2: Now assume that ⊕a is of type ∞. We will prove that z∗ b0 > 0 such that (1, b0)a = 1. Then, (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) w0(cid:107)w0(cid:107) , b0z0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)a (cid:19) (cid:28)(cid:18) w0(cid:107)w0(cid:107) , b0z0 (cid:18) w∗ (cid:18)(cid:18) w0(cid:107)w0(cid:107) , b0z0 (cid:18) w∗ (cid:19) (cid:18) w0(cid:107)w0(cid:107) , b0z0 (cid:18) (cid:19) (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) w0(cid:107)w0(cid:107) , b0z0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)a (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S This implies that 0(cid:107) , 0 0(cid:107)w∗ 0(cid:107)w∗ and = 1 So, , , (cid:54) (1, b0)a = 1. = 1 (cid:19)(cid:29) 0(cid:107) , 0 (cid:19)(cid:19) (cid:18) w∗ (cid:19) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:54) v((cid:101)S) = 1. 0(cid:107) , 0 0(cid:107)w∗ , (w0, z0) = 1 − 1 b0(cid:107)w0(cid:107) (w0, 0) + 1 b0(cid:107)w0(cid:107) (w0, b0(cid:107)w0(cid:107)z0). ∈ Π(W ⊕a Z). Since ⊕a∗ is of type 1 (see Lemma 1.4.d), there is K > 0 such that (cid:107)w∗ 0(cid:107) + K(cid:107)z∗ 0(cid:107) (cid:54) (cid:107)(w∗ 0, z∗ 0 )(cid:107)a∗ = 1. Set Then, 1 = (cid:104)(cid:101)S(w0, z0), (w∗ 0, 0)(cid:105) + (cid:104)(cid:101)S(w0, z0), (0, z∗ = (cid:104)(cid:101)S(w0, z0), (w∗ 0 )(cid:105) 0, z∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:19) 0 )(cid:105) (cid:19) (cid:18) (cid:104)(cid:101)S(w0, 0), (w∗ b0(cid:107)w0(cid:107) (cid:18)(cid:18) (w0, 0) + 1 − 1 − (cid:54) (cid:54) 1 1 b0(cid:107)w0(cid:107) (cid:19) (cid:29)(cid:12)(cid:12)(cid:12)(cid:12) + (cid:104)(cid:101)S(w0, z0), (0, z∗ 0 )(cid:105) 0, 0)(cid:105) + (cid:104)(cid:101)S(w0, z0), (0, z∗ 0 )(cid:105) 1 b0(cid:107)w0(cid:107) (w0, b0(cid:107)w0(cid:107)z0) 0, 0)(cid:105) + b0(cid:107)w0(cid:107)(cid:104)(cid:101)S(w0, b0(cid:107)w0(cid:107)z0), (w∗ , (w∗ 0, 0) 1 THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 1 1 , , 1 − 1 − (cid:107)w∗ 0(cid:107)w∗ 0(cid:107) , 0 (cid:18) (cid:18) (cid:19) (cid:19) b0(cid:107)w0(cid:107) b0(cid:107)w0(cid:107) (cid:18) w∗ b0(cid:107)w0(cid:107)z0 (cid:107)w0(cid:107) (cid:19) (cid:18) w∗ 0(cid:107)(cid:107)w0(cid:107) (cid:107)w∗ 0(cid:107) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18) w0(cid:107)w0(cid:107) , 0 (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) w0(cid:107)w0(cid:107) , (cid:18) w∗ (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S (cid:18) w0(cid:107)w0(cid:107) , b0z0 0(cid:107)(cid:107)w0(cid:107) 0(cid:107)(cid:107)w0(cid:107) 0(cid:107)v((cid:101)S) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:18) z∗ (cid:19)(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:107)z∗ (cid:12)(cid:12)(cid:12)(cid:12) (cid:54) (cid:107)z∗ (cid:19)(cid:101)S2(w0, z0) 0(cid:107)(cid:107)(cid:101)S2(cid:107) < ε(cid:107)z∗ 0(cid:107). 0, 0)(cid:105) + (cid:104)(cid:101)S(w0, z0), (0, z∗ b0(cid:107)w0(cid:107)(cid:104)(cid:101)S(w0, b0(cid:107)w0(cid:107)z0), (w∗ 0 )(cid:105) b0(cid:107)w0(cid:107)(cid:107)w∗ b0(cid:107)w0(cid:107)(cid:107)w∗ b0(cid:107)w0(cid:107)(cid:107)w∗ b0(cid:107)w0(cid:107)(cid:107)w∗ 0(cid:107). 0(cid:107)z∗ 0(cid:107) 0(cid:107)w∗ 0(cid:107) , 0 0(cid:107)w∗ 0(cid:107) , 0 (cid:19) 1 1 1 0(cid:107) 1 , 15 (cid:3) (cid:18) But we have that 1 − 1 b0(cid:107)w0(cid:107) (cid:19) (cid:104)(cid:101)S(w0, 0), (w∗ 0, 0)(cid:105) = (cid:54) and b0(cid:107)w0(cid:107)(cid:104)(cid:101)S(w0, b0(cid:107)w0(cid:107)z0), (w∗ 1 0, 0)(cid:105) = = Moreover, by using (5), we have that (cid:104)(cid:101)S(w0, z0), (0, z∗ Therefore, for all ε ∈ (0, K), we have (cid:18) (cid:18) 1 (cid:54) 1 − 1 b0(cid:107)w0(cid:107) 1 1 − (cid:54) b0(cid:107)w0(cid:107) 0(cid:107) + ε(cid:107)z∗ = (cid:107)w∗ This contradiction gives z∗ Now, since ((w0, b0z0), (w∗ (cid:54) 1 1 0, (cid:18) z∗ 0(cid:107)z∗ 0(cid:107) (cid:12)(cid:12)(cid:12)(cid:12)(cid:28)(cid:101)S(w0, z0), 0(cid:107) 0 )(cid:105) = (cid:107)z∗ (cid:19) (cid:104)(cid:101)S(w0, 0), (w∗ (cid:19) 0, 0)(cid:105) + b0(cid:107)w0(cid:107)(cid:107)w∗ 0(cid:107) + (cid:107)w∗ 0(cid:107) + ε(cid:107)z∗ 0(cid:107) 0(cid:107) (cid:54) 1. 0(cid:107) + K(cid:107)z∗ 0(cid:107) < (cid:107)w∗ 1 = (cid:104)(cid:101)S(w0, z0), (w∗ 1 = (cid:104)(cid:101)S(w0, z0), (w∗ 0, 0)) ∈ Π(W ), we have that (cid:19) (cid:18) 0, 0)(cid:105) (cid:104)(cid:101)S(w0, 0), (w∗ (cid:18) (cid:19) (cid:18) (cid:19) 0(Sw0) + 0 = 0. So, (w0, w∗ w∗ (cid:54) (cid:54) 1 − 1 b0 1 − 1 b0 1 − 1 b0 v(S) + 1 b0 (cid:54) 0) ∈ Π(W ) and 0, 0)(cid:105) = w∗ 0((cid:101)S1(w0, z0)). (cid:104)(cid:101)S(w0, b0z0), (w∗ 0, 0)(cid:105) 1 b0 0, 0)(cid:105) + v((cid:101)S) 1 v((cid:101)S) (cid:54) 1. b0 So, w∗ 0(Sw0) = v(S) = 1 and this finishes the proof. As far as we know, Proposition 4.7 is new even for L- and M -summands. Corollary 4.8. Let X be Banach space. (a) If W is an L-summand of X and NRA(X) = L(X), then NRA(W ) = L(W ) (b) If W is an M -summand of X and NRA(X) = L(X), then NRA(W ) = L(W ) Adapting Proposition 4.7 to the compact case, we have the following result. Proposition 4.9. Let X be a Banach space and let W be an absolute summand of X of type 1 or ∞. If NRA(X) ∩ K(X) is dense in K(X), then NRA(W ) ∩ K(W ) is dense in K(W ). Let us finally discuss on the validity of reciprocal results. In [23, Remark 2.5] that the sets NRA(C[0, 1] ⊕1 L1[0, 1]) and NRA(C[0, 1] ⊕∞ L1[0, 1]) are not dense, respectively, in L(C[0, 1] ⊕1 L1[0, 1]) and L(C[0, 1] ⊕∞ L1[0, 1]) was observed. As both C[0, 1] and L1[0, 1] have the BPBp for numerical radius [6, 23], this shows that there is no possible valid reciprocal results for Theorem 4.1 and Proposition 4.7. On the other hand, the sets NRA(C[0, 1]⊕1 L1[0, 1])∩K(C[0, 1]⊕1 L1[0, 1]) and NRA(C[0, 1]⊕∞ L1[0, 1])∩K(C[0, 1]⊕∞ L1[0, 1]) are dense 16 CHOI, DANTAS, JUNG, AND MART´IN in K(C[0, 1] ⊕1 L1[0, 1]) and K(C[0, 1] ⊕∞ L1[0, 1]), respectively (see [12, Example 3.4]). Nevertheless, we do not know if there is some reciprocal result for the BPBp-nu for compact operators. References [1] M. D. Acosta, Denseness of norm attaining mappings, RACSAM 100 (2006), 9 -- 30. [2] M. D. Acosta, R. M. Aron, D. Garc´ıa, and M. Maestre, The Bishop-Phelps-Bollob´as theorem for operators, J. Funct. Anal. 294 (2008), 2780 -- 2899. [3] M. D. Acosta, M. Fakhar, and M. Soleimani-Mourchehkhorti, The Bishop-Phelps-Bollob´as property for numerical radius of operators on L1(µ), J. Math. Anal. Appl. 458 (2018), 925 -- 936. [4] M. D. Acosta, M. Masty(cid:32)lo, and M. Soleimani-Mourchehkhorti, The Bishop-Phelps-Bollob´as and approximate hyperplane series properties, J. Funct. Anal. 274 (2018), no. 9, 2673 -- 2699. [5] R. M. Aron, Y. S. Choi, S. K. Kim, H. J. Lee, and M. Mart´ın, The Bishop-Phelps-Bollob´as version of Lindenstrauss properties A and B, Trans. Amer. Math. Soc. 367 (2015), 6085 -- 6101. [6] A. Avil´es, A. J. Guirao, and J. Rodr´ıguez, On the Bishop-Phelps-Bollob´as property for numerical radius in C(K) spaces, J. Math. Anal. Appl. 419 (2014), 395 -- 421. [7] E. Bishop and R. R. Phelps, A proof that every Banach space is reflexive, Bull. Amer. Math. Soc. 67 (1961), 97 -- 98. [8] B. Bollob´as, An extension to the Theorem of Bishop and Phelps, Bull. London Math. Soc. 2 (1970), 181 -- 182. [9] F. F. Bonsall and J. Duncan, Numerical ranges of operators on normed spaces and of elements of normed algebras, London Mathematical Society Lecture Note Series, 2. Cambridge University Press, London, New York (1971). [10] F. F. Bonsall and J. Duncan, Numerical ranges II, London Mathematical Society Lecture Notes Series, No. 10. Cambridge University Press, New York, London (1973). [11] J. Bourgain, On dentability and the Bishop-Phelps property, Israel J. Math. 78 (1977), 265 -- 271. [12] A. Capel, M. Mart´ın, and J. Mer´ı, Numerical radius attaining compact linear operators, J. Math. Anal. Appl. 445, 1258 -- 1266. [13] M. Chica, V. Kadets, M. Mart´ın, F. Rambla-Barreno, and S. Moreno-Pulido, Bishop-Phelps-Bollob´as modudi of a Banach space, J. Math. Anal. Appl. 412 (2014), 697 -- 719. [14] M. Chica, M. Mart´ın, and J. Mer´ı, Numerical radius of rank-1 operators on Banach spaces, Quart. J. Math. 65 (2014), 89 -- 100. [15] D. H. Cho and Y. S. Choi, The Bishop-Phelps-Bollob´as theorem on bounded closed convex sets, J. Lond. Math. Soc. 93 (2016), 502 -- 518. [16] S. Dantas, D. Garc´ıa, M. Maestre, and M. Mart´ın, The Bishop-Phelps-Bollob´as property for compact operators, Can. J. Math. 70 (2018), no. 1, 53-73. [17] S. Dantas, V. Kadets, S. K. Kim, H. J. Lee, and M. Mart´ın, On the pointwise Bishop -- Phelps -- Bollob´as property for operators, accepted in Can. J. Math. doi:10.4153/S0008414X18000032. [18] S. Dantas, S. K. Kim, and H. J. Lee, The Bishop-Phelps-Bollob´as point property, J. Math. Anal. Appl. 444 (2016), 1739 -- 1751. [19] F. J. Garc´ıa-Pacheco, The AHSP is inherited by E-summands, Adv. Oper. Theory 2 (2017), 17 -- 20. [20] A. Guirao and O. Kozhushkina, The Bishop-Phelp-Bollob´as property for numerical radius in (cid:96)1(C), Studia Math. 218 (2013), 41 -- 54. [21] J. Hardtke, Absolute sums of Banach spaces and some geometric properties related to rontundity and smoothness, Banach J. Math. Anal. 8 (2014), 295 -- 334. [22] P. Harmand, D. Werner, and D. Werner, M -ideals in Banach spaces and Banach algebras, Lecture Notes in Math. 1547, Springer-Verlag, Berlin 1993. [23] S. K. Kim, H. J. Lee, and M. Mart´ın, On the Bishop-Phelps-Bollob´as property for numerical radius, Abs. Appl. Anal. vol. 2014, Article ID 479208, 15 pages, 2014. [24] S. K. Kim, H. J. Lee, M. Mart´ın, and J. Mer´ı, On the second numerical index for Banach spaces, accepted in Proc. Royal Soc. Edinburgh Sect A . [25] S. K. Kim, H. J. Lee, and M. Mart´ın, The Bishop-Phelps-Bollob´as theorem for operators from (cid:96)1 sums, J. Math. Anal. Appl. 428 (2015), 920 -- 929. [26] S. K. Kim, H. J. Lee, and M. Mart´ın, On the Bishop-Phelps-Bollob´as theorem for operators and numerical radius, Studia Math. 233 (2016), 141 -- 151. [27] J. Lindenstrauss, On operators which attain their norm, Israel J. Math. 1 (1963), 139 -- 148. [28] J. F. Mena, R. Pay´a and A. Rodr´ıguez, Semisummands and semiideals in Banach spaces, Israel J. Math., 52, (1985), 33 -- 67. [29] J. F. Mena, R. Pay´a and A. Rodr´ıguez, Absolute subspaces of Banach spaces, Quart. J. Math., 40, (1989), 33 -- 37. [30] J. F. Mena, R. Pay´a and A. Rodr´ıguez, Absolutely proximinal subspaces of Banach spaces, J. Approx. Theory 65 (1991), 46 -- 72. [31] R. Pay´a, T´ecnicas de Rango Num´erico y Estructura en Espacios Normados, Ph.D. Dissertation, Universidad de Granada, Spain, 1980. Available at http://hdl.handle.net/10481/52674. [32] R. Pay´a, A counterexample on numerical radius attaining operators, Israel J. Math. 79 (1992), 83 -- 101. [33] R. Pay´a and Y. Saleh, Norm attaining operators from L1(µ) into L∞(ν), Arch. Math. 75 (2000), 380 -- 388. THE BISHOP-PHELPS-BOLLOB ´AS PROPERTY AND ABSOLUTE SUMS 17 (Choi) Department of Mathematics, POSTECH, Pohang 790-784, Republic of Korea E-mail address: [email protected] (Dantas) Department of Mathematics, Faculty of Electrical Engineering, Czech Technical University in Prague, Technick´a 2, 166 27, Prague 6, Czech Republic ORCID: 0000-0001-8117-3760 E-mail address: [email protected] (Jung) Department of Mathematics, POSTECH, Pohang 790-784, Republic of Korea ORCID: 0000-0003-2240-2855 E-mail address: [email protected] (Mart´ın) Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain ORCID: 0000-0003-4502-798X E-mail address: [email protected]
1808.01467
2
1808
2018-12-18T20:21:27
Sobolev functions on closed subsets of the real line: long version
[ "math.FA" ]
For each $p>1$ and each positive integer $m$ we give intrinsic characterizations of the restriction of the Sobolev space $W^m_p(R)$ and homogeneous Sobolev space $L^m_p(R)$ to an arbitrary closed subset $E$ of the real line. In particular, we show that the classical one dimensional Whitney extension operator is "universal" for the scale of $L^m_p(R)$ spaces in the following sense: for every $p\in(1,\infty]$ it provides almost optimal $L^m_p$-extensions of functions defined on $E$. The operator norm of this extension operator is bounded by a constant depending only on $m$. This enables us to prove several constructive $W^m_p$- and $L^m_p$-extension criteria expressed in terms of $m^{th}$ order divided differences of functions.
math.FA
math
Sobolev functions on closed subsets of the real line: long version By Pavel Shvartsman Department of Mathematics, Technion - Israel Institute of Technology 32000 Haifa, Israel e-mail: [email protected] Abstract For each p > 1 and each positive integer m we give intrinsic characterizations of the p (R) to an arbitrary p (R) and homogeneous Sobolev space Lm restriction of the Sobolev space Wm closed subset E of the real line. In particular, we show that the classical one dimensional Whitney extension operator [67] p (R) spaces in the following sense: for every p ∈ (1,∞] it is "universal" for the scale of Lm provides almost optimal Lm p -extensions of functions defined on E. The operator norm of this extension operator is bounded by a constant depending only on m. This enables us to prove several constructive Wm p -extension criteria expressed in terms of m th order divided differences of functions. p - and Lm Contents 1. Introduction. 2. Main Theorems: necessity. 2.1. Divided differences: main properties. 2.2. Proofs of the necessity part of the main theorems. 3. The Whitney extension method in R and traces of Sobolev functions. 3.1. Interpolation knots and their properties. 3.2. Lagrange polynomials and divided differences at interpolation knots. 3.3. Whitney m-fields and the Hermite polynomials. 3.4. Extension criteria in terms of sharp maximal functions: sufficiency. 4. A variational criterion for Sobolev traces. 4.1. A variational criterion for Sobolev jets. 4.2. The Main Lemma: from jets to Lagrange polynomials. 4.3. Proof of the sufficiency part of the variational criterion. 4.4. Lm 4.5. The variational criterion and the sharp maximal function-type criterion. p -functions on increasing sequences of the real line. 2 9 10 12 13 14 17 22 26 29 29 36 48 53 60 Math Subject Classification 46E35 Key Words and Phrases Sobolev space, trace space, divided difference, extension operator. This research was supported by Grant No 2014055 from the United States-Israel Binational Science Foundation (BSF). 1 5. Extension criteria for Sobolev Wm p -functions. 5.1. The variational criterion for Wm 5.2. The variational criterion for Wm 5.3. Wm 5.4. Wm 5.5. Further remarks and comments. p (R)-functions on sequences of points. p (R)-restrictions and local sharp maximal functions. p -traces: necessity. p -traces: sufficiency. 6. The Finiteness Principle for Lm∞(R) traces: multiplicative finiteness constants. 6.1. Multiplicative finiteness constants of the space Lm∞(R). 6.2. Extremal functions and finiteness constants. References 61 62 64 83 90 94 97 97 100 104 1. Introduction. In this paper we characterize the restrictions of Sobolev functions of one variable to an arbitrary closed subset of the real line. Given m ∈ N and p ∈ [1,∞], we let Lm p (R) denote the standard homogeneous Sobolev space on R. We identify Lm p (R) with the space of all real valued functions F on R such that the (m − 1)-th derivative F(m−1) is absolutely continuous on R and the weak m-th derivative F(m) ∈ Lp(R). Lm p (R) is seminormed by (cid:107)F(cid:107)Lm p (R) = (cid:107)F(m)(cid:107)Lp(R) . As usual, we let Wm p (R) denote the corresponding Sobolev space of all functions F ∈ Lm p (R) whose derivatives on R of all orders up to m belong to Lp(R). This space is normed by m(cid:88) (cid:107)F(cid:107)Wm p (R) = (cid:107)F(k)(cid:107)Lp(R). In this paper we study the following k=0 Problem 1.1 Let p ∈ (1,∞], m ∈ N, and let E be a closed subset of R. Let f be a function on E. We ask two questions: p (R) such that the restriction FE p (R) such that FE = f . How small can 1. How can we decide whether there exists a function F ∈ Wm p (R)-norms of all functions F ∈ Wm of F to E coincides with f ? 2. Consider the Wm these norms be? We denote the infimum of all these norms by (cid:107) f(cid:107)Wm p (R)E; thus (cid:107) f(cid:107)Wm p (R)E = inf{(cid:107)F(cid:107)Wm p (R) : F ∈ Wm p (R), FE = f}. (1.1) We refer to (cid:107) f(cid:107)Wm standard quotient space norm in the trace space Wm E, i.e., in the space p (R)E as the trace norm of the function f in Wm p (R)E of all restrictions of Wm p (R). This quantity provides the p (R)-functions to p (R)E = { f : E → R : there exists F ∈ Wm Wm p (R) such that FE = f}. 2 Theorem 1.2, our main contribution in this paper, provides a complete solution to Problem 1.1 for the case p ∈ (1,∞). Let us prepare the ingredients that are needed to formulate this theorem: Given a function f defined on a k + 1 point set S = {x0, ..., xk}, we let ∆k f [x0, ..., xk] denote the k th order divided difference of f on S . Recall that ∆m f [S ] coincides with the coefficient of xm in the Lagrange polynomial of degree at most m which agrees with f on S . See Section 2.1 for other equivalent definitions of divided differences and their main properties. sequence {x0, ..., xn} ⊂ E we set Everywhere in the paper we will use the following notation: Given a finite strictly increasing xi = +∞ if i > n. (1.2) Here now is the main result of our paper: Theorem 1.2 Let m be a positive integer, p ∈ (1,∞), and let E be a closed subset of R containing at least m + 1 points. A function f : E → R can be extended to a function F ∈ Wm p (R) if and only if the following quantity  m(cid:88) k=0 n−k(cid:88) i=0 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p  1 p (1.3) Wm,p( f : E) = sup {x0,...,xn}⊂E is finite. Here the supremum is taken over all finite strictly increasing sequences {x0, ..., xn} ⊂ E with n ≥ m. Furthermore, The constants in equivalence (1.4) depend only on m. (cid:107) f(cid:107)Wm p (R)E ∼ Wm,p( f : E) . (1.4) We refer to this result as a variational criterion for the traces of Wm An examination of our proof of Theorem 1.2 shows that when E is a strictly increasing sequence of points in R (finite, one-sided infinite, or bi-infinite) it is enough to take the supremum in (1.3) over a unique subsequence of E - the sequence E itself. Therefore, in the particular case of such sets E, we can obtain the following refinement of Theorem 1.2: Theorem 1.3 Let (cid:96)1, (cid:96)2 ∈ Z ∪ {±∞}, (cid:96)1 ≤ (cid:96)2, and let p ∈ (1,∞). Let E = {xi}(cid:96)2 increasing sequence of points in R, and let p (R)-functions. be a strictly i=(cid:96)1 mE = min{m, #E − 1}. 3 (1.5) A function f ∈ Wm p (R)E if and only if the following quantity (cid:102)Wm,p( f : E) =  mE(cid:88) k=0 (cid:96)2−k(cid:88) i=(cid:96)1 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p  1 p (1.6) is finite. (Note that, according to the notational convention adopted in (1.2), it must be understood that in (1.6) we have xi = +∞ for every i > (cid:96)2.) Furthermore, p (R)E ∼ (cid:102)Wm,p( f : E) (cid:107) f(cid:107)Wm with constants in this equivalence depending only on m. For a special case of Theorem 1.3 for m = 2 and strictly increasing sequences {xi}i∈Z with p (R)E see xi+1 − xi ≤ const, and other results related to characterization of the trace space Wm Est´evez [22]. p (R)E expressed in terms In Section 5.4 we give another characterization of the trace space Wm of Lp-norms of certain kinds of "sharp maximal functions" which are defined as follows. Definition 1.4 For each m ∈ N, each closed set E ⊂ R, each function f : E → R and each integer k in the range 0 ≤ k ≤ m − 1, we let f # k,E denote the maximal function associated with f which is given by f # k,E(x) = sup S⊂E, #S =k+1, dist(x,S )≤1 ∆k f [S ], x ∈ R, and, when k = m, by f # m,E(x) = sup S⊂E, #S =m+1, dist(x,S )≤1 diam S diam(S ∪ {x}) ∆m f [S ], x ∈ R. (1.7) (1.8) Here for every k = 0, ..., m, the above two suprema are taken over all (k + 1)-point subsets S ⊂ E such that dist(x, S ) ≤ 1. If the family of sets S satisfying these conditions is empty, we put f # k,E(x) = 0. Theorem 1.5 Let m ∈ N and let p ∈ (1,∞). A function f ∈ Wm k,E ∈ Lp(R) for every k = 0, ..., m. f # Furthermore, p (R)E if and only if the function m(cid:88) (cid:107) f(cid:107)Wm p (R)E ∼ (cid:107) f # k,E(cid:107)Lp(R) with constants in this equivalence depending only on m and p. k=0 We feel a strong debt to the remarkable papers of Calder´on and Scott [9, 10] which are devoted to characterization of Sobolev spaces on Rn in terms of classical sharp maximal functions. These papers motivated us to formulate and subsequently prove Theorem 1.5. corresponding trace space Lm We turn to a variant of Problem 1.1 for the homogeneous Sobolev space Lm p (R)E and trace seminorm in Lm p (R)E = { f : E → R : there exists F ∈ Lm Lm p (R)E by letting p (R) such that FE = f} p (R). We define the 4 and (cid:107) f(cid:107)Lm p (R)E = inf{(cid:107)F(cid:107)Lm p (R) : F ∈ Lm p (R), FE = f}. (1.9) Whitney [67] completely solved an analog of the part 1 of Problem 1.1 for the space Cm(R). Whitney's extension construction [67] produces a certain extension operator m,E : Cm(R)E → Cm(R) F (Wh) (1.10) which linearly and continuously maps the trace space Cm(R)E into Cm(R). An important ingredient of this construction is the classical Whitney's extension method for Cm-jets [66]. (See also Merrien [48].) The extension method developed by Whitney in [67] also provides a complete solution to Prob- lem 1.1 for the space Lm∞(R). Recall that Lm∞(R) can be identified with the space Cm−1,1(R) of all Cm−1-functions on R whose derivatives of order m− 1 satisfy the Lipschitz condition. In particular, the method of proof and technique developed in [67] and [48] lead us to the following well known description of the trace space Lm∞(R)E: A function f ∈ Lm∞(R)E if and only if the following quantity (1.11) Lm,∞( f : E) = ∆m f [S ] sup S⊂E, #S =m+1 is finite. Furthermore, C1 Lm,∞( f : E) ≤ (cid:107) f(cid:107)Lm∞(R)E ≤ C2 Lm,∞( f : E) (1.12) where C1 and C2 are positive constants depending only on m. i=(cid:96)1 We refer the reader to [41, 56] for further results in this direction. There is an extensive literature devoted to a special case of Problem 1.1 where E consists of all the elements of a strictly increasing sequence {xi}(cid:96)2 (finite, one-sided infinite, or bi-infinite). We refer the reader to the papers of Favard [23], Chui, Smith, Ward [11 -- 13, 64, 65], Karlin [42], de Boor [15 -- 20], Fisher and Jerome [32, 33], Golomb [36], Jakimovski and Russell [38], Kunkle [44, 45], Pinkus [49, 50], Schoenberg [52 -- 54] and references therein for numerous results in this direction and techniques for obtaining them. In particular, for the space Lm∞(R) Favard [23] developed a powerful linear extension method (very different from Whitney's method [67]) based on a certain delicate duality argument. Note that for any set E as above and every f : E → R, Favard's extension operator F (Favard) yields an ex- tension of f with the smallest possible seminorm in Lm∞(R). (Thus, (cid:107) f(cid:107)Lm∞(R)E ( f )(cid:107)Lm∞(R) for every function f defined on E.) Note also that Favard's approach leads to the following slight refinement of (1.12): = (cid:107)F (Favard) m,E m,E (cid:107) f(cid:107)Lm∞(R)E ∼ sup (cid:96)1≤i≤(cid:96)2−m ∆m f [xi, ..., xi+m] . See Section 6 for more detail. Modifying Favard's extension construction, de Boor [17 -- 19] characterized the traces of Lm p (R)- functions to arbitrary sequences of points in R. Theorem 1.6 ( [17]) Let p ∈ (1,∞), and let (cid:96)1, (cid:96)2 ∈ Z ∪ {±∞}, (cid:96)1 + m ≤ (cid:96)2. Let f be a function p (R)E if an only if the defined on a strictly increasing sequence of points E = {xi}(cid:96)2 following quantity . Then f ∈ Lm i=(cid:96)1 (1.13) (cid:101)Lm,p( f : E) = (cid:96)2−m(cid:88) i=(cid:96)1  1 p (xi+m − xi)∆m f [xi, ..., xi+m]p 5 is finite. Furthermore, (cid:107) f(cid:107)Lm p (R)E ∼ (cid:101)Lm,p( f : E) with constants depending only on m. For a special case of this result for sequences satisfying some global mesh ratio restrictions see Golomb [36]. See also Est´evez [22] for an alternative proof of Theorem 1.6 for m = 2. Using a certain limiting argument, Golomb [36] showed that Problem 1.1 for Lm p (R) and an arbitrary set E ⊂ R can be reduced to the same problem, but for arbitrary finite sets E. More specifically, his result (in an equivalent form) provides the following formula for the trace norm in p (R)E: Lm (cid:107) f(cid:107)Lm p (R)E = sup{(cid:107) fE(cid:48)(cid:107)Lm p (R)E(cid:48) : E ⊂ E, #E(cid:48) < ∞}. following description of the trace space Lm Let us remark that, by combining this formula with de Boor's Theorem 1.6, we can obtain the p (R)E for an arbitrary closed set E ⊂ R. p (R)-traces) Let p ∈ (1,∞) and let m be a positive inte- Theorem 1.7 (Variational criterion for Lm ger. Let E ⊂ R be a closed set containing at least m + 1 points. A function f : E → R can be extended to a function F ∈ Lm p (R) if and only if the following quantity Lm,p( f : E) = sup {x0,...,xn}⊂E (xi+m − xi)∆m f [xi, ..., xi+m]p (1.14) is finite. Here the supremum is taken over all finite strictly increasing sequences {x0, ..., xn} ⊂ E with n ≥ m. Furthermore, (cid:107) f(cid:107)Lm p (R)E ∼ Lm,p( f : E) . (1.15) The constants in equivalence (1.15) depend only on m. In the present paper we give a direct and explicit proof of Theorem 1.7 which does not use any limiting argument. Actually we show, perhaps surprisingly, that the very same Whitney extension operator F (Wh) (see (1.10)) which was introduced in [67] for characterization of the trace space Cm(R)E, provides almost optimal extensions of functions belonging to Lm p (R)E for every p ∈ (1,∞]. m,E  n−m(cid:88) i=0  1 p space Lm defined as follows: In Section 3 we prove an analogue of Theorem 1.5 which enables us to characterize the trace p (R)E in terms of of Lp-norms of certain kinds of "sharp maximal functions" which are : E → R we let For each m ∈ N, each closed set E ⊂ R with #E > m, and each function f (∆m f )(cid:93) E denote the maximal function associated with f which is given by ∆m−1 f [x0, ..., xm−1] − ∆m−1 f [x1, ..., xm] (∆m f )(cid:93) E (x) = sup {x0,...,xm}⊂E x0<x1<...<xm x − x0 + x − xm x ∈ R . , (1.16) Theorem 1.8 Let p ∈ (1,∞), m ∈ N, and let f be a function defined on a closed set E ⊂ R. The function f ∈ Lm p (R)E if and only if (∆m f )(cid:93) (cid:107) f(cid:107)Lm E ∈ Lp(R). Furthermore, p (R)E ∼ (cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) with the constants in this equivalence depending only on m and p. 6 Remark 1.9 Note that (∆m f )(cid:93) E (x) ≤ sup S⊂E, #S =m+1 ∆m f [S ] diam S diam({x} ∪ S ) ≤ 2 (∆m f )(cid:93) E (x), x ∈ R . (1.17) (See property (2.3) below.) Theorem 1.8, (1.16) and this inequality together now imply two explicit formulae for the trace norm of a function f in the space Lm p (R)E:   (cid:90) (cid:90) R R (cid:107) f(cid:107)Lm p (R)E ∼ ∼ 1 p  dx ∆m−1 f [x0, ..., xm−1] − ∆m−1 f [x1, ..., xm] p x − x0p + x − xmp sup {x0,...,xm}⊂E x0<x1<...<xm sup S⊂E, #S =m+1 (cid:32)∆m f [S ] diam S diam({x} ∪ S ) (cid:33)p  1 p dx . (cid:67) For versions of Theorems 1.7 and 1.8 for the space L1 p(Rn), n ∈ N, n < p < ∞, we refer the reader to [59, 63]. The next theorem states that there are solutions to Problem 1.1 and its analogue for the space Lm p (R) which depend linearly on the initial data, i.e., the functions defined on E. Theorem 1.10 For every closed subset E ⊂ R, every p > 1 and every m ∈ N there exists a continuous linear extension operator which maps the trace space Lm p (R). Its operator norm is bounded by a constant depending only on m. p (R)E into Lm The same statement holds for the space Wm p (R). Remark 1.11 As we have noted above, for every p ∈ (1,∞] the Whitney extension operator F (Wh) m,E (see (1.10)) provides almost optimal extensions of functions from Lm p (R). is linear, it has the properties described in Theorem 1.10. Slightly modifying F (Wh) Since F (Wh) m,E , we construct a continuous linear extension operator from Wm p (R) with the operator norm bounded by a constant depending only on m. See (1.19). p (R)E to functions from Lm p (R)E into Wm m,E (cid:67) Let us recall something of the history of Theorem 1.10. We know that for each closed E ⊂ R m,E maps Lm∞(R)E into Lm∞(R) with the operator norm (cid:107)F (Wh) m,E (cid:107) the Whitney extension operator F (Wh) bounded by a constant depending only on m. As we have mentioned above, if E is a sequence of points in R, the Favard's linear extension operator also maps Lm∞(R)E into Lm∞(R), but with the operator norm (cid:107)F (Favard) For p ∈ (1,∞) and arbitrary sequence E ⊂ R Theorem 1.10 follows from [17, Section 4]. Luli [46] gave an alternative proof of Theorem 1.10 for the space Lm p (R) and a finite set E. In the multidimensional case the existence of corresponding linear continuous extension operators for the p (Rn), n < p < ∞, was proven in [59] (m = 1, n ∈ N, E ⊂ Rn is arbitrary), [37] Sobolev spaces Lm and [60] (m = 2, n = 2, E ⊂ R2 is finite), and [30] (arbitrary m, n ∈ N and an arbitrary E ⊂ Rn). For the case p = ∞ see [7] (m = 2) and [26, 27] (m ∈ N). (cid:107) = 1. m,E Let us briefly describe the structure of the paper and the main ideas of our approach. First we note that equivalence (1.15) is not trivial even in the simplest case, i.e., for E = R; in this case (1.15) tells us that for every f ∈ Lm p (R) and every p ∈ (1,∞] (cid:107) f(cid:107)Lm p (R) ∼ Lm,p( f : R) 7 (1.18) p (R). This characterization of the space Lm with constants depending only on m. In other words, the quantity Lm,p(· : R) provides an equivalent seminorm on Lm p (R) is known in the literature; see F. Riesz [51] (m = 1 and 1 < p < ∞), Schoenberg [53] (p = 2 and m ∈ N), and Jerome and Schumaker [40] (arbitrary m ∈ N and p ∈ (1,∞)). Of course, equivalence (1.18) implies the necessity part of Theorem 1.7. Nevertheless, for the readers convenience, in Section 2.2 we give a short direct proof of this result (together with the proof of the necessity part of Theorem 1.8). In Section 3 we recall the Whitney extension method [67] for functions of one variable. We prove a series of auxiliary statements which enable us to adapt Whitney's construction to extension of Lm p (R)-functions. We then use this extension technique and a criterion for extension of Sobolev jets [63] to help us prove the sufficiency part of Theorem 1.8. (See Section 3.4.) The proof of the sufficiency part of Theorem 1.7 is given in Section 4. One of the main ingredient of this proof is a new criterion for extensions of Sobolev jets distinct from Theorem 3.17. We prove this extension criterion in Section 4.1 (see Theorem 4.1). Another important ingredient of the proof of the sufficiency is Main Lemma 4.9. For every function on E this lemma provides a controlled transition from Hermite polynomials of the function (which are basic components of the Whitney's construction) to its Lagrange polynomials. See Section 4.2. These two results, Theorem 4.1 and Main Lemma 4.9, are key elements of our proof of the sufficiency part of Theorem 1.7 which we present in Section 4.3. Section 5 of the paper is devoted to the proof of Theorem 1.2 and Theorem 1.3, the variational criterion for the traces of Wm p (R)-functions and its variant for finite sets or sequences in R. A short proof of the necessity part of Theorem 1.2 is given in Section 5.1. We prove the sufficiency using a certain modification of the Whitney's extension method. More specifically, we introduce sets (cid:101)E, G ⊂ R by Given a function f on E, we let f denote a function on (cid:101)E which coincides with f on E and equals 0 on G (if G (cid:44) ∅). We prove that f satisfies the hypothesis of Theorem 1.7 with (cid:101)E = E(cid:83) G where G = {x ∈ R : dist(x, E) ≥ 1}. Lm,p( f : E) ≤ C1 Wm,p( f : E) (see (1.3) and (1.14)) provided f satisfies the hypothesis of Theorem 1.2 (i.e., Wm,p( f Theorem 1.7, the function f can be extended to a function F = F( f ) ∈ Lm C2 Wm,p( f : E). We also show that (cid:107)F(cid:107)Lp(R) ≤ C3 Wm,p( f : E) proving that F ∈ Wm Sobolev norm is bounded by C4 Wm,p( f only on m. This completes the proof of Theorem 1.2. : E) < ∞). Therefore, by p (R) ≤ p (R) and its : E). Here C1, ..., C4 are positive constants depending p (R) with (cid:107)F(cid:107)Lm Furthermore, the function f depends linearly on f and, by Theorem 1.10, one can choose F( f ) linearly depending on f so that the extension operator f → F( f ) depends linearly on f (1.19) as well. This proves Theorem 1.10 for the space Wm proof of Theorem 1.2, in Section 5.3 we prove Theorem 1.3. p (R). Slightly modifying and simplifying the In Section 6 we discuss the dependence on m of the constants C1, C2 in inequality (1.12). We interpret this inequality as a particular case of the Finiteness Principle for traces of smooth func- tions. (See Theorem 6.1 and Theorem 6.9 below). We refer the reader to [6, 8, 24, 25, 28, 58] and references therein for numerous results related to the Finiteness Principle. For the space Lm∞(R) the Finiteness Principle implies the following statement: there exists a constant γ = γ(m) such that for every closed set E ⊂ R and every f ∈ Lm∞(R)E the following 8 inequality (cid:107) f(cid:107)Lm∞(R)E ≤ γ sup S⊂E, #S =m+1 (cid:107) fS(cid:107)Lm∞(R)S (1.20) holds. We can express this result by stating that the number m+1 is a finiteness number for the space Lm∞(R). We also refer to any constant γ which satisfies (1.20) as a multiplicative finiteness constant for the space Lm∞(R). In this context we let γ(cid:93)(Lm∞(R)) denote the infimum of all multiplicative finiteness constants for Lm∞(R) for the finiteness number m + 1. One can easily see that γ(cid:93)(L1∞(R)) = 1. We prove that γ(cid:93)(L2∞(R)) = 2 and (π/2)m−1 < γ(cid:93)(Lm∞(R)) < (m − 1) 9m for every m > 2 . (1.21) See Theorem 6.3. The proof of (1.21) relies on results of Favard [23] and de Boor [17,18] devoted to calculation of certain extension constants for the space Lm∞(R). In particular, (1.21) implies the following sharpened version of the Finiteness Principle for Lm∞(R), m > 2: Let f be a function defined on a closed set E ⊂ R. Suppose that for every (m + 1)-point subset E(cid:48) ⊂ E there exists a function FE(cid:48) ∈ Lm∞(R) with (cid:107)FE(cid:48)(cid:107)Lm∞(R) ≤ 1, such that FE(cid:48) = f on E(cid:48). Then there exists a function F ∈ Lm∞(R) with (cid:107)F(cid:107)Lm∞(R) ≤ (m − 1) 9m such that F = f Furthermore, there exists a closed set (cid:101)E ⊂ R and a function f : (cid:101)E → R such that for every on E. (m + 1)-point subset E(cid:48) ⊂ (cid:101)E there exists a function FE(cid:48) ∈ Lm∞(R) with (cid:107)FE(cid:48)(cid:107)Lm∞(R) ≤ 1, such that FE(cid:48) = f on E(cid:48), but nevertheless, (cid:107)F(cid:107)Lm∞(R) ≥ (π/2)m−1 for every F ∈ Lm∞(R) such that F = f on(cid:101)E. See Section 6 for more details. For journal versions of the results presented in this paper we refer the reader to [61], [62]. Acknowledgements. I am very thankful to M. Cwikel for useful suggestions and remarks. I am grateful to Charles Fefferman, Bo'az Klartag and Yuri Brudnyi for valuable conversations. The results of this paper were presented at the 11th Whitney Problems Workshop, Trinity College Dublin, Dublin, Ireland. I am very thankful to all participants of this conference for stimulating discussions and valuable advice. 2. Main Theorems: necessity. Let us fix some notation. Throughout the paper C, C1, C2, ... will be generic positive constants which depend only on m and p. These symbols may denote different constants in different occur- rences. The dependence of a constant on certain parameters is expressed by the notation C = C(m), C = C(p) or C = C(m, p). Given constants α, β ≥ 0, we write α ∼ β if there is a constant C ≥ 1 such that α/C ≤ β ≤ C α. Given a measurable set A ⊂ R, we let A denote the Lebesgue measure of A. If A ⊂ R is finite, by #A we denote the number of elements of A. Given δ > 0, we let [A]δ denote the δ-neighborhood of the set A. Given A, B ⊂ R, let diam A = sup{ a − a(cid:48) : a, a(cid:48) ∈ A} and dist(A, B) = inf{ a − b : a ∈ A, b ∈ B}. For x ∈ R we also set dist(x, A) = dist({x}, A). Finally, we put dist(A,∅) = +∞ provided A (cid:44) ∅. The notation A → x will mean that diam(A ∪ {x}) → 0. (2.1) 9 by M if every point x ∈ R is covered by at most M intervals from I. Given M > 0 and a family I of intervals in R we say that covering multiplicity of I is bounded Given a function g ∈ L1,loc(R) we let M[g] denote the Hardy-Littlewood maximal function of g: M[g](x) = sup I(cid:51)x 1 I g(y)dy, x ∈ R. (2.2) (cid:90) I Here the supremum is taken over all closed intervals I in R containing x. By Pm we denote the space of all polynomials of degree at most m defined on R. Finally, given a nonnegative integer k, a (k + 1)-point set S ⊂ R and a function f on S , we let LS [ f ] denote the Lagrange polynomial of degree at most k interpolating f on S ; thus LS [ f ] ∈ Pk and LS [ f ](x) = f (x) for every x ∈ S . 2.1. Divided differences: main properties. In this section we recall several useful properties of the divided differences of functions. We refer the reader to monographs [21, Ch. 4, §7] and [34, Section 1.3] for the proofs of these properties. Everywhere in this section k is a nonnegative integer and S = {x0, ..., xk} is a (k + 1)-point subset of R. In ((cid:70)1)-((cid:70)3) by f we denote a function defined on S . Then the following properties hold: ((cid:70)1) ∆0 f [S ] = f (x0) provided S = {x0} is a singleton. ((cid:70)2) If k ∈ N then ∆k f [S ] = ∆k f [x0, x1, ..., xk] = Furthermore, ∆k f [S ] = (cid:16) (cid:17) ∆k−1 f [x1, ..., xk] − ∆k−1 f [x0, ..., xk−1] k(cid:88) i=0 k(cid:88) f (xi) ω(cid:48)(xi) = f (xi)(cid:81) (xi − x j) i=0 j∈{0,...,k}, j(cid:44)i /(xk − x0). (2.3) (2.4) where ω(x) = (x − x0)...(x − xk). ((cid:70)3) We recall that LS [ f ] denotes the Lagrange polynomial of degree at most k = #S − 1 interpolating f on S . Then the following equality ∆k f [S ] = 1 k! L(k) S [ f ] (2.5) holds. Thus, ∆k f [S ] = Ak where Ak is the coefficient of xk of the polynomial LS [ f ]. ((cid:70)4) Let k ∈ N, and let x0 = min{xi : i = 0, ..., k} and xk = max{xi : i = 0, ..., k}. Then for every function F ∈ Ck[x0, xk] there exists ξ ∈ [x0, xk] such that 1 k! ∆kF[x0, x1, ..., xk] = F(k)(ξ) . (2.6) 10 ((cid:70)5) Let k ∈ N and let x0 < x1 < ... < xk. Let Mk = Mk[S ](t), t ∈ R, be the B-spline (basis spline) associated with the set S = {x0, ...xk}. We recall that given t ∈ R, the B-spline Mk[S ](t) is the divided difference of the function gt(u) = k · (u − t)k−1 Mk[S ](t) = ∆kgt[S ] = k k(cid:88) i=0 + , u ∈ R, over the set S ; thus, (xi − t)k−1 ω(cid:48)(xi) + . See [20] for this definition and various properties of B-splines. We know that 0 ≤ Mk[S ](t) ≤ and supp Mk[S ] ⊂ [x0, xk]. Furthermore, xk(cid:90) k xk − x0 for every t ∈ R, Mk[S ](t) dt = 1 . x0 ((cid:70)6) Let x0 < x1 < ... < xk, and let F be a function on [x0, xk] with absolutely continuous derivative of order k − 1. Then xk(cid:90) x0 ∆kF[S ] = 1 k! Mk[S ](t) F(k)(t) dt . (2.7) (2.8) (2.9) (2.10) (2.11) (2.12) See [20] or [21, p. 137]. This equality and inequality (2.8) tell us that ∆kF[S ] ≤ 1 (k − 1)! · Hence,  ∆kF[S ] ≤ 1 (k − 1)! xk(cid:90) 1 x0 xk − x0 xk(cid:90) 1 xk − x0 x0 F(k)(t) dt.  1 p F(k)(t)p dt for every p ∈ [1,∞). Furthermore, thanks to (2.9) and (2.10), for every {x0, ..., xm} ⊂ R, x0 < ... < xm, and every F ∈ Lm∞(R) the following inequality m!∆mF[x0, ..., xm] ≤ (cid:107)F(cid:107)Lm∞(R) (2.13) holds. function on {s0, ..., sn}, and let T = {t0, ..., tk} be a (k + 1)-point subset of Y. ((cid:70)7) Let k, n ∈ N, n ≥ k, and let {s0, ..., sn} ⊂ R, be a strictly increasing sequence. Let g be a (i) ( [34, p. 15]) There exist αi ∈ R, αi ≥ 0, i = 1, ..., n, such that α1 + ... + αn = 1 and ∆kg[T] = αi ∆kg[si, ..., si+k] . (2.14) n−k(cid:88) i=0 11 (ii) ( [20, p. 8], [14, 47]) Suppose that t0 = s0 and tk = sn. There exist numbers βi ∈ [0, (si+k − si)/(sn − s0)] for all i = 0, ..., n − k, such that ∆kg[T] = βi ∆kg[si, ..., si+k] . n−k(cid:88) i=0 2.2. Proofs of the necessity part of the main theorems. (Theorem 1.7: Necessity) Let 1 < p < ∞ and let f ∈ Lm p (R) be an arbitrary function such that FE = f . Let n ≥ m and let {x0, ..., xn} ⊂ E, x0 < ... < xn. Then, by (2.12), for every i, 0 ≤ i ≤ n − m, p (R)E. Let F ∈ Lm (xi+m − xi)∆m f [xi, ..., xi+m]p = (xi+m − xi)∆mF[xi, ..., xi+m]p xi+m(cid:90) xi 1 xi+m − xi F(m)(t)p dt ≤ (xi+m − xi) · 1 = ((m − 1)!)p 1 ((m − 1)!)p · xi+m(cid:90) n−m(cid:88) F(m)(t)p dt . xi+m(cid:90) xi 1 ((m − 1)!)p i=0 xi xn(cid:90) x0 (xi+m − xi)∆m f [xi, ..., xi+m]p ≤ F(m)(t)p dt . Clearly, the covering multiplicity of the family {(xi, xi+m) : i = 0, ..., n − m} of open intervals is bounded by m, so that (xi+m − xi)∆m f [xi, ..., xi+m]p ≤ m ((m − 1)!)p F(m)(t)p dt . Hence, n−m(cid:88) i=0 n−m(cid:88) i=0 Hence, n−m(cid:88) i=0 (xi+m − xi)∆m f [xi, ..., xi+m]p ≤ m ((m − 1)!)p (cid:107)F(cid:107)p p (R) ≤ 2p (cid:107)F(cid:107)p Lm p (R) . Lm (2.15) Taking the supremum in the left hand side of this inequality over all (n + 1)-point subsets {x0, ..., xn} ⊂ E with n ≥ m, we obtain the following: Lm,p( f : E) ≤ 2(cid:107)F(cid:107)Lm p (R) . See (1.14). Finally, taking the infimum in the right hand side of this inequality over all functions F ∈ Lm p (R)E proving the necessity part of Theorem 1.7. p (R) such that FE = f , we obtain that Lm,p( f : E) ≤ 2(cid:107) f(cid:107)Lm (cid:3) 12 (Theorem 1.8: Necessity) Let f ∈ Lm p (R) be an arbitrary function such that FE = f . Let S = {x0, ..., xm}, x0 < ... < xm, be a subset of E and let x ∈ R. Thanks to (2.3) and (2.11), p (R)E where p ∈ (1,∞), and let F ∈ Lm ∆m−1 f [x0, ..., xm−1] − ∆m−1 f [x1, ..., xm] ∆m−1F[x0, ..., xm−1] − ∆m−1F[x1, ..., xm] x − x0 + x − xm xm(cid:90) x − x0 + x − xm 1 (m − 1)! (x − x0 + x − xm) F(m)(t) dt . x0 = ≤ ∆mF[x0, ..., xm] (xm − x0) x − x0 + x − xm = Let I be the smallest closed interval containing S and x. Clearly, I ≤ x − x0 + x − xm and I ⊃ [x0, xm]. Hence, ∆m−1 f [x0, ..., xm−1] − ∆m−1 f [x1, ..., xm] x − x0 + x − xm ≤ 1 (m − 1)! 1 I F(m)(t) dt ≤ M[F(m)](x) . (cid:90) I (Recall that M denotes the Hardy-Littlewood maximal function, see (2.2).) Taking the supremum in the left hand side of this inequality over all subsets S = {x0, ..., xm} ⊂ E, x0 < ... < xm, we obtain that See (1.16). Hence, (∆m f )(cid:93) E (x) ≤ M[F(m)](x), x ∈ R . E (cid:107)Lp(R) ≤ (cid:107)M[F(m)](cid:107)Lp(R) (cid:107) (∆m f )(cid:93) so that, by the Hardy-Littlewood maximal theorem, (cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) ≤ C(p)(cid:107)F(m)(cid:107)Lp(R) = C(p)(cid:107)F(cid:107)Lm p (R) . Taking the infimum in the right hand side of this inequality over all functions F ∈ Lm that FE = f , we finally obtain the required inequality p (R) such E (cid:107)Lp(R) ≤ C(p)(cid:107) f(cid:107)Lm The proof of the necessity part of Theorem 1.8 is complete. (cid:107) (∆m f )(cid:93) p (R)E . (cid:3) 3. The Whitney extension method in R and traces of Sobolev functions. In this section we prove the sufficiency part of Theorem 1.8. Given a function F ∈ Cm(R) and x ∈ R, we let m(cid:88) k=0 T m x [F](y) = F(k)(x)(y − x)k, y ∈ R, 1 k! denote the Taylor polynomial of F of degree m at x. Let E be a closed subset of R and let P = {Px : x ∈ E} be a family of polynomials of degree at most m indexed by points of E. (Thus Px ∈ Pm for every x ∈ E.) Following [29], we refer to P as a Whitney m-field defined on E. 13 We say that a function F ∈ Cm(R) agrees with the Whitney m-field P = {Px : x ∈ E} on E, if x [F] = Px for each x ∈ E. In that case we also refer to P as the Whitney m-field on E generated T m by F or as the m-jet generated by F. We define the Lm p (R), T m−1 p -"norm" of the m-jet P = {Px : x ∈ E} by [F] = Px for every x ∈ E (3.1) We prove the sufficiency part of Theorem 1.8 in two steps. At the first step, given m ∈ N we construct a linear operator which to every function f on E assigns a certain Whitney (m − 1)-field (cid:110)(cid:107)F(cid:107)Lm (cid:107)P(cid:107)m,p,E = inf (cid:111) . p (R) : F ∈ Lm x P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} such that Px(x) = f (x) for all x ∈ E. We produce P(m,E)[ f ] by a slight modification of Whitney's extension construction [67]. See also [34, 35, 43, 48] where similar constructions have been used for characterization of traces of Lm∞(R)-functions. At the second step of the proof we show that for every p ∈ (1,∞) and every function f : E → R such that (∆m f )(cid:93) E ∈ Lp(R) (see (1.16)) the following inequality (cid:107)P(m,E)[ f ](cid:107)m,p,E ≤ C(m, p)(cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) (3.2) holds. One of the main ingredients of the proof of (3.2) is a trace criterion for jets generated by Sobolev functions. See Theorem 3.17 below. 3.1. Interpolation knots and their properties. Let E ⊂ R be a closed subset, and let k be a non-negative integer, k ≤ #E. Following [67] (see also [43, 48]), given x ∈ E we construct an important ingredient of our extension procedure, a finite set Yk(x) ⊂ E which, in a certain sense, is "well concentrated" around x. This set provides interpolation knots for Lagrange and Hermite polynomials which we use in our modification of the Whitney extension method. We will need the following notion. Let A be a nonempty finite subset of E containing at most one limit point of E. We assign to A a point aE(A) ∈ E in the closure of E \ A having minimal distance to A. More specifically: (i) If A does not contain limit points of E, the set E \ A is closed, so that in this case aE(A) is a point nearest to A on E \ A. Clearly, in this case aE(A) (cid:60) A; (ii) Suppose there exists a (unique) point a ∈ A which is a limit point of E. In this case we set aE(A) = a. Note that in both cases dist(aE(A), A) = dist(A, E \ A). Now let us construct a family of points {y0(x), y1(x), ..., ynk(x)} in E, 0 ≤ nk(x) ≤ k, using the following inductive procedure. First, we put y0(x) = x and Y0(x) = {y0(x)}. If k = 0, we put nk(x) = 0, and stop. Suppose that k > 0. If y0(x) = x is a limit point of E, we again put nk(x) = 0, and stop. If y0(x) We define a point y1(x) ∈ E by y1(x) = aE(Y0(x)), and set Y1(x) = {y0(x), y1(x)}. If k = 1 or y1(x) is an isolated point of E, we continue the procedure. is a limit point of E, we put nk(x) = 1, and stop. Let k > 1 and y1(x) is an isolated point of E. In this case we put y2(x) = aE(Y1(x)) and Y2(x) = {y0(x), y1(x), y2(x)}. 14 If k = 2 or y2(x) is a limit point of E, we set nk(x) = 2, and stop. But if k > 2 and y2(x) is an isolated point of E, we continue the procedure and define y3, etc. At the j-th step of this algorithm we obtain a j + 1-point set Y j(x) = {y0(x), ..., y j(x)}. If j = k or y j(x) is a limit point of E, we put nk(x) = j and stop. But if j < k and y j(x) is an isolated point of E, we define a point y j+1(x) and a set Y j+1(x) by the formulae y j+1(x) = aE(Y j(x)) and Y j+1(x) = {y0(x), ..., y j(x), y j+1(x)} . (3.3) (3.4) Clearly, for a certain n = nk(x), 0 ≤ n ≤ k, the procedure stops . This means that either n = k or, whenever n < k, the points y0(x), ..., yn−1(x) are isolated points of E, but We also introduce points y j(x) and sets Y j(x) for nk(x) ≤ j ≤ k by letting yn(x) is a limit point of E . (3.5) (3.6) Note that, given x ∈ E the definitions of points y j(x) and the sets Y j(x) do not depend on k, i.e, y j(x) is the same point and Y j(x) is the same set for every k ≥ j. This is immediate from (3.6). y j(x) = ynk(x)(x) and Y j(x) = Ynk(x)(x). In the next three lemmas, we describe several important properties of the points y j(x) and the sets Y j(x). Lemma 3.1 Given x ∈ E, the points y j(x) and the sets Y j(x), 0 ≤ j ≤ k, have the following properties: (a). y0(x) = x and y j(x) = aE(Y j−1(x)) for every 1 ≤ j ≤ k; (b). Let n = nk(x) ≥ 1. Then y0(x), ..., yn−1(x) are isolated points of E. Furthermore, if y ∈ Yn(x) and y is a limit point of E, then y = yn(x). In addition, if 0 < n < k, (c). #Y j(x) = min{ j, nk(x)} + 1 for every 0 ≤ j ≤ k; (d). For every j = 0, ..., k the following equality then yn(x) is a limit point of E; [min Y j(x), max Y j(x)] ∩ E = Y j(x) (3.7) holds. Furthermore, the point y j(x) is either minimal or maximal point of the set Y j(x). Proof. Properties (b)-(d) are immediate from the definitions of the points y j(x) and the sets Y j(x). Let us prove (a). We know that y0(x) = x and, by (3.3), y j(x) = aE(Y j−1(x)) for every j = 1, ..., nk(x). If nk(x) < j ≤ k, then, by (3.6), Y j(x) = Ynk(x)(x) for every j, nk(x) ≤ j ≤ k. On the other hand, since nk(x) < k, the point ynk(x) is a unique limit point of E. See (3.5). Hence, by definition of aE and (3.6), for every j, nk(x) < j ≤ k, proving property (a) in the case under consideration. (cid:3) aE(Y j−1(x)) = aE(Ynk(x)(x)) = ynk(x) = y j(x) 15 Lemma 3.2 Let x1, x2 ∈ E and let 0 ≤ j ≤ k. If x1 ≤ x2 then min Y j(x1) ≤ min Y j(x2) (3.8) and max Y j(x1) ≤ max Y j(x2) . (3.9) Proof. We proceed by induction on j. Since Y0(x1) = {x1} and Y0(x2) = {x2}, inequalities (3.8) Suppose that these inequalities hold for some j, 0 ≤ j ≤ k − 1. Let us prove that and (3.9) hold for j = 0. min Y j+1(x1) ≤ min Y j+1(x2) and max Y j+1(x1) ≤ max Y j+1(x2) . (3.10) (3.11) First we prove (3.10). Recall that, by (3.3), for each (cid:96) = 1, 2, we have y j+1(x(cid:96)) = aE(Y j(x(cid:96))) and (3.12) If Y j(x2) contains a limit point of E, then y j+1(x2) ∈ Y j(x2) so that Y j+1(x2) = Y j(x2). This Y j+1(x(cid:96)) = Y j(x(cid:96)) ∪ {y j+1(x(cid:96))} . equality and the assumption (3.8) imply that min Y j+1(x1) ≤ min Y j(x1) ≤ min Y j(x2) = min Y j+1(x2) proving (3.10) in the case under consideration. Now suppose that all points of Y j(x2) are isolated points of E. In particular, this implies that 0 ≤ j ≤ nk(x2), see part (b) of Lemma 3.1 and definitions (3.5) and (3.6). Hence, by part (c) of Lemma 3.1, #Y j(x2) = j + 1. Consider two cases. First we assume that (3.13) Then for each point a ∈ R nearest to Y j(x2) on the set E \ Y j(x2) we have a ≥ min Y j(x1). This inequality, definition of aE and (3.3) tell us that min Y j(x1) < min Y j(x2) . aE(Y j(x2)) = y j+1(x2) ≥ min Y j(x1) . Combining this inequality with (3.12) and (3.13), we obtain the required inequality (3.10). Now prove (3.10) whenever min Y j(x1) = min Y j(x2). This equality and inequality (3.9) imply the following inclusion: Y j(x1) ⊂ I = [min Y j(x2), max Y j(x2)] . In turn, equality (3.7) tells us that I ∩ E = Y j(x2) proving that Y j(x1) ⊂ Y j(x2). Recall that in the case under consideration all points of Y j(x2) are isolated points of E. Therefore, all points of Y j(x1) are isolated points of E as well. Now, using the same argument as for the set Y j(x2), we conclude that Thus Y j(x1) ⊂ Y j(x2) and #Y j(x1) = #Y j(x2) proving that Y j(x1) = Y j(x2). This equality implies #Y j(x1) = j + 1 = #Y j(x2). (3.10) completing the proof of inequality (3.10). In the same fashion we prove inequality (3.11). The proof of the lemma is complete. (cid:3) 16 Lemma 3.3 ( [43, p. 231]) Let x1, x2 ∈ E, and let Yk(x1) (cid:44) Yk(x2). Then for all 0 ≤ i, j ≤ k the following inequality max{yi(x1) − y j(x1),yi(x2) − y j(x2)} ≤ max{i, j}x1 − x2 holds. This lemma implies the following Corollary 3.4 For every x1, x2 ∈ E such that Yk(x1) (cid:44) Yk(x2) the following inequality diam Yk(x1) + diam Yk(x2) ≤ 2 k x1 − x2 holds. 3.2. Lagrange polynomials and divided differences at interpolation knots. In this section we describe main properties of the Lagrange polynomials on finite subsets of the set E. Lemma 3.5 Let k be a nonnegative integer, and let P ∈ Pk. Suppose that P has k real distinct roots which lie in a set S ⊂ R. Let I ⊂ R be a closed interval. Then for every i, 0 ≤ i ≤ k, the following inequality P(i) ≤ (diam(I ∪ S ))k−i P(k) max I holds. Proof. Let x j, j = 1, ..., k, be the roots of P, and let X = {x1, ..., xk}. By the lemma's hypothesis, X ⊂ S . Clearly, so that for every i, 0 ≤ i ≤ k, P(x) = P(k) k! k(cid:89) (x − xi), (cid:88) (cid:89) i=1 x ∈ R, P(i)(x) = i! k! P(k) X(cid:48)⊂X, #X(cid:48)=k−i y∈X(cid:48) (x − y), x ∈ R. Hence, max I proving the lemma. P(i) ≤ i! k! (cid:3) k! i!(k − i)! (diam(I ∪ X))k−i P(k) ≤ (diam(I ∪ S ))k−i P(k) We recall that, given S ⊂ R with #S = k + 1 and a function f : S → R, by LS [ f ] we denote the Lagrange polynomial of degree at most k interpolating f on S . Lemma 3.6 Let S 1, S 2 ⊂ R, S 1 (cid:44) S 2, and let #S 1 = #S 2 = k + 1 where k is a nonnegative integer. Let I ⊂ R be a closed interval. Then for every function f : S 1 ∪ S 2 → R and every i, 0 ≤ i ≤ k, L(i) S 1[ f ] − L(i) S 2[ f ] ≤ (k + 1)! (diam(I ∪ S 1 ∪ S 2))k−i A (3.14) max I where ∆k+1 f [S (cid:48)] diam S (cid:48) ; A = max S (cid:48)⊂S 1∪S 2 #S (cid:48)=k+2 17 n−1(cid:88) i=0 Proof. Let n = k + 1 − #(S 1 ∩ S 2); then n ≥ 1 because S 1 (cid:44) S 2. Let {Y j : j = 0, ..., n} be a family of (k + 1)-point subsets of S such that Y0 = S 1, Yn = S 2, and #(Y j ∩ Y j+1) = k for every j = 0, ..., n − 1. Let P j = LY j[ f ], j = 0, ..., n. Then L(i) S 1[ f ] − L(i) S 2[ f ] = max I max I P(i) 0 − P(i) n ≤ P(i) j − P(i) j+1 . max I (3.15) Note that each point y ∈ Y j ∩ Y j+1 is a root of the polynomial P j − P j+1 ∈ Pk. Thus, if the polynomial P j − P j+1 is not identically 0, it has precisely k distinct real roots which belong to the set S 1∪ S 2. We apply Lemma 3.5, taking P = P j− P j+1 and S = S 1∪ S 2, and obtain the following: (3.16) j+1 ≤ (diam(I ∪ S 1 ∪ S 2))k−i P(k) j − P(k) j+1. P(i) j − P(i) max I Thanks to (2.5), j+1 = L(k) which together with (2.3) imply that j − P(k) P(k) Y j [ f ] − L(k) Y j+1[ f ] = k!∆k f [Y j] − ∆k f [Y j+1] j − P(k) P(k) j+1 ≤ k!∆k+1 f [Y j ∪ Y j+1] diam(Y j ∪ Y j+1) ≤ k! max S (cid:48)⊂S 1∪S 2 #S (cid:48)=k+2 ∆k+1 f [S (cid:48)] diam S (cid:48) = k! A . This inequality, (3.15) and (3.16) together imply the required inequality (3.14) proving the (cid:3) lemma. Lemma 3.7 Let k be a nonnegative integer, (cid:96) ∈ N, k < (cid:96), and let Y = {y j}(cid:96) increasing sequence in R. Let I = [y0, y(cid:96)], S 1 = {y0, ..., yk}, S 2 = {yl−k, ..., y(cid:96)}, and let j=0 be a strictly S ( j) = {y j, ..., yk+ j+1}, j = 0, ..., (cid:96) − k − 1. (3.17) : Y → R, every i, 0 ≤ i ≤ k, and every p ∈ [1,∞) the following Then for every function f inequality (cid:96)−k−1(cid:88) j=0 (cid:96)−k ≤ (cid:96)−k−1(cid:88) i=0 L(i) S 1[ f ] − L(i) S 2[ f ]p ≤ ((k + 2)!)p (diam I)(k−i+1)p−1 max I holds. ∆k+1 f [S ( j)]p (diam S ( j)) (3.18) Proof. Let Y j = {y j, ..., y j+k}, j = 0, ..., (cid:96) − k, so that S 1 = Y0, S 2 = Y(cid:96)−k, and S ( j) = Y j ∪ Y j+1, j = 0, ..., (cid:96) − k − 1. Let P j = LY j[ f ], j = 0, ..., (cid:96) − k. Then S 1[ f ] − L(i) L(i) S 2[ f ] = max I max I P(i) 0 − P(i) j − P(i) P(i) j+1 . max I (3.19) Note that every y ∈ Y j ∩ Y j+1 is a root of the polynomial P j − P j+1 ∈ Pk. Thus, if the polynomial P j − P j+1 is not identically 0, it has precisely k distinct real roots on I. Then, by Lemma 3.5, P(i) j − P(i) j+1 ≤ (diam I)k−i P(k) j − P(k) j+1. max I (3.20) 18 (cid:96)−k−1(cid:88) j=0 j=0 (cid:96)−k−1(cid:88) (cid:96)−k−1(cid:88) (cid:96)−k−1(cid:88) j=0 j=0 p p−1 (cid:96)−k−1(cid:88) (cid:96)−k−1(cid:88) j=0 Thanks to (2.5), P(k) j − P(k) j+1 = L(k) Y j [ f ] − L(k) Y j+1[ f ] = k!∆k f [Y j] − ∆k f [Y j+1] so that, by (2.3), P(k) j − P(k) j+1 = k!∆k+1 f [Y j ∪ Y j+1] diam(Y j ∪ Y j+1) = k!∆k+1 f [S ( j)] diam S ( j) . This inequality, (3.19) and (3.20) together imply that L(i) S 1[ f ] − L(i) S 2[ f ] ≤ k! (diam I)k−i max I ∆k+1 f [S ( j)] diam S ( j) . (3.21) Let I j = [y j, yk+ j+1]. Then, thanks to (3.17), diam I j = diam S j = yk+ j+1 − y j. Furthermore, since j=0 is a strictly increasing sequence and #S j = k + 2, the covering multiplicity of the family {y j}(cid:96) {I j : j = 0, ..., (cid:96) − k − 1} is bounded by k + 2. Hence, (cid:96)−k−1(cid:88) (cid:96)−k−1(cid:88) diam S ( j) = diam I j = j=0 j=0 I j ≤ (k + 2)I = (k + 2) diam I . Finally, this inequality, the Holder inequality and (3.21) together imply that L(i) S 1[ f ] − L(i) S 2[ f ]p ≤ (k!)p (diam I)(k−i)p max I ∆k+1 f [S ( j)] diam S ( j) ≤ (k!)p (diam I)(k−i)p diam S ( j) ∆k+1 f [S ( j)]p diam S ( j) ≤ (k!)p(k + 2)p−1 (diam I)(k−i+1)p−1 ∆k+1 f [S ( j)]p diam S ( j) proving inequality (3.18). Lemma 3.8 Let k be a nonnegative integer and let 1 < p < ∞. Let f be a function defined on a closed set E ⊂ R with #E > k + 1. Suppose that (cid:3) j=0 λ = sup S⊂E, #S =k+2 ∆k+1 f [S ] (diam S ) 1 p < ∞ . Then for every limit point x of E and every i, 0 ≤ i ≤ k, there exists a limit Recall that the notation S → x means diam(S ∪ {x}) → 0 (see (2.1)). S→x, S⊂E, #S =k+1 Furthermore, let Px ∈ Pk be a polynomial such that fi(x) = lim L(i) S [ f ](x) . (3.22) (3.23) (3.24) Then for every δ > 0 and every set S ⊂ E such that #S = k + 1 and diam(S ∪{x}) < δ the following inequality P(i) x (x) = fi(x) for every i, 0 ≤ i ≤ k . P(i) x − L(i) S [ f ] ≤ C(k) λ δk+1−i−1/p, 0 ≤ i ≤ k, (3.25) max [x−δ,x+δ] holds. 19 Proof. Let δ > 0 and let S 1, S 2 be two subsets of E such that #S j = k + 1 and diam(S j∪{x}) < δ, j = 1, 2. Hence, S = S 1 ∪ S 2 ⊂ I = [x − δ, x + δ]. We apply Lemma 3.6 and obtain that ∆k+1 f [S (cid:48)] diam S (cid:48) S 2[ f ](x) ≤ (k + 1)! (diam I)k−i L(i) S 1[ f ](x) − L(i) max . S (cid:48)⊂S , #S (cid:48)=k+2 Thanks to (3.22), for every (k + 2)-point subset S (cid:48) ⊂ E, so that ∆k+1 f [S (cid:48)] ≤ λ (diam S (cid:48))− 1 p L(i) S 1[ f ](x) − L(i) S 2[ f ](x) ≤ (k + 1)! λ (2δ)k−i max S (cid:48)⊂S , #S (cid:48)=k+2 ≤ (k + 1)! λ (2δ)k−i (2δ)1−1/p. Hence, L(i) S 1[ f ](x) − L(i) S 2[ f ](x) ≤ C(k) λ δk+1−i−1/p . (diam S (cid:48))1−1/p (3.26) (3.27) Since p > 1, S 2[ f ](x) → 0 as δ → 0 L(i) S 1[ f ](x) − L(i) Let us prove inequality (3.25). Thanks to (3.27), for every two sets S ,(cid:101)S ∈ E, with #S = #(cid:101)S = proving the existence of the limit in (3.23). k + 1 such that diam(S ∪ {x}), diam(S ∪ {x}) < δ, the following inequality holds. Passing to the limit in this inequality whenever the set(cid:101)S → x (i.e., diam((cid:101)S ∪ {x}) → 0) we S [ f ](x) ≤ C(k) λ δk+1−i−1/p [ f ](x) − L(i) L(i)(cid:101)S obtain the following: S [ f ](x) ≤ C(k) λ δk+1−i−1/p . See (3.23) and (3.24). Therefore, for each y ∈ [x − δ, x + δ], we have P(i) x (x) − L(i) P(i) x (y) − L(i) S [ f ](y) = (P(i+ j) (x) − L(i+ j) S [ f ](x)) (y − x) j P(i+ j) x (x) − L(i+ j) S [ f ](x) δ j 1 j! ≤ C(k) λ δk+1−i− j−1/p δ j ≤ C(k) λ δk+1−i−1/p . The proof of the lemma is complete. (cid:3) j=0 Lemma 3.9 Let k, p, E, f, λ and x be as in the statement of Lemma 3.8. Then for every i, 0 ≤ i ≤ k, lim S→x, S⊂E, #S =i+1 i! ∆i f [S ] = fi(x) . Proof. Let δ > 0 and let S ⊂ E be a finite set such that #S = i + 1 and diam({x}, S ) < δ. Since x is a limit point of E, there exists a set Y ⊂ E ∩ [x − δ, x + δ] with #Y = k + 1 such that S ⊂ Y. Then, thanks to (3.25), P(i) x − L(i) Y [ f ] ≤ C(k) λ δk+1−i−1/p . max [x−δ,x+δ] (3.28) 20 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k−i(cid:88) j=0 1 j! x k−i(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ k−i(cid:88) j=0 Since the Lagrange polynomial LY[ f ] interpolates f on S , we have ∆i f [S ] = ∆i(LY[ f ])[S ] so that, by (2.6), there exists ξ ∈ [x − δ, x + δ] such that i! ∆i f [S ] = L(i) Y [ f ](ξ). This equality and (3.28) imply that P(i) x (ξ) − i!∆i f [S ] = P(i) x (ξ) − L(i) Y [ f ](ξ) ≤ C(k) λ δk+1−i−1/p. Hence, fi(x) − i!∆i f [S ] = P(i) ≤ P(i) x (x) − i!∆i f [S ] ≤ P(i) x (x) − P(i) x (x) − P(i) x (ξ) + C(k) λ δk+1−i−1/p. x (ξ) + P(i) x (ξ) − i!∆i f [S ] Since P(i) x (cid:3) is a continuous function and p > 1, the right hand side of this inequality tends to 0 as δ → 0 proving the lemma. Lemma 3.10 Let p ∈ (1,∞), k ∈ N, and let f be a function defined on a closed set E ⊂ R with #E > k + 1. Suppose that f satisfies condition (3.22). Let x ∈ E be a limit point of E, and let S be a subset of E with #S ≤ k containing x. Then for every i, 0 ≤ i ≤ k + 1 − #S , lim S (cid:48)\S→x S⊂S (cid:48)⊂E, #S (cid:48)=k+1 L(i) S (cid:48)[ f ](x) = fi(x) . Proof. For S = {x} the statement of the lemma follows from Lemma 3.8. Suppose that #S > 1. Let I0 = [x − 1/2, x + 1/2] so that diam I0 = 1. We prove that for every i, 0 ≤ i ≤ k − 1, the family of functions {L(i+1) Y [ f ] : Y ⊂ I0 ∩ E, #Y = k + 1} is uniformly bounded on I0 provided condition (3.22) holds. Indeed, fix a subset Y0 ⊂ I0 ∩ E with #Y0 = k + 1. Then for arbitrary Y ⊂ I0 ∩ E, Y (cid:44) Y0, with #Y = k + 1, by Lemma 3.6 and (3.14), L(i+1) Y [ f ] − L(i+1) Y0 max I0 [ f ] ≤ (k + 1)! (diam(I0 ∪ Y ∪ Y0))k−i−1 A = (k + 1)! A where ∆k+1 f [S (cid:48)] diam S (cid:48) . A = max S (cid:48)⊂Y0∪Y #S (cid:48)=k+2 Therefore, thanks to (3.26), [ f ] − L(i+1) L(i+1) max Y0 I0 Y [ f ] ≤ (k + 1)!λ max S (cid:48)⊂Y∪Y0, #S (cid:48)=k+2 (diam S (cid:48))1−1/p ≤ (k + 1)!λ . We apply this inequality to an arbitrary set Y ⊂ I0∩E with #Y = k+1 and to every i, 0 ≤ i ≤ k−1, and obtain that where L(i+1) Y [ f ] ≤ Bi max I0 (3.29) Bi = max I0 L(i+1) Y0 [ f ] + (k + 1)!λ. Fix ε > 0. By Lemma 3.9, there exists δ ∈ (0, 1/2] such that for an arbitrary set V ⊂ E, with diam({x}, V) < δ and #V = i + 1, the following inequality i! ∆i f [V] − fi(x) ≤ ε/2 (3.30) 21 holds. Let S (cid:48) be an arbitrary subset of E such that S ⊂ S (cid:48), #S (cid:48) = k + 1, and (3.31) Recall that #S (cid:48) − #S = k + 1 − #S ≥ i, so that there exists a subset V ⊂ (S (cid:48) \ S ) ∪ {x} with diam({x}, S (cid:48) \ S ) < δ = min{δ, ε/(2Bi)} . #V = i + 1. Thanks to (2.6), there exists ξ ∈ [x − δ, x + δ] such that On the other hand, since the polynomial LS (cid:48)[ f ] interpolates f on V, we have i! ∆i(LS (cid:48)[ f ])[V] = L(i) S (cid:48)[ f ](ξ). ∆i f [V] = ∆i(LS (cid:48)[ f ])[V] so that, i!∆i f [V] = L(i) S (cid:48)[ f ](ξ). This and (3.30) imply that L(i) S (cid:48)[ f ](ξ) − fi(x) ≤ ε/2 . Clearly, thanks to (3.31) and (3.29), (cid:32) (cid:33) [ f ] S (cid:48)[ f ](ξ) − L(i) L(i) Finally, we obtain that S (cid:48)[ f ](x) ≤ fi(x) − L(i) proving the lemma. (cid:3) max [x−1/2,x+1/2] L(i+1) S (cid:48) · x − ξ ≤ Bi δ ≤ Bi (ε/(2Bi)) = ε/2 . S (cid:48)[ f ](x) ≤ fi(x) − L(i) S (cid:48)[ f ](ξ) + L(i) S (cid:48)[ f ](ξ) − L(i) S (cid:48)[ f ](x) ≤ ε/2 + ε/2 = ε 3.3. Whitney m-fields and the Hermite polynomials. Let m ∈ N and let p ∈ (1,∞). In this section, given a function f on E satisfying condition (3.32), we construct a certain Whitney (m − 1)-field P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} such that Px(x) = f (x) for all x ∈ E. In the next section we apply to P(m,E)[ f ] a criterion for extensions of Sobolev jets given in E ∈ Theorem 3.17 below. This criterion will enable us to show that f ∈ Lm Lp(R). This will complete the proof of the sufficiency part of Theorem 1.8. p (R)E provided (∆m f )(cid:93) We turn to constructing the Whitney field P(m,E)[ f ]. Everywhere in this section we assume that the function f satisfies the following condition: sup Let k = m − 1. Given x ∈ E, let S⊂E, #S =m+1 ∆m f [S ] (diam S ) 1 p < ∞ . Sx = Yk(x) = {y0(x), ..., ynk(x)(x)} and let sx = ynk(x) . (3.32) (3.33) (3.34) We recall that the points y j(x) and the sets Y j(x) are defined by formulae (3.3)-(3.6). The next two propositions describe the main properties of the sets {Sx : x ∈ E} and the points {sx : x ∈ E}. These properties are immediate from Lemmas 3.1, 3.2 and Corollary 3.4. 22 Proposition 3.11 (i) x ∈ Sx and #Sx ≤ m for every x ∈ E. Furthermore, [min Sx , max Sx] ∩ E = Sx ; (ii) For every x1, x2 ∈ E such that Sx1 diam Sx1 (cid:44) Sx2 the following inequality + diam Sx2 ≤ 2 mx1 − x2 holds; (iii). If x1, x2 ∈ E and x1 < x2 then min Sx1 ≤ min Sx2 and max Sx1 ≤ max Sx2. Proposition 3.12 (i) The point (3.35) (3.36) sx belongs to Sx (3.37) for every x ∈ E. This point is either minimal or maximal point of the set Sx . (ii) All points of the set Sx \ {sx} are isolated points of E provided #S x > 1. If y ∈ Sx and y is a limit point of E, then y = sx; (iii) If #Sx < m then sx is a limit point of E. Remark 3.13 Let E = {xi}(cid:96)2 where (cid:96)1, (cid:96)2 ∈ Z ∪ {−∞, +∞}, (cid:96)1 + m ≤ (cid:96)2, be a strictly increasing sequence of points in R. In this case, for each i ∈ Z, (cid:96)1 ≤ i ≤ (cid:96)2, the set Sxi consists of m consecutive elements of the sequence E. In other words, there exists ν ∈ Z, (cid:96)1 ≤ ν ≤ (cid:96)2, such that i=(cid:96)1 = {xν, ..., xν+m−1} . Sxi (3.38) Indeed, let k = m − 1. Since all points of E are isolated, nk(x) = k for every x ∈ E. See (3.4). In Thus, thanks to (3.33), Sx = Yk(x) = {y0(x), ..., yk(x)} so that #Sx = k + 1 = m. On the other hand, particular, in this case sx = yk(x). by (3.35), Sx = [min Sx, max Sx] ∩ E proving (3.38). Definition 3.14 Given a function f : E → R satisfying condition (3.32), we define the Whitney (m − 1)-field P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} as follows: (3.32) and Lemma 3.8, for every i, 0 ≤ i ≤ m − 1, there exists a limit (i) If #Sx < m, part (iii) of Proposition 3.12 tells us that sx is a limit point of E. Then, thanks to (cid:67) fi(sx) = lim S→sx S⊂E, #S =m L(i) S [ f ](sx) . (3.39) We define a polynomial Px ∈ Pm−1 as the Hermite polynomial satisfying the following condi- tions: and Px(y) = f (y) for every y ∈ Sx, P(i) x (sx) = fi(sx) for every i, 1 ≤ i ≤ m − #Sx . (ii) If #Sx = m, we put Px = LSx[ f ]. 23 (3.40) (3.41) (3.42) The next lemma shows that the Whitney (m−1)-field P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} determined by Definition 3.14 is well defined. Lemma 3.15 For each x ∈ E there exists the unique polynomial Px satisfying conditions (3.40) and (3.41) provided condition (3.32) holds. Proof. In case (i) (#Sx < m) the existence and uniqueness of Px satisfying (3.40) and (3.41) is immediate from [2, Ch. 2, Section 11]. See also formula (3.45) below. (cid:3) Clearly, in case (ii) (#Sx = m) the property (3.40) holds as well, and (3.41) holds vacuously. We also note that x ∈ Sx and Px = f on Sx (see (3.40)) which imply that Px(x) = f (x) for every x ∈ E . (3.43) For the case #Sx < m, m > 1, we present an explicit formula for the Hermite polynomials Px, x ∈ E, from Definition 3.14. This formula follows from general properties of the Hermite polynomials given in [2, Ch. 2, Section 11]. Let n = #Sx − 1 and let yi = yi(x), i = 0, ..., n, so that Sx = {y0, ..., yn}. See (3.33). (Note also that In this case the Hermite polynomial Px satisfying (3.40) and (3.41) can be represented as a linear in these settings sx = yn.) combination of polynomials H0, ..., Hn,(cid:101)H1, ....,(cid:101)Hm−n−1 ∈ Pm−1 which are uniquely determined by the following conditions: i (yn) = 0 for every i, 0 ≤ i ≤ n . (i) Hi(yi) = 1 for every i, 0 ≤ i ≤ n, and H j(yi) = 0 for every 0 ≤ i, j ≤ n, i (cid:44) j, and (ii) (cid:101)H j(yi) = 0 for every 0 ≤ i ≤ n, 1 ≤ j ≤ m − n − 1, and for every 1 ≤ j ≤ m − n − 1, j (yn) = 0 for every (cid:96), 1 ≤ (cid:96) ≤ m − n − 1, (cid:96) (cid:44) j . (cid:101)H j, 0 ≤ i ≤ n, 1 ≤ j ≤ m − n − 1, i (yn) = ... = H(m−n−1) H(cid:48) and (cid:101)H((cid:96)) The existence and uniqueness of the polynomials j (yn) = 1 (cid:101)H( j) Hi and are proven in [2, Ch. 2, Section 11]. It is also shown in [2] that for every P ∈ Pm−1 the following unique representation n(cid:88) n(cid:88) i=0 m−n−1(cid:88) P( j)(yn)(cid:101)H j(y), m−n−1(cid:88) f j(yn)(cid:101)H j(y), j=1 y ∈ R, y ∈ R . (3.44) (3.45) P(y) = holds. In particular, P(yi) Hi(y) + Px(y) = f (yi) Hi(y) + i=0 j=1 24 Clearly, Px meets conditions (3.40) and (3.41). Let I ⊂ R be a bounded closed interval, and let Cm(I) be the space of all m-times continuously differentiable functions on I. We norm Cm(I) by m(cid:88) i=0 (cid:107) f(cid:107)Cm(I) = f (i) . max I We will need the following important property of the polynomials {Px : x ∈ E}. Lemma 3.16 Let f be a function defined on a closed set E ⊂ R with #E > m + 1, and satisfying condition (3.32). Let I be a bounded closed interval in R. Then for every x ∈ E (cid:107)LS (cid:48)[ f ] − Px(cid:107)Cm(I) = 0 . lim S (cid:48)\Sx→sx Sx⊂ S (cid:48)⊂E, #S (cid:48)=m Proof. The lemma is obvious whenever #Sx = m because in this case LSx[ f ] = Px. In particular, the lemma is trivial for m = 1. Let now m > 1 and let #Sx < m. In this case the polynomial Px can be represented in the form (3.45). Since in this case sx is a limit point of E (see part (iii) of Proposition 3.12), Lemma 3.10 and (3.41) imply that L(i) S (cid:48)[ f ](sx) = fi(sx) = P(i) x (sx) for every 1 ≤ i ≤ m − n − 1 . (3.46) lim S (cid:48)\Sx→sx Sx⊂ S (cid:48)⊂E, #S (cid:48)=m Let n = #Sx − 1 and let Sx = {y0, ..., yn} where yi = yi(x), i = 0, ..., n. Then, thanks to (3.44), for every set S (cid:48) ⊂ E with #S (cid:48) = m such that Sx ⊂ S (cid:48), the polynomial LS (cid:48)[ f ] has the following representation: LS (cid:48)[ f ](y) = f (yi) Hi(y) + S (cid:48) [ f ](sx)(cid:101)H j(y), L( j) y ∈ R . This representation, (3.45) and (3.46) imply that n(cid:88) i=0 j=1 m−n−1(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)m−n−1(cid:88) m−n−1(cid:88) (cid:12)(cid:12)(cid:12)L( j) j=1 I j=1 LS (cid:48)[ f ] − Px = max max I S (cid:48) [ f ](sx) − P( j) (L( j) ≤ S (cid:48) [ f ](sx) − P( j) x (sx) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) x (sx))(cid:101)H j (cid:12)(cid:12)(cid:12) → 0 (cid:12)(cid:12)(cid:12)(cid:101)H j (cid:12)(cid:12)(cid:12) max I as S (cid:48) \ Sx → sx provided Sx ⊂ S (cid:48) ⊂ E and #S (cid:48) = m. Since the uniform norm on I and the Cm(I)-norm are equivalent norms on the finite dimensional space Pm, convergence of LS (cid:48)[ f ] to Px in the uniform norm on I implies convergence of LS (cid:48)[ f ] to Px in the Cm(I)-norm proving the lemma. (cid:3) 25 3.4. Extension criteria in terms of sharp maximal functions: sufficiency. E ∈ Lp(R). See (1.16) and (1.17). Let f be a function on E such that (∆m f )(cid:93) Let us prove that f satisfies condition (3.32). Indeed, let S = {x0, ..., xm} ⊂ E, x0 < ... < xm. Clearly, for every x ∈ [x0, xm], diam({x} ∪ S ) = diam S = xm − x0 so that, thanks to (1.17), ∆m f [S ]p diam S = (xm − x0) ≤ 2p ((∆m f )(cid:93) E (x))p (xm − x0). ∆m f [S ]p(diam S )p (diam({x} ∪ S ))p xm(cid:90) Integrating this inequality (with respect to x) over the interval [x0, xm], we obtain the following: ∆m f [S ]p diam S ≤ 2p ((∆m f )(cid:93) E (x))p dx ≤ 2p (cid:107) (∆m f )(cid:93) E (cid:107)p Lp(R). x0 Hence, sup S⊂E, #S =m+1 ∆m f [S ] (diam S ) 1 p ≤ 2(cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) < ∞ proving (3.32). This condition and Lemma 3.15 guarantee that the Whitney (m − 1)-field P(m,E)[ f ] from Defini- tion 3.14 is well defined. Let us to show that inequality (3.2) holds. Its proof relies on Theorem 3.17 below which provides a criterion for the restrictions of Sobolev jets. For each family P = {Px ∈ Pm−1 : x ∈ E} of polynomials we let P(cid:93) "sharp maximal function" associated with P which is defined by Pa1(x) − Pa2(x) x − a1m + x − a2m , P(cid:93) m,E(x) = a1, a2∈E, a1(cid:44)a2 sup m,E denote a certain kind of a x ∈ R. Theorem 3.17 ( [63]) Let m ∈ N, p ∈ (1,∞), and let E be a closed subset of R. Suppose we are given a family P = {Px : x ∈ E} of polynomials of degree at most m − 1 indexed by points of E. for every x ∈ E if and Then there exists a Cm−1-function F ∈ Lm p (R) such that T m−1 [F] = Px x only if P(cid:93) m,E ∈ Lp(R). Furthermore, (cid:107)P(cid:107)m,p,E ∼ (cid:107)P(cid:93) m,E(cid:107)Lp(R) (3.47) with the constants in this equivalence depending only on m and p. We recall that the quantity (cid:107)P(cid:107)m,p,E is defined by (3.1). Lemma 3.18 Let f be a function on E such that (∆m f )(cid:93) following inequality E ∈ Lp(R). Then for every x ∈ R the (P(m,E)[ f ])(cid:93) m,E(x) ≤ C(m) (∆m f )(cid:93) E (x) (3.48) holds. 26 Proof. Let x ∈ R, a1, a2 ∈ E, a1 (cid:44) a2, and let r = x − a1 + x − a2. Let(cid:101)S j = S a j and let s j = sa j, j = 1, 2. See (3.33) and (3.34). We know that a j, s j ∈(cid:101)S j, j = 1, 2 (see Propositions 3.11 and 3.12). Suppose that(cid:101)S1 (cid:44)(cid:101)S2. Then inequality (3.36) tells us that diam(cid:101)S1 + diam(cid:101)S2 ≤ 2 ma1 − a2 . Fix an ε > 0. Lemma 3.16 produces m-point subsets S j ⊂ E, j = 1, 2, such that(cid:101)S j ⊂ S j, diam({s j} ∪ (S j \(cid:101)S j)) ≤ r (3.50) (3.49) and Recall that s j ∈(cid:101)S j, j = 1, 2, so that, thanks to (3.50), Pa j(x) − LS j[ f ](x) ≤ ε rm/2m+1 . diam S j ≤ diam(cid:101)S j + diam({s j} ∪ (S j \(cid:101)S j)) ≤ diam(cid:101)S j + r. This inequality together with (3.49) imply that (3.51) Let I be the smallest closed interval containing S 1 ∪ S 2 ∪ {x}. Since a j ∈(cid:101)S j ⊂ S j, j = 1, 2, diam S j ≤ 2ma1 − a2 + r, j = 1, 2. (3.52) diam I ≤ x − a1 + x − a2 + diam S1 + diam S2 so that, by (3.52), diam I ≤ x − a1 + x − a2 + 4ma1 − a2 + 2r ≤ (4m + 3) r. (3.53) (Recall that r = x − a1 + x − a2.) This inequality and inequality (3.51) imply the following: Pa1(x) − Pa2(x) ≤ Pa1(x) − LS1[ f ](x) + LS1[ f ](x) − LS2[ f ](x) + Pa2(x) − LS2[ f ](x) = J + εrm/2m where J = LS1[ f ](x) − LS2[ f ](x). Let us estimate J. We may assume that S1 (cid:44) S2; otherwise J = 0. We apply Lemma 3.6 taking k = m − 1 and i = 0, and get J ≤ max LS1[ f ] − LS2[ f ] ≤ m! (diam I)m−1 I max S (cid:48)⊂S , #S (cid:48)=m+1 ∆m f [S (cid:48)] diam S (cid:48) where S = S1 ∪ S2. This inequality together with (3.53) and (1.17) imply that J ≤ C(m) rm−1 (∆m f )(cid:93) E(x) max S (cid:48)⊂S , #S (cid:48)=m+1 diam({x} ∪ S (cid:48)) ≤ C(m) rm (∆m f )(cid:93) E(x) . We are in a position to prove inequality (3.48). We have: Pa1(x) − Pa2(x) x − a1m + x − a2m ≤ 2m r−m Pa1(x) − Pa2(x) ≤ 2m r−m (J + εrm/2m) E(x) + ε = C(m) (∆m f )(cid:93) ≤ C(m)2m r−m rm (∆m f )(cid:93) E(x) + ε . 27 Since ε > 0 is arbitrary, we obtain that Pa1(x) − Pa2(x) x − a1m + x − a2m ≤ C(m) (∆m f )(cid:93) E(x) provided (cid:101)S1 (cid:44) (cid:101)S2. Clearly, this inequality also holds whenever (cid:101)S1 = (cid:101)S2 because in this case = Pa2. Pa1 Taking the supremum in the right hand side of inequality (3.54) over all a1, a2 ∈ E, a1 (cid:44) a2, we (3.54) obtain (3.48). The proof of the lemma is complete. (cid:3) We finish the proof of Theorem 1.8 as follows. Let f be a function on E such that (∆m f )(cid:93) the Whitney (m − 1)-field from Definition 3.14. Lemma 3.18 tells us that E (cid:107)Lp(R) . m,E(cid:107)Lp(R) ≤ C(m)(cid:107) (∆m f )(cid:93) (cid:107)(P(m,E)[ f ])(cid:93) E ∈ Lp(R), and let P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} be This inequality and definition (3.1) imply the existence of a function F ∈ Lm Combining this inequality with equivalence (3.47), we obtain inequality (3.2). T m−1 p (R) such that p (R) ≤ 2(cid:107)P(m,E)[ f ](cid:107)m,p,E ≤ C(m, p)(cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) . (3.55) x [F] = Px on E and (cid:107)F(cid:107)Lm We also note that Px(x) = f (x) on E, see (3.43), so that F(x) = T m−1 x [F](x) = Px(x) = f (x), x ∈ E . Thus F ∈ Lm p (R) and FE = f proving that f ∈ Lm p (R)E. Furthermore, thanks to (1.9) and (3.55), p (R)E ≤ (cid:107)F(cid:107)Lm The proof of Theorem 1.8 is complete. (cid:107) f(cid:107)Lm (cid:3) p (R) ≤ C(m, p)(cid:107) (∆m f )(cid:93) E (cid:107)Lp(R). Remark 3.19 (i) In [63] given a Whitney (m − 1)-field P = {Px : x ∈ E} satisfying conditions of Theorem 3.17, we construct a corresponding extension F of f using the standard Whitney's extension method [66]. (ii) Examining the proof of Lemma 3.18, we conclude that the construction of the Whitney field P(m,E)[ f ] described in Definition 3.14, can be generalized considerably. More specifically, we can modify this definition as follows: x (sx) = fi(sx) = fi(x), i = 0, ..., m − 1.) (i(cid:48)) If #Sx = 1 (i.e., sx = x), we construct Px in the same way as in part (i) of Definition 3.14. (ii(cid:48)) Let 1 < #Sx ≤ m, and let εx = diam Sx(> 0). Fix a constant γ ≥ 1. We pick an m-point set S+ which belongs to the (γεx)-neighborhood of x. (Thus y − x ≤ γεx (Thus, P(i) for every y ∈ S+.) One can easily see that Lemma 3.18 holds after such a modification of the Whitney field P(m,E)[ f ] with the constant C in the right hand side of (3.48) depending on m and γ. This remark shows that there exist a rather wide variety of Whitney-type extension operators providing almost optimal Sobolev extensions of functions defined on closed subsets of R. Each of these operators enables us to prove the trace criterion given in Theorem 1.8. However, only one of these operators enables us to prove the ("stronger") variational trace crite- rion presented in Theorem 1.7, namely, the original extension operator with (m − 1)-jets described in Definition 3.14. See the next section. (cid:67) 28 4. A variational criterion for Sobolev traces. 4.1. A variational criterion for Sobolev jets. Our proof of the sufficiency part of Theorem 1.7 relies on the following extension theorem for Sobolev jets. Theorem 4.1 Let m ∈ N, p ∈ (1,∞), and let E be a closed subset of R. Suppose we are given a Whitney (m − 1)-field P = {Px : x ∈ E} defined on E. Then there exists a Cm−1-function F ∈ Lm p (R) such that T m−1 x [F] = Px for every x ∈ E if and only if the following quantity Nm,p,E(P) = sup  k−1(cid:88) j=1 m−1(cid:88) i=0 P(i) x j (x j) − P(i) x j+1(x j)p (x j+1 − x j)(m−i)p−1 1/p (4.1) (4.2) is finite. Here the supremum is taken over all integers k > 1 and all finite strictly increasing sequences {x j}k Furthermore, j=1 ⊂ E. (cid:107)P(cid:107)m,p,E ∼ Nm,p,E(P) . The constants of equivalence in (4.3) depend only on m. (4.3) Theorem 4.1 is a refinement of Theorem 3.17. Proof. (Necessity.) Let {x j}k j=1 ⊂ E be a strictly increasing sequence. Let P = {Px : x ∈ E} be a Whitney (m− 1)-field on E, and let F ∈ Lm p (R) be a function satisfying condition (4.1). The Taylor formula with the reminder in the integral form tells us that for every x ∈ R and every a ∈ E the following equality F(x) − T m−1 a [F](x) = 1 (m − 1)! F(m)(t) (x − t)m−1 dt holds. Let 0 ≤ i ≤ m−1. Differentiating this equality i times (with respect to x) we obtain the following: x(cid:90) a x(cid:90) a 1 (m − 1 − i)! x(cid:90) a 29 F(i)(x) − (T m−1 a From this and (4.1), we have [F])(i)(x) = F(m)(t) (x − t)m−1−i dt. x (x) − P(i) P(i) a (x) = 1 (m − 1 − i)! F(m)(t) (x − t)m−1−i dt for every x ∈ E. x j (x j) − P(i) P(i) x j+1(x j) = (cid:12)(cid:12)(cid:12)(cid:12)P(i) x j (x j) − P(i) x j+1(x j) 1 x j+1 (m − 1 − i)! ((m − 1 − i)!)p x j(cid:90) (cid:12)(cid:12)(cid:12)(cid:12)p ≤ (x j+1 − x j)(m−1−i)p  x j+1(cid:90) p k−1(cid:88) m−1(cid:88)  m−1(cid:88) F(m)(t) dt ≤ i=0 j=1 = x j i=0 F(m)(t) (x j − t)m−1−i dt  x j+1(cid:90) x j p F(m)(t) dt ≤ 1 ((m − 1 − i)!)p x j+1(cid:90) 1 ((m − 1 − i)!)p F(m)(t)p dt x j F(m)(t)p dt  xk(cid:90) x1 1 ((m − 1 − i)!)p x j+1(cid:90) x j F(m)(t)p dt. holds. Hence, proving that Consequently, k−1(cid:88) m−1(cid:88) j=1 i=0 which implies that P(i) x j (x j) − P(i) (x j+1 − x j)(m−i)p−1 ≤ (x j+1 − x j)1−p ((m − 1 − i)!)p x j+1(x j)p P(i) x j+1(x j)p x j (x j) − P(i) (x j+1 − x j)(m−i)p−1 Therefore, for every j ∈ {1, ..., k − 1} the following equality k−1(cid:88) m−1(cid:88) j=1 i=0 P(i) x j (x j) − P(i) (x j+1 − x j)(m−i)p−1 ≤ ep (cid:107)F(cid:107)p x j+1(x j)p p (R) . Lm Taking the supremum in the left hand side of this inequality over all finite strictly increasing p (R) j=1 ⊂ E, and then the infimum in the right hand side over all function F ∈ Lm sequences {x j}k satisfying (4.1), we obtain the required inequality Nm,p,E(P) ≤ e(cid:107)P(cid:107)m,p,E . The proof of the necessity is complete. (Sufficiency.) Let P = {Px : x ∈ E} be a Whitney (m − 1)-field defined on E such that λ = Nm,p,E(P) < ∞. See (4.2). Thus, for every strictly increasing sequence {x j}k x j+1(x j)p m−1(cid:88) k−1(cid:88) P(i) x j (x j) − P(i) (x j+1 − x j)(m−i)p−1 ≤ λp holds. j=1 i=0 j=1 ⊂ E the following inequality (4.4) 30 Our aim is to prove the existence of a function F ∈ Lm p (R) such that T m−1 x [F] = Px for every x ∈ E and (cid:107)F(cid:107)Lm p (R) ≤ C(m) λ. We construct F with the help of the classical Whitney extension method [66]. It is proven in [63] that this method provides an almost optimal extension of the restrictions of Whitney (m − 1)-fields generated by Sobolev Wm In this paper we will use a special one dimensional version of this method suggested by Whitney in [67, Section 4]. Since E is a closed subset of R, the complement of E, the set R \ E, can be represented as a p (Rn)-functions. union of a certain finite or countable family JE = {Jk = (ak, bk) : k ∈ K} (4.5) of pairwise disjoint open intervals (bounded or unbounded). Thus, ak, bk ∈ E ∪{±∞} for all k ∈ K, (4.6) R \ E = ∪{Jk = (ak, bk) : k ∈ K} and Jk(cid:48)∩ Jk(cid:48)(cid:48) = ∅ for every k(cid:48) , k(cid:48)(cid:48) ∈ K, k(cid:48) (cid:44) k(cid:48)(cid:48) To each interval J ∈ JE we assign a polynomial HJ ∈ P2m−1 as follows: (•1) Let J = (a, b) be an unbounded open interval, i.e., either a = −∞ and b is finite, or a is finite and b = +∞. In the first case (i.e., J = (a, b) = (−∞, b)) we set HJ = Pb, while in the second case (i.e., J = (a, b) = (a, +∞)) we set HJ = Pa. (•2) Let J = (a, b) ∈ JE be a bounded interval so that a, b ∈ E. In this case we define the polynomial HJ ∈ P2m−1 as the Hermite polynomial satisfying the following conditions: . H(i) J (a) = P(i) a (a) and H(i) J (b) = P(i) b (b) for all i = 0, ..., m − 1. (4.7) For the proof of the existence and uniqueness of the polynomial HJ we refer the reader to [2, Ch. 2, Section 11]. Let us note an explicit formula for HJ proven in [1, Lemma 1, p. 316]. To its formulation given a ∈ E and non-negative integer k ≤ m − 1 we let Pa,k denote a polynomial of degree at most k defined by Pa,k(x) = P(i) a (a) i! (x − a)i, x ∈ R. Clearly, Pa,m−1 = Pa. Proposition 4.2 For every bounded interval J = (a, b) ∈ JE the following equality HJ(x) = + holds. Finally, we define the extension F by the formula: F(x) = HJ(x) χJ(x), x ∈ R \ E. x ∈ E, (cid:32)m + k − 1 (cid:32)m + k − 1 m − 1 (cid:33)(cid:18) x − a (cid:33)(cid:32)b − x b − a (cid:19)k (cid:33)k m − 1 b − a k(cid:88) i=0 k=0 (cid:33)m m−1(cid:88) m−1(cid:88) (cid:19)m b − a b − a (cid:32)b − x (cid:18) x − a  Px(x), (cid:88) k=0 J∈JE 31 Pa,m−k−1(x) Pb,m−k−1(x) (4.8) Let us note that inequality (4.4) implies the following: for every x, y ∈ E and every i, 0 ≤ i ≤ m − 1, P(i) x (x) − P(i) y (x) ≤ λx − ym−i−1/p. Hence, x (x) − P(i) P(i) y (x) = o(x − ym−1−i) provided x, y ∈ E and 0 ≤ i ≤ m − 1. (4.9) Whitney [67] proved that for every (m − 1)-field P = {Px : x ∈ E} satisfying (4.9), the extension x (x) (Recall that p > 1.) F defined by formula (4.8) is a Cm−1-function on R which agrees with P on E, i.e., F(i)(x) = P(i) for every x ∈ E and every i, 0 ≤ i ≤ m − 1. Let us show that p (R) ≤ C(m) λ . Our proof of these facts relies on the following description of L1 p (R) and (cid:107)F(cid:107)Lm F ∈ Lm p(R)-functions. (4.10) Theorem 4.3 Let p > 1 and let τ > 0. Let G be a continuous function on R satisfying the following condition: There exists a constant A > 0 such that for every finite family I = {I = [uI, vI]} of pairwise disjoint closed intervals of diameter at most τ the following inequality holds. Then G ∈ L1 p(R) and I=[uI ,vI]∈I (cid:88) G(uI) − G(vI)p (vI − uI)p−1 ≤ A (cid:107)G(cid:107)L1 p(R) ≤ C A 1 p where C is an absolute constant. Proof. The Riesz theorem [51] tells us that G ∈ L1 For τ = ∞ inequality (4.11) follows from [4, Theorem 2] and [5, Theorem 4]. description of Sobolev spaces obtained in [5, § 4, 3◦].) For the case 0 < τ < ∞ we refer the reader to [63, Section 7, Theorem 7.3]. We will also need the following auxiliary lemmas. p(R). See also [40]. (cid:3) Lemma 4.4 Let J = (a, b) ∈ JE be a bounded interval. Then for every n ∈ {0, ..., m} and every x ∈ [a, b] the following inequality (4.11) (See also a holds. Here and Y1(x) = P(n) a (x) + Y2(x) = P(n) b (x) + J (x) ≤ C(m) min{Y1(x), Y2(x)} H(n)  m−1(cid:88)  m−1(cid:88) i=0 i=0  · (x − a)m−n  · (b − x)m−n . P(i) b (b) − P(i) (b − a)m−i a (b) P(i) b (a) − P(i) (b − a)m−i a (a) (Recall that HJ ∈ P2m−1 is the Hermite polynomial defined by equalities (4.7).) 32 Proof. Definition (4.7) implies the existence of γm, γm+1, ..., γ2m−1 ∈ R such that 2m−1(cid:88) k=m HJ(x) = Pa(x) + Hence, for every n ∈ {0, ..., m} and every x ∈ [a, b], for every x ∈ [a, b] . 1 k! γk (x − a)k 2m−1(cid:88) (k − n)! γk (x − a)k−n . 2m−1(cid:88) k=m 1 1 (k − n)! γk (b − a)k−n k=m H(n) J (x) = P(n) a (x) + (4.12) In particular, H(n) J (b) = P(n) a (b) + which together with (4.7) implies that 2m−1(cid:88) k=m (b − a)k−n (k − n)! γk = P(n) b (b) − P(n) a (b), for all n = 0, ..., m − 1 . Thus the tuple (γm, γm+1, ..., γ2m−1) is a solution of the above system of m linear equations with respect to m unknowns. It is proven in [67] that this solution can be represented in the following form: k = m, ..., 2m − 1, , (4.13) m−1(cid:88) i=0 γk = Kk,i b (b) − P(i) P(i) (b − a)k−i a (b) where Kk,i are certain constants depending only on m. This representation enables us to estimate H(n) P(i) b (b) − P(i) (b − a)k−i γk ≤ C(m) a (b) i=0 On the other hand, (4.12) tells us that m−1(cid:88) 2m−1(cid:88) J as follows: Thanks to (4.13), k = m, ..., 2m − 1 . , H(n) J (x) ≤ P(n) a (x) + γk (x − a)k−n for every x ∈ [a, b]. Hence, k=m J (x) ≤ P(n) H(n) a (x) + C(m) = P(n) a (x) + C(m) (x − a)k−n (x − a)k−n a (b) a (b) b (b) − P(i) P(i) (b − a)k−i b (b) − P(i) P(i) (b − a)k−i a (b) (x − a)m−n (b − a)m−i · a (b) (x − a)m−n (b − a)m−i 2m−1(cid:88) k=m (cid:19)k−m (cid:18) x − a b − a = P(n) a (x) + C(m) b (b) − P(i) P(i) ≤ P(n) a (x) + C(m) m b (b) − P(i) P(i) 1 (k − n)! m−1(cid:88) 2m−1(cid:88) i=0 k=m 2m−1(cid:88) m−1(cid:88) m−1(cid:88) m−1(cid:88) i=0 i=0 k=m i=0 33 proving that H(n) J (x) ≤ C(m) m Y1(x) for all x ∈ [a, b]. By interchanging the roles of a and b we obtain that H(n) J (x) ≤ C(m) Y2(x) on [a, b] proving the lemma. Lemma 4.5 Let J = (a, b) ∈ JE be a bounded interval. Then for every x ∈ [a, b] the following inequality (cid:3)  m−1(cid:88) i=0 b (b) − P(i) P(i) (b − a)m−i a (b) , b (a) − P(i) P(i) (b − a)m−i a (a) m−1(cid:88) i=0  H(m) J (x) ≤ C(m) min holds. (uI, vI) ⊂ R \ E for every I ∈ I. Then(cid:88) Proof. The proof is immediate from Lemma 4.4 because Pa and Pb belong to Pm−1. (cid:3) Lemma 4.6 Let I be a finite family of pairwise disjoint closed intervals I = [uI, vI] such that F(m−1)(vI) − F(m−1)(uI)p (vI − uI)p−1 I=[uI ,vI]∈I ≤ C(m)p λp . (4.14) Here F is the function defined by (4.8). Proof. Let I = [uI, vI] ∈ I. Since (uI, vI) ⊂ R \ E, there exist an interval J = (a, b) ∈ JE containing (uI, vI). (Recall that the family JE is defined by (4.5)). The extension formula (4.8) tells us that FJ = HJ. Therefore, by Lemma 4.5, F(m−1)(uI) − F(m−1)(vI) = H(m−1) J (uI) − H(m−1) ≤ C(m) (vI − uI) J m−1(cid:88) i=0 I (vI) ≤ (max P(i) a (a) − P(i) (b − a)m−i H(m)) · (vI − uI) b (a) . m−1(cid:88) i=0 a (a) − P(i) P(i) (b − a)(m−i)p b (a)p ≤ C(m)p m (vI − uI)1−p(vI − uI)p Hence, so that F(m−1)(uI) − F(m−1)(vI)p (vI − uI)p−1 m−1(cid:88) i=0 F(m−1)(uI) − F(m−1)(vI)p (vI − uI)p−1 ≤ C(m)p m (vI − uI) P(i) a (a) − P(i) (b − a)(m−i)p b (a)p . (4.15) For every J = (a, b) ∈ JE by IJ we denote a subfamily of I defined by Let (cid:101)J = {J ∈ J : IJ (cid:44) ∅}. Then, thanks to (4.15), for every J = (aJ, bJ) ∈ (cid:101)J IJ = {I ∈ I : I ⊂ [a, b]} . (cid:88) QJ = I=[uI ,vI]∈IJ F(m−1)(vI) − F(m−1)(uI)p (vI − uI)p−1 ≤ C p 34 (cid:88) I∈IJ   m−1(cid:88) i=0 diam I aJ(aJ) − P(i) P(i) (bJ − aJ)(m−i)p bJ (aJ)p  where C = C(m) is a constant depending only on m. Since the intervals of the family IJ are pairwise disjoint (because the intervals of the family I are pairwise disjoint), QJ ≤ C p (bJ − aJ) P(i) aJ(aJ) − P(i) (bJ − aJ)(m−i)p bJ P(i) aJ(aJ) − P(i) (aJ)p (bJ − aJ)(m−i)p−1 bJ .  m−1(cid:88) i=0 i=0 (aJ)p m−1(cid:88) (cid:88)  = C p (cid:88) J=(aJ ,bJ)∈(cid:101)J I=[uI ,vI]∈IJ P(i) aJ(aJ) − P(i) (aJ)p (bJ − aJ)(m−i)p−1 = bJ . Finally, Q = = (cid:88) (cid:88) J=(aJ ,bJ)∈(cid:101)J I=[uI ,vI]∈I F(m−1)(vI) − F(m−1)(uI)p m−1(cid:88) (vI − uI)p−1 (cid:88) J=(aJ ,bJ)∈(cid:101)J QJ ≤ C p F(m−1)(vI) − F(m−1)(uI)p (vI − uI)p−1 i=0 Since the intervals of the family (cid:101)J are pairwise disjoint, assumption (4.4) implies that Q ≤ C p λp proving the lemma. Lemma 4.7 Let I = {I = [uI, vI]} be a finite family of closed intervals such that uI, vI ∈ E for each I ∈ I. Suppose that the open intervals {(uI, vI) : I ∈ I} are pairwise disjoint. Then inequality (4.14) holds. (cid:3) Proof. Since F agrees with the Whitney (m − 1)-field P = {Px : x ∈ E}, we have F(m−1)(x) = P(m−1) x (x) for every x ∈ E. Hence, F(m−1)(uI) − F(m−1)(vI)p (vI − uI)p−1 = P(m−1) uI (uI) − P(m−1) (vI − uI)p−1 vI (vI)p . (cid:88) A = I=[uI ,vI]∈I (cid:88) I=[uI ,vI]∈I Since the intervals {(uI, vI) : I ∈ I} are pairwise disjoint, assumption (4.4) implies that A ≤ λp proving the lemma. We are in a position to finish the proof of the sufficiency. Let I be a finite family of pairwise disjoint closed intervals. We introduce the following notation: given an interval I = [u, v], u (cid:44) v, we put (cid:3) Y(I; F) = F(m−1)(uI) − F(m−1)(vI)p (vI − uI)p−1 . We put Y(I; F) = 0 whenever u = v, i.e., I = [u, v] is a singleton. Let I = [uI, vI] ∈ I be an interval such that I ∩ E (cid:44) ∅ and {uI, vI} (cid:49) E . Thus either uI or vI belongs to R \ E. Let u(cid:48) I and v(cid:48) I be the points of E nearest to uI and vI on I ∩ E respectively. Then [u(cid:48) I, v(cid:48) I] ⊂ [uI, vI]. Let Note that u(cid:48) I, v(cid:48) I ∈ E and (uI, u(cid:48) I, v(cid:48) I(2) = [u(cid:48) I] and I(3) = [v(cid:48) I(1) = [uI, u(cid:48) I], I, vI) ⊂ R \ E provided uI (cid:60) E and vI (cid:60) E. Furthermore, I), (v(cid:48) Y(I(1); F) + Y(I(2); F) + Y(I(3); F) Y(I; F) ≤ 3p (cid:110) I, vI] . (cid:111) . 35 If I ∈ I and (uI, vI) ⊂ R \ E, or uI, vI ∈ E, we put I(1) = I(2) = I(3) = I. Clearly, in all cases (cid:88) Y(I; F) ≤ 3p (cid:88) (cid:110) A(F;I) = (cid:111) I∈I A(F;I) ≤ 3p (cid:88) I∈(cid:101)I Y(I(1); F) + Y(I(2); F) + Y(I(3); F) I∈I Y(I; F) where (cid:101)I = (cid:110) I(1), I(2), I(3) : I ∈ I(cid:111) . proving that We know that for each I = [uI, vI] ∈(cid:101)I either (uI, vI) ∈ R \ E, or uI, vI ∈ E, or uI = vI (and so Y(I; F) = 0). Furthermore, the sets {(uI, vI) : I ∈(cid:101)I} are pairwise disjoint. We apply Lemmas 4.6 and 4.7 to the family(cid:101)I and obtain that A(F;I) ≤ C(m)p λp. Since I is an arbitrary finite family of pairwise disjoint closed intervals, the function G = F(m−1) satisfies the p(R) ≤ C(m) λ. hypothesis of Theorem 4.3. This theorem tells us that F(m−1) ∈ L1 Hence we conclude that (4.10) holds proving the sufficiency part of Theorem 4.1. p(R) and (cid:107)F(m−1)(cid:107)L1 The proof of Theorem 4.1 is complete. (cid:3) 4.2. The Main Lemma: from jets to Lagrange polynomials. Until the end of Section 4.3, we assume that f is a function defined on a closed set E ⊂ R with #E ≥ m + 1, which satisfies the hypothesis of Theorem 1.7. This hypothesis tells us that λ = Lm,p( f : E) < ∞. (4.16) See (1.14). This enables us to make the following assumption. Assumption 4.8 For every finite strictly increasing sequence of points {x0, ..., xn} ⊂ E, n ≥ m, the following inequality n−m(cid:88) (xi+m − xi)∆m f [xi, ..., xi+m]p ≤ λp (4.17) i=0 holds. Our aim is to prove that there exists F ∈ Lm p (R) such that FE = f and (cid:107)F(cid:107)Lm p (R) ≤ C(m) λ . Clearly, by (4.17), sup S⊂E, #S =m+1 ∆m f [S ] (diam S ) 1 p ≤ λ, so that inequality (3.32) holds. In Section 3 we have proved that for any function f satisfying this inequality the Whitney (m − 1)-field : E → R P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} satisfying conditions (3.39)-(3.42) of Definition 3.14 (4.18) is well defined. We prove the existence of the function F with the help of Theorem 4.1 which we apply to the field P(m,E)[ f ]. To enable us to do this, we first have to check that the hypothesis of this theorem 36 holds, i.e., we must show that for every integer k > 1 and every strictly increasing sequence {x j}k in E the following inequality j=1 P(i) x j+1(x j) − P(i) (x j+1 − x j)(m−i)p−1 ≤ C(m)p λp x j (x j)p (4.19) k−1(cid:88) m−1(cid:88) j=1 i=0 holds. We prove this inequality in Lemma 4.16 below. One of the main ingredients of this proof is the following Main Lemma 4.9 Let k ∈ N, ε > 0, and let X = {x1, ..., xk} ⊂ E, x1 < ... < xk, be a sequence of points in E. There exist a positive integer (cid:96) ≥ m, a finite strictly increasing sequence V = {v1, ..., v(cid:96)} of points in E, and a mapping H : X → 2V such that: (1) For every x ∈ X the set H(x) consists of m consecutive points of the sequence V. Thus, H(x) = {v j1(x), ..., v j2(x)} where 1 ≤ j1(x) ≤ j2(x) = j1(x) + m − 1 ≤ (cid:96); (2) x ∈ H(x) for each x ∈ X. In particular, X ⊂ V. Furthermore, given i ∈ {1, ..., k − 1} let xi = vκi and xi+1 = vκi+1. Then 0 < κi+1 − κi ≤ 2m. (3) Let x(cid:48), x(cid:48)(cid:48) ∈ X, x(cid:48) < x(cid:48)(cid:48). Then min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)) and max H(x(cid:48)) ≤ max H(x(cid:48)(cid:48)) ; (4) For every x(cid:48), x(cid:48)(cid:48) ∈ X such that H(x(cid:48)) (cid:44) H(x(cid:48)(cid:48)) the following inequality diam H(x(cid:48)) + diam H(x(cid:48)(cid:48)) ≤ 2(m + 1)x(cid:48) − x(cid:48)(cid:48) holds; (5) For every x, y ∈ X and every i, 0 ≤ i ≤ m − 1, we have H(x)[ f ](y) < ε. P(i) x (y) − L(i) (4.20) (4.21) Proof. We proceed by steps. STEP 1. At this step we introduce the sequence V and the mapping H. We recall that, given x ∈ E, by Sx and sx we denote a subset of E and a point in E whose properties are described in Propositions 3.11 and 3.12. In particular, #Sx ≤ m. Let SX = Sx and let n = #SX. (4.22) Clearly, one can consider SX as a finite strictly increasing sequence of points {ui}n i=1 in E. Thus (4.23) If #Sx = m for every x ∈ X, we set V = SX and H(x) = Sx, x ∈ X. In this case the required SX = {u1, ..., un} and u1 < u2 < ... < un . properties (1)-(5) of the Main Lemma are immediate from Propositions 3.11 and 3.12. 37 (cid:91) x∈X However, in general, the set X may have points x with #Sx < m. For those x we construct the required set H(x) by adding to Sx a certain finite set (cid:101)H(x) ⊂ E. In other words, we define H(x) as where (cid:101)H(x) is a subset of E such that x ∈ X, (4.24) H(x) = (cid:101)H(x) ∪ Sx, (cid:101)H(x) ∩ Sx = ∅ and #(cid:101)H(x) = m − #Sx. Finally, we set V = ∪{H(x) : x ∈ X} . We construct (cid:101)H(x) by picking (m − #Sx) points of E in a certain small neighborhood of sx. We turn to the precise definition of the mapping (cid:101)H. First, we set (cid:101)H(x) = ∅ whenever #Sx = m. Thus, Propositions 3.11, 3.12, and Lemma 3.16 enable us to prove that this neighborhood can be chosen so small that V and H will satisfy conditions (1)-(5) of the Main Lemma. (4.25) H(x) = Sx provided #Sx = m . (4.26) Note, that in this case Px = LSx[ f ] = LH(x)[ f ] (see (3.42)), so that (4.21) trivially holds. Let us define the sets H(x) for all points x ∈ X such that #Sx < m. We recall that part (iii) of Proposition 3.12 tells us that for each x ∈ X with #Sx < m the point sx is a limit point of E. (4.27) In turn, part (i) of this proposition tells us that either sx = min Sx or sx = max Sx . ZX = {sx : x ∈ X, #Sx < m}. every point z ∈ ZX is a limit point of E. Let Then, thanks to (4.27), Given z ∈ ZX, let K(z) = {x ∈ X : sx = z} . The following lemma describes main properties of the sets K(z), z ∈ ZX. Lemma 4.10 Let z ∈ ZX. Suppose that K(z) (cid:44) {z}. Then: (1) The set K(z) lies on one side of z, i.e., either max K(z) ≤ z or min K(z) ≥ z ; (2) If min K(z) ≥ z , then for every r > 0 the interval (z − r, z) contains an infinite number of points of E . 38 (4.28) (4.29) (4.30) (4.31) If max K(z) ≤ z then each interval (z, z + r) contains an infinite number of points of E; (3) If min K(z) ≥ z then [z, y] ∩ X ⊂ K(z) for every y ∈ K(z) . Furthermore, and min Sy = z for all y ∈ K(z), (4.32) (4.33) Sx ⊂ Sy for every y ∈ K(z) and every x ∈ [z, y] ∩ E. and every x ∈ [y, z] ∩ E. In addition, max Sy = z for all y ∈ K(z); (4.34) If max K(z) ≤ z then [y, z] ∩ X ⊂ K(z) for each y ∈ K(z). Moreover, Sx ⊂ Sy for every y ∈ K(z) (4) Assume that min K(z) ≥ z. Let ¯z = max K(z). Then K(z) = [z, ¯z] ∩ X. (4.35) Furthermore, in this case K(z) ⊂ S ¯z. If max K(z) ≤ z then K(z) = [z, z] ∩ X where z = min K(z). In this case K(z) ⊂ S z; (5) #K(z) ≤ m. Proof. (1) Suppose that (4.30) does not hold so that there exist z(cid:48), z(cid:48)(cid:48) ∈ K(z) such that z(cid:48)(cid:48) < z < z(cid:48). Thanks to (4.29), z = sz(cid:48)(cid:48) = sz(cid:48). We also know that z(cid:48), sz(cid:48) ∈ Sz(cid:48), see part (i) of Proposition 3.11 and part (i) of Proposition 3.12. This property and (3.35) tell us that [z, z(cid:48)] ∩ E = [sz(cid:48), z(cid:48)] ∩ E ⊂ [min Sz(cid:48), max Sz(cid:48)] ∩ E = Sz(cid:48). Part (i) of Proposition 3.11 also tells us that #Sz(cid:48) ≤ m proving that the interval (z, z(cid:48)) contains at most m points of E. (4.36) (4.37) In the same way we show that the interval (z(cid:48)(cid:48), z) contains at most m points of E. Thus, the interval (z(cid:48)(cid:48), z(cid:48)) contains a finite number of points of E proving that z is an isolated point of E. On the other hand, z ∈ ZX so that, thanks to (4.28), z is a limit point of E, a contradiction. (2) Let z(cid:48) ∈ K(z), z(cid:48) (cid:44) z. Then z < z(cid:48) so that, thanks to (4.37), the interval (z, z(cid:48)) contains at most m points of E. But z is a limit point of E, see (4.28), so that the interval (z−r, z) contains an infinite number of points of E. This proves (4.31). In the same fashion we prove the second statement of part (2). (3) Let min K(z) ≥ z, and let y ∈ K(z), y (cid:44) z. Prove that sx = z for every x ∈ [z, y] ∩ E. We know that z = sy < y. Furthermore, property (4.36) tells us that (4.38) We recall that, thanks to (4.28), z is a limit point of E so that Sz = {z}. Part (iii) of Proposition [z, y] ∩ E ⊂ Sy. 3.11 tells us that so that z = min Sz ≤ min Sx and max Sx ≤ max Sy, Sx ⊂ [z, max Sy] ∩ E. 39 (4.39) In particular, z ≤ min Sy. But z = sy ∈ Sy, so that z = min Sy proving (4.33). In turn, thanks to (3.35), [min Sy, max Sy] ∩ E = [z, max Sy] ∩ E = Sy. This and (4.39) imply that Sx ⊂ Sy for every x ∈ [z, y] ∩ E proving (4.34). Moreover, thanks to (4.27), if #Sx < m then sx is a limit point of E. But sx ∈ Sx ⊂ Sy, therefore, If #Sx = m then Sx = Sy (because Sx ⊂ Sy and #Sy ≤ m). Hence, z = sy ∈ Sx. But z is a limit part (ii) of Proposition 3.12 implies that sx = sy = z. point of E which together with part (ii) of Proposition 3.12 imply that sx = z = sy. Thus, in all cases sx = z proving property (3) of the lemma in the case under consideration. Using the same ideas we prove the second statement of the lemma related to the case max K(z) ≤ z. (4) Thanks to (4.32) [z, ¯z] ∩ X ⊂ K(z) . On the other hand, K(z) ⊂ [z, ¯z] because z ≤ min K(z) and ¯z = max K(z). Since K(z) ⊂ X, see (4.29), K(z) ⊂ [z, ¯z] ∩ X proving (4.35). Furthermore, thanks to (4.38) (with y = ¯z), [z, ¯z] ∩ E ⊂ S¯z so that K(z) = [z, ¯z] ∩ X ⊂ [z, ¯z] ∩ E ⊂ S¯z . In the same way we prove the last statement of part (4) related to the case z ≥ max K(z). (5) We recall that #Sx ≤ m for every x ∈ E, see part (i) of Proposition 3.11. Part (4) of the present lemma tells us that K(z) ⊂ Sy where y = max K(z) or y = min K(z). Hence #K(z) ≤ #Sy ≤ m. The proof of Lemma 4.10 is complete. Let us fix a point z ∈ ZX and define the set H(x) for every x ∈ K(z). Thanks to property (4.30), (cid:3) it suffices to consider the following three cases: Case ((cid:70)1). Suppose that K(z) (cid:44) {z} and min K(z) ≥ z. [z, y] ∩ X ⊂ K(z). 4.10. Furthermore, part (5) of this lemma tells us that #K(z) ≤ m. (4.40) Let y ∈ K(z). Thus y ∈ X and sy = z; we also know that y ≥ z. Then property (4.32) tells us that We also note that K(z) = [z, ¯z] ∩ X where ¯z = max K(z), and K(z) ⊂ S ¯z, see part (4) of Lemma Let us fix several positive constants which we need for definition of the sets {H(x) : x ∈ X}. We recall that X = {x1, ...xk} and x1 < ... < xk. Let IX = [x1, xk]. We also recall that inequality (3.32) holds, and sx = z provided x ∈ K(z). This enables us to apply Lemma 3.16 to the interval I = IX and the point x ∈ K(z). This lemma tells us that (cid:107)LS (cid:48)[ f ] − Px(cid:107)Cm(IX) = 0 . lim S (cid:48)\Sx→z Sx⊂ S (cid:48)⊂E, #S (cid:48)=m Thus, there exists a constant δx = δx(ε) > 0 satisfying the following condition: for every m-point set S (cid:48) such that Sx ⊂ S (cid:48) ⊂ E and S (cid:48) \ Sx ⊂ (z − δx, z + δx) we have S (cid:48)[ f ](y) < ε for every i, 0 ≤ i ≤ m − 1, and every y ∈ IX . P(i) x (y) − L(i) 40 We recall that SX = {u1, ..., un} is the set defined by (4.22) and (4.23). Let τX = 1 4 min i=1,...,n−1 (ui+1 − ui). Thus, Finally, we set x − y ≥ 4τX provided x, y ∈ SX, x (cid:44) y . δz = min{τX, min x∈K(z) δx} . (4.41) (4.42) (4.43) Clearly, δz > 0 (because K(z) is finite). Definition (4.43) implies the following: Let x ∈ K(z). Then for every i, 0 ≤ i ≤ m− 1, and every m-point set S (cid:48) such that Sx ⊂ S (cid:48) ⊂ E and S (cid:48) \ Sx ⊂ (z − δz, z + δz) the following inequality P(i) x (y) − L(i) S (cid:48)[ f ](y) < ε, y ∈ X, (4.44) holds. Inequality by (4.31) tells us that the interval (z − δz, z) contains an infinite number of points of E. Let us pick m − 1 distinct points a1 < a2 < ... < am−1 in (z − δz, z) ∩ E and set W(z) = {a1, a2, ..., am−1}. Thus, In particular, W(z) = {a1, a2, ..., am−1} ⊂ (z − δz, z) ∩ E. (4.45) z − τX < a1 < a2 < ... < am−1 < z (4.46) (cid:101)H(x) = ∅ and H(x) = Sx provided (cid:96)x = m. Let x ∈ K(z), and let (cid:96)x = #Sx. We introduce the set (cid:101)H(x) as follows: we set (because δz ≤ τX). If (cid:96)x < m, we define (cid:101)H(x) by letting(cid:101)H(x) = {a(cid:96)x, a(cid:96)x+1, ..., am−1}. Clearly, #(cid:101)H(x) + #Sx = m. Then we define H(x) by formula (4.24), i.e., we set H(x) = (cid:101)H(x) ∪ Sx. This definition, property (4.45) and inequality (4.44) imply that for every y ∈ X and every i, 0 ≤ i ≤ m − 1, the following inequality (4.47) (4.48) P(i) x (y) − L(i) H(x)[ f ](y) < ε 41 (4.49) holds. Furthermore, property (4.34) tells us that Sx(cid:48) ⊂ Sx(cid:48)(cid:48) provided x(cid:48), x(cid:48)(cid:48) ∈ K(z), x(cid:48) < x(cid:48)(cid:48). This property and definition (4.48) imply that min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)) , x(cid:48)(cid:48) ∈ K(z), x(cid:48) for every x(cid:48) < x(cid:48)(cid:48) (4.50) . Let us also note the following property of the set H(x) which directly follows from its definition: Let Then and (cid:98)H(x) = (cid:101)H(x) ∪ {sx} . [min H(x), max H(x)] = [min Sx, max Sx] ∪ [min(cid:98)H(x), max(cid:98)H(x)] [min Sx, max Sx] ∩ [min(cid:98)H(x), max(cid:98)H(x)] = {sx}. (4.51) (4.52) (4.53) Case ((cid:70)2). Suppose that K(z) (cid:44) {z} and max K(z) ≤ z. (4.54) Using the same approach as in Case ((cid:70)1), see (4.40), given x ∈ K(z), we define a corresponding constant δz, a set W(z) = {a1, ..., am−1} and sets (cid:101)H(x) and H(x). More specifically, we pick a strictly increasing sequence W(z) = {a1, a2, ..., am−1} ⊂ (z, z + δz) ∩ E. (4.55) In particular, this sequence has the following property: z < a1 < a2 < ... < am−1 < z + τX. (Recall that τX is defined by (4.41).) Then we set (cid:101)H(x) = ∅ and H(x) = Sx if (cid:96)x = m, and As in Case ((cid:70)1), see (4.40), our choice of δz, (cid:101)H(x) and H(x) provides inequality (4.49) and (Recall that (cid:96)x = #Sx.) Finally, we define the set H(x) by formula (4.24). (cid:101)H(x) = {a1, a2, ..., am−(cid:96)x} if (cid:96)x < m. (4.56) (4.57) properties (4.50), (4.52) and (4.53). Case ((cid:70)3). Suppose that (4.58) Note that in this case z ∈ X is a limit point of E and Sz = z. This enables us to pick an m − 1 K(z) = {z}. point set W(z) = {a1, ..., am−1} ⊂ E such that either (4.45) or (4.55) hold. (We may assume that the sequence {ai}m−1 sing so that inequalities (4.46) and (4.56) hold as well.) i=1 is strictly increa- 42 We set (cid:101)H(z) = W(z). Thus, in this case the set H(z) is defined by formula (4.24) with (cid:101)H(z) = {a1, a2, ..., am−1}, (4.59) i.e., H(z) = {z, a1, a2, ..., am−1}. choice of the set W(z) provides inequality (4.49) with x = z and H(x) = {z, a1, a2, ..., am−1}. It is also clear that properties (4.52), (4.53) hold in the case under consideration. Moreover, our We have defined the set H(x) for every x ∈ X. Then we define the set V by formula (4.25). Clearly, V is a finite subset of E. Let us enumerate the points of this set in increasing order: thus, we represent V in the form V = {v1, v2, ..., v(cid:96)} j=1 is a strictly increasing sequence of points in E. STEP 2. At this step we prove two auxiliary lemmas which describe a series of important where (cid:96) is a positive integer and {v j}(cid:96) properties of the mappings (cid:101)H and H. Lemma 4.11 (i) For each x ∈ X the following inclusion (cid:98)H(x) ⊂ (sx − τX, sx + τX) holds. (Recall that (cid:98)H(x) = (cid:101)H(x) ∪ {sx}, see (4.51).) (ii) The following property (4.60) (4.61) [min H(x), max H(x)] ⊂ [min S(x) − τX, max S(x) + τX] holds for every x ∈ X. Proof. Property (i) is immediate from (4.46), (4.51), (4.56). In turn, property (ii) is immediate from (3.37), (4.24) and (4.60). Lemma 4.12 Let x, y ∈ X. Suppose that #Sx < m and (cid:3) [min(cid:98)H(x), max(cid:98)H(x)] ∩ [min H(y), max H(y)] (cid:44) ∅ . Then sx = sy. Proof. Part (iii) of Proposition 3.12 tells us that the point sx is a limit point of E. If y = sx, part (ii) of Proposition 3.12 implies that sy = y so that in this case the lemma holds. Let us prove the lemma for y (cid:44) sx. To do so, we assume that Prove that sx (cid:44) sy. sx (cid:60) [min Sy, max Sy]. Indeed, otherwise, part (i) of Proposition 3.11 tells us that sx ∈ [min Sy, max Sy] ∩ E = Sy. 43 (4.62) (4.63) But sx is a limit point of E. In this case part (ii) of Proposition 3.12 tells us that sx = sy. This contradicts (4.62) proving (4.63). In particular, sx (cid:44) min Sy and sx (cid:44) max Sy. Furthermore, sx, min Sy, max Sy ∈ SX, see (4.22). Therefore, thanks to (4.42), sx − min Sy ≥ 4τX and sx − max Sy ≥ 4τX. Hence, On the other hand, part (i) and part (ii) of Lemma 4.11 tell us that (cid:98)H(x) ⊂ (sx − τX, sx + τX) and dist(sx, [min Sy, max Sy]) ≥ 4τX. Hence, [min H(y), max H(y)] ⊂ [min Sy − τX, max Sy + τX]. [min(cid:98)H(x), max(cid:98)H(x)] ∩ [min H(y), max H(y)] = ∅. This contradicts (4.61) proving that assumption (4.62) does not hold. The proof of the lemma is complete. (cid:3) STEP 3. We are in a position to prove properties (1)-(5) of the Main Lemma 4.9. (cid:4) Proof of property (1). This property is equivalent to the following statement: for every x ∈ X the following equality [min H(x), max H(x)] ∩ V = H(x) (4.64) holds. to a contradiction. Let us assume that (4.64) does not hold for certain x ∈ X, and show that this assumption leads Thanks to definition (4.25), if (4.64) does not hold then there exist y ∈ X and u ∈ H(y) such that (4.65) u ∈ [min H(x), max H(x)] \ H(x). Prove that #Sx < m. Indeed, otherwise, Sx = H(x) (see (4.26)). In this case (3.35) implies that so that [min H(x), max H(x)] ∩ E = H(x) [min H(x), max H(x)] \ H(x) = ∅. This contradicts (4.65) proving that #Sx < m. Furthermore, property (3.35) tells us that [min Sx, max Sx] ∩ E = Sx ⊂ H(x) so that Hence, thanks to (4.52), u (cid:60) [min Sx, max Sx]. (4.66) (4.67) We conclude that the hypothesis of Lemma 4.12 holds for x and y because #Sx < m, u ∈ H(y) and (4.67) holds. This lemma tells us that sx = sy. u ∈ [min(cid:98)H(x), max(cid:98)H(x)]. 44 Let z = sx so that x, y ∈ K(z), see (4.29). Clearly, K(z) (cid:44) {z}; otherwise x = y which contradicts We know that either min K(z) ≥ z (i.e., z satisfies the condition of the case ((cid:70)1) of STEP 1) or (4.65). max K(z) ≤ z (the case ((cid:70)2) of STEP 1 holds). Suppose that min K(z) ≥ z, see ((cid:70)1). Then min Sx = sx = z so that [min Sx, max Sx] = [z, max Sx]. Moreover, thanks to (4.48) and (4.51), (cid:101)H(x) = {a(cid:96)x, a(cid:96)x+1, ..., am−1} and (cid:98)H(x) = {a(cid:96)x, ..., am−1, z}. Here (cid:96)x = #Sx and a1, ..., am−1 are m − 1 distinct points of E satisfying inequality (4.46). Properties (4.66) and (4.68) tell us that u (cid:44) z = sx. This, (4.67) and (4.69) imply that u ∈ [a(cid:96)x, z). On the other hand, u ∈ H(y). Since y ∈ K(z), definitions (4.24) and (4.48) tell us that H(y) = {a(cid:96)y, ..., am−1} ∪ Sy. But min Sy = z, see (4.33). This and (4.70) imply that u ∈ {a(cid:96)y, ..., am−1} and u ∈ [a(cid:96)x, z). (4.68) (4.69) (4.70) Hence, u ∈ {a(cid:96)x, ..., am−1} ⊂ H(x) which contradicts (4.65). In the same way we obtain a contradiction whenever z satisfies the condition of the case ((cid:70)2). The proof of property (1) of the Main Lemma is complete. (cid:4) Proof of property (2). Part (i) of Proposition 3.12 tells us that x ∈ Sx for every x ∈ E. In turn, definition (4.24) implies that Sx ⊂ H(x) so that x ∈ H(x). Hence, x ∈ V for each x ∈ X, see (4.25), proving that X ⊂ V. Let us prove that 0 < κi+1 − κi ≤ 2m provided xi = vκi and xi+1 = vκi+1. The first inequality is obvious because xi < xi+1 and V = {v j}(cid:96) j=1 is a strictly increasing sequence. Our proof of the second inequality relies on the following fact: V ∩ [xi, xi+1] ⊂ H(xi) ∪ H(xi+1). (4.71) Indeed, let v ∈ V ∩ [xi, xi+1]. Then definition (4.25) implies the existence of a point ¯x ∈ X such that H( ¯x) (cid:51) v. Hence, v ≤ max H( ¯x). Suppose that ¯x < xi. In this case property (3) of the Main Lemma 4.9 (which we prove below) tells us that max H( ¯x) ≤ max H(xi) so that xi ≤ v ≤ max H(xi). Hence, v ∈ [min H(xi), max H(xi)] (because xi ∈ H(xi)). We also know that v ∈ V. This and property (4.64) (which is equivalent to property (1) of the Main Lemma) imply that v ∈ [min H(xi), max H(xi)] ∩ V = H(xi). In the same way we show that v ∈ H(xi+1) provided ¯x > xi+1, and the proof of (4.71) is complete. Since #H(xi) = #H(xi+1) = m, property (4.71) tells us that the interval [xi, xi+1] contains at most 2m points of the set V. This implies the required second inequality κi+1 − κi ≤ 2m completing the proof of part (2) of the Main Lemma. 45 (cid:4) Proof of property (3). Let x(cid:48), x(cid:48)(cid:48) ∈ X, x(cid:48) < x(cid:48)(cid:48). Prove that min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)). We recall that H(x(cid:48)) = (cid:101)H(x(cid:48)) ∪ Sx(cid:48) and H(x(cid:48)(cid:48)) = (cid:101)H(x(cid:48)(cid:48)) ∪ Sx(cid:48)(cid:48), see (4.24). Here (cid:101)H is the set defined by formulae (4.47), (4.48), (4.57) and (4.59). Part (iii) of Proposition 3.11 tells us that min Sx(cid:48) ≤ min Sx(cid:48)(cid:48). Since sx(cid:48)(cid:48) ∈ Sx(cid:48)(cid:48), we have min Sx(cid:48) ≤ min Sx(cid:48)(cid:48) ≤ sx(cid:48)(cid:48) . (4.72) (4.73) We proceed the proof of (4.72) by cases. Case 1. Assume that min Sx(cid:48) < sx(cid:48)(cid:48). Since the points min Sx(cid:48) and sx(cid:48)(cid:48) belong to the set SX (see (4.22)), inequality (4.42) tells us that (4.74) sx(cid:48)(cid:48) − Sx(cid:48) > 4τX . On the other hand, part (i) of Lemma 4.11 implies that (cid:101)H(x(cid:48)(cid:48)) ⊂ (cid:98)H(x(cid:48)(cid:48)) ⊂ (sx(cid:48)(cid:48) − τX, sx(cid:48)(cid:48) + τX). (Recall that (cid:98)H(x(cid:48)(cid:48)) = (cid:101)H(x(cid:48)(cid:48)) ∪ {sx(cid:48)(cid:48)}, see (4.51).) This inclusion and (4.74) imply that min(cid:101)H(x(cid:48)(cid:48)) > sx(cid:48)(cid:48) − τX > min Sx(cid:48) + 3τX > min Sx(cid:48) . This inequality and (4.73) tell us that min H(x(cid:48)(cid:48)) = min ((cid:101)H(x(cid:48)(cid:48)) ∪ Sx(cid:48)(cid:48)) = min{min(cid:101)H(x(cid:48)(cid:48)), min Sx(cid:48)(cid:48)} ≥ min Sx(cid:48) ≥ min ((cid:101)H(x(cid:48)) ∪ Sx(cid:48)) = min H(x(cid:48)) proving (4.72) in the case under consideration. Case 2. Suppose that min Sx(cid:48) = sx(cid:48)(cid:48), and consider two cases. Case 2.1: #Sx(cid:48)(cid:48) = m. Then H(x(cid:48)(cid:48)) = Sx(cid:48)(cid:48), see (4.26), which together with (4.73) imply that min H(x(cid:48)) ≤ min Sx(cid:48) ≤ min Sx(cid:48)(cid:48) = min H(x(cid:48)(cid:48)) proving (4.72). Case 2.2: #Sx(cid:48)(cid:48) < m. In this case part (iii) of Proposition 3.12 tells us that sx(cid:48)(cid:48) is a limit point of E. Thanks to the assumption of Case 2, sx(cid:48)(cid:48) = min Sx(cid:48) so that the point min Sx(cid:48) is a limit point of E as well. Hence, min Sx(cid:48) = sx(cid:48), see part (ii) of Proposition 3.12. Thus, sx(cid:48) = sx(cid:48)(cid:48) = min Sx(cid:48). Let z = sx(cid:48) = sx(cid:48)(cid:48). We know that < x(cid:48)(cid:48) z = min Sx(cid:48) ≤ x(cid:48) In particular, K(z) (cid:44) {z} proving that the point z satisfies In this case inequality (4.48) tells us that so that x(cid:48), x(cid:48)(cid:48) ∈ K(z), see (4.29). the condition of Case ((cid:70)1) of STEP 1, see (4.40). min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)) proving (4.72) in Case 2.2. Thus, (4.72) holds in all cases. 46 Since each of the sets H(x(cid:48)) and H(x(cid:48)(cid:48)) consists of m +1 consecutive points of the strictly increa- i=1 and min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)), we conclude that max H(x(cid:48)) ≤ max H(x(cid:48)(cid:48)). sing sequence V = {vi}(cid:96) The proof of property (3) is complete. (cid:4) Proof of property (4). Let x(cid:48), x(cid:48)(cid:48) ∈ X, x(cid:48) (cid:44) x(cid:48)(cid:48), and let H(x(cid:48)) (cid:44) H(x(cid:48)(cid:48)). (4.75) Part (i) of Lemma 3.11 and definition (4.22) tell us that x(cid:48) ∈ Sx(cid:48), x(cid:48)(cid:48) ∈ Sx(cid:48)(cid:48) and Sx(cid:48), Sx(cid:48)(cid:48) ∈ SX. Hence, x(cid:48), x(cid:48)(cid:48) ∈ SX so that, thanks to (4.42), x(cid:48) − x(cid:48)(cid:48) ≥ 4τX. In turn, definition (4.24) implies that diam H(x(cid:48)) ≤ diam(cid:101)H(x(cid:48)) + diam Sx(cid:48) and diam H(x(cid:48)(cid:48)) ≤ diam(cid:101)H(x(cid:48)(cid:48)) + diam Sx(cid:48)(cid:48) . Furthermore, part (i) of Lemma 4.11 tells us that max{diam(cid:101)H(x(cid:48)), diam(cid:101)H(x(cid:48)(cid:48))} ≤ 2τX < x(cid:48) − x(cid:48)(cid:48). Hence, diam H(x(cid:48)) ≤ diam Sx(cid:48) + x(cid:48) − x(cid:48)(cid:48) and diam H(x(cid:48)(cid:48)) ≤ diam Sx(cid:48)(cid:48) + x(cid:48) − x(cid:48)(cid:48) . (4.76) Prove that Sx(cid:48) (cid:44) Sx(cid:48)(cid:48). Indeed, suppose that Sx(cid:48) = Sx(cid:48)(cid:48) and prove that this equality contradicts If #Sx(cid:48) = #Sx(cid:48)(cid:48) = m then Sx(cid:48) = Hx(cid:48) and Sx(cid:48)(cid:48) = Hx(cid:48)(cid:48), see (4.26), which implies the required (4.75). contradiction H(x(cid:48)) = H(x(cid:48)(cid:48)). part (ii) of Proposition 3.12. A similar statement is true for sx(cid:48)(cid:48) and Sx(cid:48)(cid:48). Hence sx(cid:48) = sx(cid:48)(cid:48). Let now #Sx(cid:48) = #Sx(cid:48)(cid:48) < m. In this case sx(cid:48) is the unique limit point of E which belongs to Sx(cid:48), see Let z = sx(cid:48) = sx(cid:48)(cid:48). Thus, x(cid:48), x(cid:48)(cid:48) ∈ K(z), see (4.29). Since x(cid:48) (cid:44) x(cid:48)(cid:48), we have K(z) (cid:44) {z}, so that the (4.54)) holds. Furthermore, since #Sx(cid:48) = #Sx(cid:48)(cid:48) < m, the sets (cid:101)H(x(cid:48)),(cid:101)H(x(cid:48)(cid:48)) are determined by the point z satisfies either the condition of Case ((cid:70)1) (see (4.40)) or the condition of Case ((cid:70)2) (see formula (4.48) or (4.57) respectively. In both cases the definitions of the sets (cid:101)H(x(cid:48)),(cid:101)H(x(cid:48)(cid:48)) depend Thus, (cid:101)H(x(cid:48)) = (cid:101)H(x(cid:48)(cid:48)) proving that only on the point z (which is the same for x(cid:48) and x(cid:48)(cid:48) because z = sx(cid:48) = sx(cid:48)(cid:48)) and the number of points in the sets Sx(cid:48) and Sx(cid:48)(cid:48) (which of course is also the same because Sx(cid:48) = Sx(cid:48)(cid:48)). H(x(cid:48)) = (cid:101)H(x(cid:48)) ∪ Sx(cid:48) = (cid:101)H(x(cid:48)(cid:48)) ∪ Sx(cid:48)(cid:48) = H(x(cid:48)(cid:48)), a contradiction. This contradiction proves that Sx(cid:48) (cid:44) Sx(cid:48)(cid:48). In this case part (ii) of Proposition 3.11 tells us that diam Sx(cid:48) + diam Sx(cid:48)(cid:48) ≤ 2 mx(cid:48) − x(cid:48)(cid:48) . Combining this inequality with (4.76), we obtain the required inequality (4.20) proving the prop- erty (4) of the Main Lemma. (cid:4) Proof of property (5). Constructing the sets H(x), x ∈ X, we have noted that in all cases of STEP 1 (Case ((cid:70)1) (see (4.40)), Case ((cid:70)2) (see (4.54)), Case ((cid:70)3) (see (4.58))) inequality (4.49) holds for all y ∈ X and all i, 0 ≤ i ≤ m−1. This inequality coincides with inequality (4.21) proving property (5) of the Main Lemma. 47 (cid:3) The proof of Main Lemma 4.9 is complete. Remark 3.13 tells us that whenever E is a sequence of points in R, for each x ∈ E the set Sx con- sists of m consecutive terms of the sequence E. This enables us to set H(x) = Sx in formulation of the Main Lemma 4.9 which leads to the following version of this lemma for the case of sequences. Lemma 4.13 (Main Lemma for sequences) Let E = {xi}(cid:96)2 where (cid:96)1, (cid:96)2 ∈ Z∪{−∞, +∞}, (cid:96)1 + m ≤ (cid:96)2, be a strictly increasing sequence of points in R. i=(cid:96)1 There exists a mapping H : E → 2E having the following properties: (1) For every x ∈ E the set H(x) consists of m consecutive points of the sequence E; (2) x ∈ H(x) for each x ∈ E; (3) Let x(cid:48), x(cid:48)(cid:48) ∈ E, x(cid:48) < x(cid:48)(cid:48). Then min H(x(cid:48)) ≤ min H(x(cid:48)(cid:48)) and max H(x(cid:48)) ≤ max H(x(cid:48)(cid:48)); (4) For every x(cid:48), x(cid:48)(cid:48) ∈ E such that H(x(cid:48)) (cid:44) H(x(cid:48)(cid:48)) the following inequality diam H(x(cid:48)) + diam H(x(cid:48)(cid:48)) ≤ 2mx(cid:48) − x(cid:48)(cid:48) holds; (5) Px = LH(x) for every x ∈ E. Proof. The proof is immediate from Proposition 3.11 and definition (3.42). (cid:3) 4.3. Proof of the sufficiency part of the variational criterion. We will need the following result from the graph theory. Lemma 4.14 Let (cid:96) ∈ N and let A = {Aα : α ∈ I} be a family of subsets of R such that every set Aα ∈ A has common points with at most (cid:96) sets Aβ ∈ A. Then there exist subfamilies Ai ⊂ A, i = 1, ..., n with n ≤ (cid:96) + 1, each consisting of pairwise n(cid:91) disjoint sets such that A = A j. i=1 (cid:3) Proof. The proof is immediate from the following well-known statement (see, e.g. [39]): Every graph can be colored with one more color than the maximum vertex degree. j=1 ⊂ E. Let P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} be the Whitney (m − 1)-field determined by (4.18). We recall that our aim is to prove inequality (4.19) for an arbitrary finite strictly increasing sequence X = {x j}k We fix ε > 0 and apply Main Lemma 4.9 to the set X = {x1, ..., xk} and the Whitney (m− 1)-field P(m,E)[ f ]. The Main Lemma 4.9 produces a finite strictly increasing sequence V = {v j}(cid:96) j=1 ⊂ E and a mapping H : X → 2V which to every x ∈ X assigns m consecutive points of V possessing properties (1)-(5) of the Main Lemma. Using these objects, the sequence V and the mapping H, in the next two lemmas we prove the required inequality (4.19). 48 Lemma 4.15 Let k−1(cid:88) m−1(cid:88) j=1 i=0 A+ = H(x j+1)[ f ](x j) − L(i) L(i) H(x j)[ f ](x j)p (x j+1 − x j)(m−i)p−1 . (4.77) Then A+ ≤ C(m)p λp. (We recall that Assumption 4.8 holds for the function f .) Proof. Let I j be the smallest closed interval containing H(x j)∪ H(x j+1), j = 1, ..., k − 1. Clearly, diam I j = diam(H(x j) ∪ H(x j+1)) and x j+1, x j ∈ I j (because x j ∈ H(x j) and x j+1 ∈ H(x j+1), see property (2) of the Main Lemma 4.9). Furthermore, property (3) of the Main Lemma 4.9 tells us that Let us prove that for every j = 1, ..., k − 1, and every i = 0, ..., m − 1, the following inequality I j = [min H(x j), max H(x j+1)]. (4.78) H(x j+1)[ f ](x j) − L(i) L(i) H(x j)[ f ](x j)p max I j ≤ C(m)p L(i) H(x j+1)[ f ] − L(i) (diam I j)(m−i)p−1 H(x j)[ f ]p (x j+1 − x j)(m−i)p−1 (4.79) holds. Indeed, this inequality is obvious if H(x j) = H(x j+1). Prove it for every j ∈ {1, ..., k − 1} such that H(x j) (cid:44) H(x j+1). Properties (2) and (4) of the Main Lemma 4.9 tell us that x j ∈ H(x j), x j+1 ∈ H(x j+1), and diam H(x j) + diam H(x j+1) ≤ 2(m + 1)(x j+1 − x j) . Hence, diam I j = diam(H(x j) ∪ H(x j+1)) ≤ diam H(x j) + diam H(x j+1) + (x j+1 − x j) ≤ 2(m + 1)(x j+1 − x j) + (x j+1 − x j) = (2m + 3)(x j+1 − x j) proving (4.79). This inequality and definition (4.77) imply that k−1(cid:88) m−1(cid:88) j=1 i=0 A+ ≤ C(m)p max I j L(i) H(x j+1)[ f ] − L(i) (diam I j)(m−i)p−1 H(x j)[ f ]p . (4.80) Note that each summand in the right hand side of inequality (4.80) equals zero provided H(x j) = H(x j+1). Therefore, in our proof of the inequality A+ ≤ C(m)p λp, without loss of generality, we may assume that H(x j) (cid:44) H(x j+1) for all j = 1, ..., k − 1 . (4.81) Property (1) of the Main Lemma 4.9 tells us that for every j = 1, ..., k, the set H(x j) consists of m consecutive points of the sequence V = {vn}(cid:96) n=1 ⊂ E. 49 (4.82) Let n j be the index of the minimal point of H(x j) in the sequence V. Thus H(x j) = {vn j, ..., vn j+m−1}, j = 1, ..., k − 1. In particular, thanks to (4.78), I j = [vn j, vn j+1+m−1], j = 1, .., k − 1. (4.83) Let us apply Lemma 3.7 with k = m − 1 to the sequence Y = {vi}n j+1+m−1 , the sets S 1 = H(x j) = {vn j, ..., vn j+m−1}, S 2 = H(x j+1) = {vn j+1, ..., vn j+1+m−1}, i=n j and the closed interval I = I j = [vn j, vn j+1+m−1]. Let = {vn, ..., vn+m}, n j ≤ n ≤ n j+1 − 1 . S (n) j Lemma 3.7 tells us that L(i) H(x j+1)[ f ] − L(i) H(x j)[ f ]p ≤ ((m + 1)!)p (diam I j)(m−i)p−1 max I j ∆m f [S (n) j ]p diam S (n) j . n j+1(cid:88) n=n j This inequality and (4.80) imply that k−1(cid:88) n j+1(cid:88) A+ ≤ C(m)p (vn+m − vn)∆m f [vn, ..., vn+m]p. j=1 n=n j To apply Assumption 4.8 to the right hand side of this inequality and prove in this way the In required inequality A+ ≤ C(m)p λp, we need some additional properties of the intervals I j. particular, let us prove that each interval I j contains at most 4m elements of the sequence V, i.e., n j+1 + m − n j ≤ 4m for every j = 1, .., k − 1. (4.84) Indeed, let x j = vκ j and x j+1 = vκ j+1. Property (2) of the Main Lemma 4.9 tells us that 0 < κ j+1 − κ j ≤ 2m and x j ∈ H(x j), x j+1 ∈ H(x j+1). But #H(x j) = #H(x j+1) = m so that n j ≤ κ j ≤ n j + m − 1 and n j+1 ≤ κ j+1 ≤ n j+1 + m − 1. These inequalities imply that n j+1 + m − n j ≤ n j+1 + m − κ j + m − 1 ≤ n j+1 − (κ j+1 − 2m) + 2m − 1 ≤ 4m proving (4.84). Property (3) of Main Lemma 4.9 and (4.81) tell us that {vn j}k−1 j=1 is a strictly increasing subsequence of the sequence V. (4.85) Let I = {I j : j = 1, ..., k − 1}. Properties (4.83), (4.84) and (4.85) imply the following: every interval I j0 ∈ I has common points with at most 8m intervals I j ∈ I. This property and Lemma 50 4.14 tell us that there exist subfamilies Iν ⊂ I, ν = 1, ..., υ, with υ ≤ 8m + 1, each consisting of pairwise disjoint intervals, such that I = ∪{Iν : ν = 1, ..., υ}. This and (4.80) imply the following: A+ ≤ C(m)p Aν (4.86) where υ(cid:88) Aν = (cid:88) n j+1(cid:88) Aν ≤ (cid:96)−m(cid:88) j:I j∈Iν n=n j ν=1 (vn+m − vn)∆m f [vn, ..., vn+m]p. (vn+m − vn)∆m f [vn, ..., vn+m]p Since the intervals of each family Iν, ν = 1, ..., υ, are pairwise disjoint, the following inequality n=1 holds. (We recall that (cid:96) = #V, see (4.82).) Applying Assumption 4.8 to the right hand side of this inequality, we obtain that Aν ≤ λp for every ν = 1, ..., υ. This and (4.86) imply that A+ ≤ C(m)p υ λp ≤ (8m + 1) C(m)p λp (cid:3) proving the lemma. Lemma 4.16 Inequality (4.19) holds for every finite strictly increasing sequence X = {x j}k Proof. Using property (5) of the Main Lemma 4.9, let us replace the Hermite polynomials j=1 ⊂ E. {Px j : j = 1, ..., k} in the left hand side of inequality (4.19) with corresponding Lagrange polynomials LH(x j). For every j = 1, ..., k − 1 and every i = 0, ..., m − 1 we have H(x j)[ f ](x j) + L(i) P(i) x j (x j) − P(i) x j+1(x j) ≤ P(i) + L(i) x j (x j) − L(i) H(x j+1)[ f ](x j) − P(i) x j+1(x j). H(x j)[ f ](x j) − L(i) H(x j+1)[ f ](x j) Property (5) of the Main Lemma (see (4.21)) tells us that P(i) x j (x j) − L(i) P(i) x j (x j) − P(i) H(x j)[ f ](x j) + L(i) x j+1(x j) ≤ L(i) H(x j+1)[ f ](x j) − P(i) x j+1(x j) ≤ 2ε H(x j)[ f ](x j) − L(i) H(x j+1)[ f ](x j) + 2ε . proving that Hence, x j (x j) − P(i) P(i) x j+1(x j)p ≤ 2p L(i) H(x j)[ f ](x j) − L(i) H(x j+1)[ f ](x j)p + 4pεp. (4.87) Let A1 be the left hand side of inequality (4.19), i.e., k−1(cid:88) m−1(cid:88) j=1 i=0 A1 = x j (x j)p P(i) x j+1(x j) − P(i) (x j+1 − x j)(m−i)p−1 , 51 and let k−1(cid:88) m−1(cid:88) j=1 i=0 A2 = 4p (x j+1 − x j)1−(m−i)p. We apply inequality (4.87) to each summand from the right hand side of (4.19), and get the following estimate of A1: A1 ≤ 2p A+ + εp A2 . Here A+ is the quantity defined by (4.77). Lemma 4.15 tells us that A+ ≤ C(m)p λp so that A1 ≤ C(m)p λp + εp A2. Since ε is an arbitrary positive number, A1 ≤ C(m)p λp proving the required inequality (4.19). The proof of Lemma 4.16 is complete. We recall that the quantity Nm,p,E is defined by (4.2). We also recall that (cid:3) P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} is the Whitney (m − 1)-field determined by formulae (3.39)-(3.42), see Definition 3.14. We know that for every x ∈ E. We are in a position to complete the proof of Theorem 1.7. Proof of the sufficiency part of Theorem 1.7. Lemma 4.16 and (4.2) tell us that Px(x) = f (x) (cid:17) ≤ C(m) λ = C(m)Lm,p( f : E) . Nm,p,E P(m,E)[ f ] (cid:16) (4.88) (4.89) See (1.14) and (4.16). This inequality and the sufficiency part of Theorem 4.1 imply that (cid:107)P(m,E)[ f ](cid:107)m,p,E ≤ C(m)Nm,p,E (4.90) We recall that the quantity (cid:107) · (cid:107)m,p,E is defined by (3.1). This definition and inequality (4.90) tell P(m,E)[ f ] (cid:17) ≤ C(m)Lm,p( f : E) . (cid:16) us that there exists a function F ∈ Lm p (R) such that T m−1 x [F] = Px on E and (cid:107)F(cid:107)Lm p (R) ≤ 2(cid:107)P(m,E)[ f ](cid:107)m,p,E ≤ C(m)Lm,p( f : E) . Since Px(x) = f (x) on E, see (4.88), we have (4.91) (4.92) F(x) = T m−1 [F](x) = Px(x) = f (x) p (R) and FE = f proving that f ∈ Lm x Thus, F ∈ Lm inequality (4.92) imply that for all x ∈ E . p (R)E. Furthermore, definition (1.9) and (cid:107) f(cid:107)Lm p (R)E ≤ (cid:107)F(cid:107)Lm p (R) ≤ C(m)Lm,p( f : E) (4.93) proving the sufficiency. The proof of Theorem 1.7 is complete. (cid:3) 52 Remark 4.17 Note that the extension algorithm providing given a function f on E the extension F from (4.91), includes three main steps: Step 1. We construct the family of sets {Sx : x ∈ E} and the family of points {sx : x ∈ E} Step 2. At this step we construct the Whitney (m − 1)-field P(m,E)[ f ] = {Px ∈ Pm−1 : x ∈ E} satisfying conditions of Proposition 3.11 and Proposition 3.12. satisfying conditions (i), (ii) of Definition 3.14. Step 3. We define the extension F by the formula (4.8). We denote the extension F by Clearly, F depends on f linearly proving that ExtE(· : Lm Theorem 1.7 states that its operator norm is bounded by a constant depending only on m. p (R)) is a linear extension operator. (cid:67) F = ExtE( f : Lm p (R)). (4.94) p -functions on increasing sequences of the real line. 4.4. Lm In this section we give an alternative proof of of Theorem 1.6. (Necessity.) The necessity part of Theorem 1.6 is immediate from the necessity part of Theorem 1.7. More specifically, let p ∈ (1,∞), and let (cid:96)1, (cid:96)2 ∈ Z ∪ {±∞}, (cid:96)1 + m ≤ (cid:96)2. Let f be a function Definitions (1.13) and (1.14) imply that (cid:101)Lm,p( f defined on a strictly increasing sequence of points E = {xi}(cid:96)2 necessity part of Theorem 1.7 and equivalence (1.15) tell us that for every function f ∈ Lm the following inequality holds. Hence, (cid:101)Lm,p( f : E) ≤ C(m)(cid:107) f(cid:107)Lm (cid:3) (Sufficiency) We assume that f is a function defined on a strictly increasing sequence E = {xi}(cid:96)2 with (cid:96)1, (cid:96)2 ∈ Z ∪ {±∞}, (cid:96)1 + m ≤ (cid:96)2 such that This inequality and definition (1.13) of (cid:101)Lm,p( f : E) enable us to make the following assumption. (cid:101)λ = (cid:101)Lm,p( f : E) < ∞. : E). On the other hand, the p (R)E p (R)E proving the necessity part of Theorem 1.6. Lm,p( f : E) ≤ C(m)(cid:107) f(cid:107)Lm (4.95) . i=(cid:96)1 : E) ≤ Lm,p( f p (R)E i=(cid:96)1 Assumption 4.18 The following inequality (cid:96)2−m(cid:88) (xi+m − xi)∆m f [xi, ..., xi+m]p ≤(cid:101)λp holds. i=(cid:96)1 Our task is to show that there exists of a function p (R) such that FE = f F ∈ Lm p (R) ≤ C(m)(cid:101)λ. and (cid:107)F(cid:107)Lm We prove the existence of F by a certain modification of the proof of the sufficiency part of Theorem 1.7 given in Sections 4.1 - 4.3. We will see that, for the case of sequences that proof can be simplified considerably. In particular, in this case we can replace the Main Lemma 4.9 (the most technically difficult part of our proof) with its simpler version given in Lemma 4.13. We begin with a version of Theorem 4.1 for the case of sequences. 53 Theorem 4.19 Let m ∈ N, p ∈ (1,∞), (cid:96)1, (cid:96)2 ∈ Z∪{±∞}, (cid:96)1+m ≤ (cid:96)2, and let E = {xi}(cid:96)2 be a strictly increasing sequence of points. Suppose we are given a Whitney (m − 1)-field P = {Px : x ∈ E} defined on E. i=(cid:96)1 p (R) which agrees with the field P on E if and only if Then there exists a Cm−1-function F ∈ Lm the following quantity (cid:101)Nm,p,E(P) = m−1(cid:88)  (cid:96)2−1(cid:88) P(i) x j (x j) − P(i) x j+1(x j)p (x j+1 − x j)(m−i)p−1 1/p (4.96) is finite. Furthermore, (cid:107)P(cid:107)m,p,E ∼ (cid:101)Nm,p,E(P) with constants of equivalence depending only on m. (Recall that the quantity (cid:107)P(cid:107)m,p,E is defined in (3.1).) Proof. The necessity part of this theorem and the inequality (cid:101)Nm,p,E(P) ≤ C(m)(cid:107)P(cid:107)m,p,E are j=(cid:96)1 i=0 immediate from the necessity part of Theorem 4.1. We prove the sufficiency using the same extension construction as in the proof of the sufficiency part of Theorem 4.1. More specifically, we introduce a family of open intervals by letting J j = (x j, x j+1), j = (cid:96)1, ..., (cid:96)2 − 1. Then we define a smooth function F interpolating f on E as follows: If (cid:96)1 > −∞, we introduce an interval J(cid:9) = (−∞, (cid:96)1), and set If (cid:96)2 < ∞, we put J⊕ = ((cid:96)2, +∞), and set FJ(cid:9) = Px(cid:96)1 . (4.97) FJ⊕ = Px(cid:96)2 . (4.98) We define F on each intervals J j, j = (cid:96)1, ..., (cid:96)2−1, in the same way as in (4.7): we let HJ j ∈ P2m−1 denote the Hermite polynomial such that x j (x j) and H(i) i = 0, ..., m − 1. H(i) J j (x j) = P(i) J j (x j+1) = P(i) x j+1(x j+1) (4.99) for all (We recall that the existence and uniqueness of the Hermite polynomial HJ j satisfying (4.99) fol- lows from a general result proven in [2, Ch. 2, Section 11].) Finally, we set F[x j,x j+1] = HJ j, j = (cid:96)1, ..., (cid:96)2 − 1. (4.100) Now, the function F is well defined on all of R. Furthermore, definitions (4.97), (4.98), (4.100) and (4.99) tell us that F is Cm−1-smooth function on R which agrees with the Whitney (m− 1)-field P = {Px : x ∈ E} on E. Note that F is a piecewise polynomial function which coincides with a polynomial of degree at most 2m− 1 on each subinterval J j = (x j, x j+1). Let us estimate the Lm p (R)-seminorm of F. Clearly, F(m)J(cid:9) = F(m)J⊕ ≡ 0 because the restrictions of F to J(cid:9) and to J⊕ are polynomials of degree at most m − 1. In turn, Lemma 4.5 tells us that F(m)(x) = H(m) J j (x) ≤ C(m) x j (x j) − P(i) P(i) (x j+1 − x j)m−i x j+1(x j) m−1(cid:88) i=0 54 so that (cid:90) R J j F(m)(x)p dx = (cid:96)2(cid:88) (cid:90) j=(cid:96)1 J j F(m)(x)p dx ≤ C(m)p m−1(cid:88) (cid:96)2(cid:88) x j (x j) − P(i) P(i) x j+1(x j)p (x j+1 − x j)(m−i)p−1 p (R) ≤ C(m)(cid:101)Nm,p,E(P). j=(cid:96)1 i=0 . for every j ∈ Z, (cid:96)1 ≤ j < (cid:96)2 and every x ∈ J j = (x j, x j+1). Hence, (cid:90) F(m)(x)p dx ≤ C(m)p x j (x j) − P(i) x j+1(x j)p P(i) (x j+1 − x j)(m−i)p−1 m−1(cid:88) i=0 This inequality and definition (4.96) imply that (cid:107)F(cid:107)Lm quantity (cid:107)P(cid:107)m,p,E implies that (cid:107)P(cid:107)m,p,E ≤ C(m, p)(cid:101)Nm,p,E(P), proving the sufficiency part of Theorem Since F agrees with the Whitney (m − 1)-field P = {Px : x ∈ E} on E, definition (3.1) of the 4.19. The proof of Theorem 4.19 is complete. (cid:3) Remark 4.20 As we have noted above, the extension F is a piecewise polynomial function which coincides with a polynomial of degree at most 2m − 1 on each interval (xi, xi+1). This enables us to reformulate this property of F in terms of Spline Theory as follows: The extension F is an interpolating Cm−1-smooth spline of order 2m with knots {xi}(cid:96)2 (cid:67) . We continue the proof of the sufficiency part of Theorem 1.6 as follows. Given x ∈ E we let Sx and sx denote the subset of E and the point in E determined by formulae (3.33) and (3.34) respectively. We know that the sets {Sx : x ∈ E} and the points {sx : x ∈ E} possesses properties given in Proposition 3.11 and Proposition 3.12. Furthermore, Remark 3.13 tells us that if i=(cid:96)1 E = {xi}(cid:96)2 i=(cid:96)1 , (cid:96)1, (cid:96)2 ∈ Z ∪ {−∞, +∞}, (cid:96)1 + m ≤ (cid:96)2, is a strictly increasing sequence of points, each set Sxi consists of m consecutive points of E. Now, let us define a Whitney (m − 1)-field P(m,E)[ f ] = {Px : x ∈ E} on E by letting (4.101) Thus Px ∈ Pm−1(R), and Px(y) = f (y) for all y ∈ Sx. In particular, Px(x) = f (x) (because x ∈ Sx). We see that the Whitney (m − 1)-field P(m,E)[ f ] = {Px : x ∈ E} is defined in the same way as in Px = LSx[ f ], the proof of the sufficiency part of Theorem 1.7, see formulae (4.18) and definition (3.42). x ∈ E. The following lemma is an analogue of Lemma 4.16. Lemma 4.21 Let P(m,E)[ f ] be the Whitney (m − 1)-field determined by (4.101). Then Recall that the quantities (cid:101)Nm,p,E and(cid:101)λ are determined by (4.96) and (4.95) respectively. (4.102) Proof. We follow the same scheme as in the proof of Lemma 4.16. In particular, we use the version of the Main Lemma 4.9 for sequences given in Lemma 4.13. 55 (cid:101)Nm,p,E(P(m,E)[ f ]) ≤ C(m)(cid:101)λ. Hence, (cid:101)Nm,p,E(P) =  (cid:96)2−1(cid:88) j=(cid:96)1 m−1(cid:88) i=0 1/p LH(x j+1)[ f ](x j) − LH(x j)[ f ](x j)p (x j+1 − x j)(m−i)p−1 . (4.103) See (4.96). This equality and definition (4.77) tell us that(cid:101)Nm,p,E(P)p is an analog of the quantity A+ defined by formula (4.77). This enables us to prove the present lemma by repeating the proof of Lemma 4.15 (with minor changes in the notation). Of course, in this modification of the proof of Lemma 4.15 we use the Lemma 4.13 rather than the Main Lemma 4.9, and Assumption 4.18 rather than Assumption 4.8. We leave the details of this obvious modification to the interested reader. For the reader's convenience, in this long version of our paper we give a complete proof of Prove that for every j ∈ Z, (cid:96)1 ≤ i ≤ (cid:96)2, every i = 0, ..., m − 1, the following inequality Lemma 4.21. L(i) H(x j+1)[ f ](x j) − L(i) H(x j)[ f ](x j)p (x j+1 − x j)(m−i)p−1 ≤ C(m)p L(i) H(x j+1)[ f ](x j) − L(i) H(x j)[ f ](x j)p (diam I j)(m−i)p−1 (4.104) holds. Here I j = [min(H(x j) ∪ H(x j+1)), max(H(x j) ∪ H(x j+1))] is the smallest closed interval containing H(x j)∪H(x j+1). Clearly, diam I j = diam(H(x j)∪H(x j+1)). Furthermore, property (3) of Lemma 4.13 tells us that I j = [min H(x j), max H(x j+1)]. (4.105) Note that inequality (4.104) is obvious whenever H(x j) = H(x j+1), so we may assume that In this case properties (2) and (4) of Lemma 4.13 tell us that x j ∈ H(x j), H(x j) (cid:44) H(x j+1). x j+1 ∈ H(x j+1) and diam H(x j) + diam H(x j+1) ≤ 2(m + 1)(x j+1 − x j) . Hence, diam I j = diam(H(x j) ∪ H(x j+1)) ≤ diam H(x j) + diam H(x j+1) + (x j+1 − x j) ≤ 2m (x j+1 − x j) + (x j+1 − x j) = (2m + 1)(x j+1 − x j) We recall that, whenever E is a sequence of points, H(x) = Sx for each x ∈ E, see Lemma 4.13. This property and definition (4.101) imply that Px j = LS x j [ f ] = LH(x j)[ f ] for every j ∈ Z, (cid:96)1 ≤ j ≤ (cid:96)2. proving (4.104). In turn, this inequality implies that (cid:96)2−1(cid:88) m−1(cid:88) (cid:101)Nm,p,E(P)p ≤ C(m)p See (4.103). j=(cid:96)1 i=0 max I j L(i) H(x j+1)[ f ] − L(i) (diam I j)(m−i)p−1 H(x j)[ f ]p . (4.106) 56 Note that each summand in the right hand side of inequality (4.106) equals zero provided H(x j) = H(x j+1). Therefore, in the proof of inequality (4.102), without loss of generality, one may assume that H(x j) (cid:44) H(x j+1) for all j = (cid:96)1, ..., (cid:96)2 − 1. Property (1) of Lemma 4.13 tells us that for every j ∈ Z, (cid:96)1 ≤ i ≤ (cid:96)2, the set H(x j) consists of m consecutive points of the sequence E = {x j}(cid:96)2 . j=(cid:96)1 Let n j be the index of the minimal point of H(x j) in the sequence E. Thus H(x j) = {xn j, ..., xn j+m−1}, (cid:96)1 ≤ j ≤ (cid:96)2 . In particular, thanks to (4.105), I j = [xn j, xn j+1+m−1], (cid:96)1 ≤ j ≤ (cid:96)2. Let us apply Lemma 3.7 with k = m − 1 to the sequence Y = {xi}n j+1+m−1 , the sets S 1 = H(x j) = {xn j, ..., xn j+m−1}, S 2 = H(x j+1) = {xn j+1, ..., xn j+1+m−1}, i=n j and the closed interval I = I j = [xn j, xn j+1+m−1]. Let Lemma 3.7 tells us that for every i = 0, ..., m − 1, the following inequality = {xn, ..., xn+m}, n j ≤ n ≤ n j+1 − 1. S (n) j n j+1(cid:88) n=n j L(i) H(x j+1)[ f ] − L(i) H(x j)[ f ]p ≤ ((m + 1)!)p (diam I j)(m−i)p−1 max I j ∆m f [S (n) j ]p diam S (n) j holds. This inequality and (4.106) imply that (cid:101)Nm,p,E(P)p ≤ C(m)p (cid:96)2−1(cid:88) n j+1(cid:88) j=(cid:96)1 n=n j (xn+m − xn)∆m f [xn, ..., xn+m]p. To apply Assumption 4.18 to the right hand side of this inequality and prove in this way the required inequality (4.102) we need some additional properties of the intervals I j. In particular, since each set H(x j) consists of m consecutive points of E, each interval I j (see (4.105)) contains at most 2m elements of the sequence E. This property implies the following: Let I = {I j : (cid:96)1 ≤ j ≤ (cid:96)2}. Then every interval I j0 ∈ I has common points with at most 4m intervals I j ∈ I. In turn, this property and Lemma 4.14 tell us that there exist subfamilies Iν ⊂ I, ν = 1, ..., υ, with υ ≤ 4m + 1, each consisting of pairwise disjoint intervals, such that I = ∪{Iν : ν = 1, ..., υ}. This and (4.106) imply that Aν (4.107) where inequality Aν = (xn+m − xn)∆m f [xn, ..., xn+m]p. υ(cid:88) ν=1 (cid:101)Nm,p,E(P)p ≤ C(m)p (cid:88) n j+1(cid:88) Aν ≤ (cid:96)−m(cid:88) j:I j∈Iν n=n j n=(cid:96)1 57 (xn+m − xn)∆m f [xn, ..., xn+m]p Since the intervals of the family Iν are pairwise disjoint, for each ν = 1, ..., υ, the following holds. Applying Assumption 4.18 to the right hand side of this inequality, we obtain that Aν ≤(cid:101)λp for every ν = 1, ..., υ. This and (4.107) imply that (cid:101)Nm,p,E(P)p ≤ C(m)p υ(cid:101)λp ≤ (4m + 1) C(m)p(cid:101)λp proving the lemma. (cid:3) We complete the proof of the sufficiency part of Theorem 1.6 in the same way as we did this at the end of the proof of Theorem 1.7. More specifically, we simply replace in inequalities (4.89), (4.90), (4.92) and (4.93) the quantities Nm,p,E and Lm,p( f : E) with the quantities(cid:101)Nm,p,E (see (4.96)) and (cid:101)Lm,p( f : E) (see (1.13)) respectively. This leads us to the existence of a function F ∈ Lm such that FE = f and (cid:107)F(cid:107)Lm p (R) ≤ C(m)(cid:101)Lm,p( f : E) proving the sufficiency. p (R) (cid:3) The proof of Theorem 1.6 is complete. Below we give an alternative proof of the sufficiency part of Theorem 1.6 by showing that the (weaker) hypothesis of Theorem 1.6 implies the (stronger) hypothesis of Theorem 1.7 provided E is a sequences in R. (cid:3) We will need the following lemma. Lemma 4.22 Let k, (cid:96) ∈ N, k ≤ (cid:96). Let {s j}(cid:96) be a strictly increasing subsequence of the sequence {s j}(cid:96) Then for every function g defined on the set S = {s0, ..., s(cid:96)} the following inequality j=0 such that t0 = s0 and tk = s(cid:96). j=0 be a strictly increasing sequence in R, and let {ti}k i=0 (cid:96)−k(cid:88) (tk − t0)∆kg[t0, ..., tk]p ≤ kp−1 (s j+k − s j)∆kg[s j, ..., s j+k]p j=0 holds. Proof. We apply the statement ((cid:70)7), (ii), of Section 2.1 to the sets {s0, ..., s(cid:96)} and T = {t0, ..., tk} (with n = (cid:96)). This statement tells us that there exist numbers β j ∈ [0, (s j+k − s j)/(s(cid:96) − s0)], j = 0, ..., (cid:96) − k, such that the following equality ∆kg[t0, ..., tk] = holds. Hence, ∆kg[t0, ..., tk] ≤ (s(cid:96) − s0)−1 βi ∆kg[s j, ..., s j+k] (s j+k − s j)∆kg[s j, ..., s j+k]. (cid:96)−k(cid:88) (cid:96)−k(cid:88) j=0 j=0 58 Therefore, by Holder's inequality, ∆kg[t0, ..., tm]p ≤ (sk − s0)−p (s j+k − s j) p−1 (cid:96)−k(cid:88) · (cid:96)−k(cid:88) j=0 (s j+k − s j)∆kg[s j, ..., s j+k]p (s j+k − s j)∆kg[s j, ..., s j+k]p  (cid:96)−k(cid:88) (cid:96)−k(cid:88) j=0 ≤ kp−1(sk − s0)−p · (sk − s0)p−1 = kp−1(tk − t0)−1 j=0 (s j+k − s j)∆kg[s j, ..., s j+k]p proving the lemma. (cid:3) j=0 i=(cid:96)1 be a strictly increasing sequence of points. The sufficiency part of Theorem 1.6. An alternative proof. Let m ∈ N, p ∈ (1,∞), (cid:96)1, (cid:96)2 ∈ Let f be a function on E such that Assumption 4.18 holds with(cid:101)λ determined by (4.95). Z ∪ {±∞}, (cid:96)1 + m ≤ (cid:96)2, and let E = {xi}(cid:96)2 for every n ∈ N, n ≥ m, and every strictly increasing subsequence {yν}n E = {xi}(cid:96)2 (yν+m − yν)∆m f [yν, ..., yν+m]p ≤ C(m)p(cid:101)λp Let us prove that f satisfies the hypothesis of Theorem 1.7. More specifically, let us show that ν=0 of the sequence the following inequality n−m(cid:88) = {xiν}n (4.108) A = i=(cid:96)1 ν=0 ν=0 holds. Fix ν ∈ {0, ..., n − m} and set s j = xiν+ j, j = 0, ..., (cid:96)ν, with (cid:96)ν = iν+m − iν, and ti = yν+i, i = 0, ..., m. Clearly, (cid:96)ν ≥ m because {yν} is a subsequence of {xi}. It is also clear that t0 = s0 and tm = s(cid:96)ν. In these settings, Iν = (yν+m − yν)∆m f [yν, ..., yν+m]p = (tm − t0)∆m f [t0, ..., tm]p . Lemma 4.22 tells us that Iν ≤ mp−1 j=0 (cid:96)ν−m(cid:88) (cid:96)ν−m(cid:88) (cid:96)ν−m(cid:88) j=0 = mp−1 n−m(cid:88) This inequality and (4.108) imply that (s j+m − s j)∆m f [s j, ..., s j+m]p (xiν+ j+m − xiν+ j)∆m f [xiν+ j, ..., xiν+ j+m]p . A ≤ mp−1 (xiν+ j+m − xiν+ j)∆m f [xiν+ j, ..., xiν+ j+m]p. (4.109) ν=0 j=0 59 Let Tν = [yν, yν+m], and let T = {Tν : ν = 0, ..., n − m}. Clearly, each interval Tν0 ∈ T has at most 2m + 2 common points with each interval Tν ∈ T . (4.110) This observation and Lemma 4.14 imply the existence of subfamilies Tk ⊂ T , k = 1, ..., κ, with κ ≤ 2m + 3, each consisting of pairwise disjoint intervals, such that T = ∪{Tk : k = 1, ..., κ}. This property and (4.109) tell us that A ≤ mp−1 Ak (4.111) κ(cid:88) k=1 (cid:88) ν:Tν∈Tk (cid:96)ν−m(cid:88) ivn−m(cid:88) j=0 j=iv0 where Ak = (xiν+ j+m − xiν+ j)∆m f [xiν+ j, ..., xiν+ j+m]p. Since the intervals of each family Tk, k = 1, ..., κ, are pairwise disjoint, the following inequality Ak ≤ (x j+m − x j)∆m f [x j, ..., x j+m]p holds. We apply Assumption 4.18 to the right hand side of this inequality and obtain that Ak ≤(cid:101)λp for every k = 1, ..., κ. This and (4.111) imply that A ≤ mp−1 κ(cid:101)λp ≤ (2m + 2) mp−1(cid:101)λp proving (4.108). In turn, (4.108) and definition (1.14) of Lm,p( f : E) tell us that proving that f satisfies the hypothesis of Theorem 1.7. The sufficiency part of this theorem pro- duces the required function F ∈ Lm Lm,p( f : E) ≤ C(m)(cid:101)λ p (R) ≤ C(m)(cid:101)λ = C(m)(cid:101)Lm,p( f : E) p (R) with (cid:107)F(cid:107)Lm (see (4.95)) whose restriction to E coincides with f . The alternative proof of the sufficiency part of Theorem 1.6 is complete. (cid:3) 4.5. The variational criterion and the sharp maximal function-type criterion. In this section we compare the variational trace criterion for the space Lm p (R) given in Theorem 1.7 with the trace criterion in terms of sharp maximal functions, see Theorem 1.8. Statement 4.23 Let p ∈ [1,∞), and let E be a closed subset of R with #E ≥ m + 1. Let f be a function on E. Then the following inequality Lm,p( f : E) ≤ C(m)(cid:107) (∆m f )(cid:93) E (cid:107)Lp(R) (4.112) holds. See (1.14). 60 sequence of points in E. Let i ∈ {0, ..., n − m}, and let S i = {xi, ..., xi+m}. Proof. Let n ∈ N, n ≥ m + 1, and let S = {x0, ..., xn} ⊂ E, x0 < ... < xn, be a strictly increasing Clearly, x − xi + x − xi+m = xi+m − xi for every x ∈ [xi, xi+m]. In this case definition (1.16) tells us that Hence, ∆m f [S i] ≤ [(∆m f )(cid:93) E](x) ∆m f [S i]p ≤ [(∆m f )(cid:93) for every x ∈ [xi, xi+m]. E]p(x) on [xi, xi+m]. Integrating this inequality on the interval [xi, xi+m] (with respect to x), we obtain that xi+m(cid:90) xi (cid:90) (xi+m − xi) ∆m f [S i]p ≤ [(∆m f )(cid:93) E]p(x) dx . n−m(cid:88) i=0 This implies the following inequality: Here, given x ∈ R the function ϕ(x) [(∆m f )(cid:93) E]p(x) dx . (xi+m − xi) ∆m f [S i]p ≤ n−m(cid:88) ϕ(x) = R χTi(x) denotes the number of intervals from the family i=0 T = {Ti = [xi, xi+m] : i = 0, ..., n − m} containing x. It follows from (4.110) that ϕ(x) ≤ 2m + 3 on R, so that n−m(cid:88) (cid:90) (xi+m − xi) ∆m f [S i]p ≤ (2m + 3) [(∆m f )(cid:93) E]p(x) dx = (2m + 3)(cid:107) (∆m f )(cid:93) E (cid:107)p Lp(R) . i=0 R Taking the supremum in the left hand side of this inequality over all strictly increasing sequences {x0, ..., xn} ⊂ E with n ≥ m, and recalling definition (1.14) of the quantity Lm,p( f : E), we obtain the required inequality (4.112). The proof of the statement is complete. (cid:3) Remark 4.24 Inequality (4.112) tells us that the necessity part of Theorem 1.8 implies the neces- sity part of Theorem 1.7, and the sufficiency part of Theorem 1.7 implies the sufficiency part of Theorem 1.8. However, for the reader's convenience, in Sections 2-4 we present the direct and independent proofs of the necessity and sufficiency parts of these theorems. (cid:67) 5. Extension criteria for Sobolev Wm p -functions. In this section we prove Theorem 1.2 (see Sections 5.1, 5.2) and Theorem 1.3 (Section 5.3). Everywhere in this section we assume that m is a positive integer and p ∈ (1,∞). In Section 5.1 and 5.2 we also assume that E is a closed subset of R containing at least m + 1 points. 61 p -traces: necessity. 5.1. The variational criterion for Wm In this section we proof the necessity part of Theorem 1.2. p (R) be a function such that FE = f . Let n ≥ m and let p (R)E, and let F ∈ Wm Let f ∈ Wm {x0, ..., xn} be a finite strictly increasing sequences in E. We have to prove that for each k = 0, ..., m, the following inequality n−k(cid:88) (cid:12)(cid:12)(cid:12)p ≤ C(m)p (cid:107)F(cid:107)p min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] n−m(cid:88) (cid:12)(cid:12)(cid:12)p ≤ min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+m] i=0 Lk = n−m(cid:88) holds. Note that Lm = (xi+m − xi) ∆m f [xi, ..., xi+m]p . Wm p (R) (5.1) i=0 i=0 This inequality and inequality (2.15) tell us that Lm ≤ 2p (cid:107)F(cid:107)p p (R) ≤ 2p (cid:107)F(cid:107)p Lm Wm p (R) proving (5.1) for k = m. We turn to the proof of inequality (5.1) for k ∈ {0, ..., m − 1}. Lemma 5.1 Let k ∈ {0, ..., m − 1}, and let S be a (k + 1)-point subset of a closed interval I ⊂ R (bounded or unbounded). Let G ∈ Ck[I] and let G(k) be absolutely continuous on I. Then for every q ∈ [1,∞) the following inequality (cid:90) (cid:16)G(k)(y)q + G(k+1)(y)q(cid:17) dy (5.2) min{1,I} · ∆kG[S ]q ≤ 2q holds. I Proof. Let S = {y0, ..., yk} with y0 < ... < yk. Property (2.6) tells us that there exists s ∈ [y0, yk] such that Let t = min{1,I}. Since s ∈ I and t ≤ I, there exists a closed interval J ⊂ I with J = t such that s ∈ J. Then, for every y ∈ J, k! ∆kG[S ] = G(k)(s). ∆kG[S ]q ≤ G(k)(s)q ≤ 2q G(k)(s) − G(k)(y)q + 2q G(k)(y)q ≤ 2q G(k+1)(x) dx + 2q G(k)(y)q. q  (cid:90) J (cid:90) Hence, by Holder's inequality, ∆kG[S ]q ≤ 2q Jq−1 (cid:90) G(k+1)(x)qdx + 2q G(k)(y)q, y ∈ J. Integrating this inequality on J with respect to y, we obtain that G(k+1)(x)qdx + 2q J∆kG[S ]q ≤ 2q Jq G(k)(y)q dy. J (cid:90) J J 62 Since J = t = min{1,I} ≤ 1 and J ⊂ I, this inequality implies (5.2) proving the lemma. Fix k ∈ {0, ..., m − 1}. Let i ∈ {0, ..., n − k}, and let Ti = [xi, xi+m] and (cid:3) ti = min{1,Ti} = min{1, xi+m − xi}. We recall convention (1.2) which tells us that (5.3) Let We apply Lemma 5.1 to G = F, t = ti, q = p, and S = {xi, ..., xi+k} and obtain that . xi+m = +∞ if i + m > n. (cid:90) Ai ≤ 2p (cid:12)(cid:12)(cid:12)p Ai = min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:16)F(k)(y)p + F(k+1)(y)p(cid:17) n−k(cid:88) (cid:16)F(k)(y)p + F(k+1)(y)p(cid:17) n−k(cid:88) Ai ≤ 2p (cid:90) (cid:16)F(k)(y)p + F(k+1)(y)p(cid:17) (cid:90) dy. i=0 i=0 Ti Ti = 2p ϕ(y) dy. dy Hence, Lk = Here, given y ∈ R the function R n−k(cid:88) ϕ(y) = χTi(y) i=0 denotes the number of intervals from the family T = {Ti : i = 0, ..., n − k} containing y. Let us see that ϕ(y) ≤ 2m for every y ∈ R. Indeed, thanks to (5.3), the number of intervals Ti = [xi, xi+m], 0 ≤ i ≤ n, such that xi+m > xn is bounded by m. Let now y ∈ Ti = [xi, xi+m] where 0 ≤ i ≤ i + m ≤ n. Thus xi ≤ y ≤ xi+m. Let y ∈ [x j, x j+1] for some j ∈ {0, ..., n}. Then i ≤ j ≤ i + m− 1 so that j− m + 1 ≤ i ≤ j. Clearly, the number of indexes i satisfying these inequalities is bounded by m. This proves that 0 ≤ ϕ(y) ≤ m + m = 2m for each y ∈ R. Hence, (cid:90) (cid:16)F(k)(y)p + F(k+1)(y)p(cid:17) Lk ≤ 2p 2m dy ≤ C(m)p (cid:107)F(cid:107)p Wm p (R) R proving (5.1). This inequality and definition (1.3) tell us that Wm,p( f : E) ≤ C(m)(cid:107)F(cid:107)Wm p (R) for every F ∈ Wm p (R) such that FE = f . Hence, Wm,p( f : E) ≤ C(m)(cid:107) f(cid:107)Wm p (R)E , and the proof of the necessity part of Theorem 1.2 is complete. (cid:3) 63 5.2. The variational criterion for Wm p -traces: sufficiency. In this section we proof the sufficiency part of Theorem 1.2. Let f be a function on E such that λ = Wm,p( f : E) < ∞. (5.4) See definition (1.3). This definition enables us to make the following assumption. Assumption 5.2 For every finite strictly increasing sequence {xi}n inequality i=0 ⊂ E, n ≥ m, the following m(cid:88) n−k(cid:88) min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤ λp (5.5) k=0 i=0 holds. Our aim is to prove that f ∈ Wm p (R)E and (cid:107) f(cid:107)Wm p (R)E ≤ C λ where C is a constant depending only on m. In this section we will need the following extended version of convention (1.2). Convention 5.3 Given a finite strictly increasing sequence {yi}n i=0 ⊂ R we put j > n y j = −∞ if y j = +∞ if j < 0, and (as in (1.2)). Our proof of the sufficiency relies on a series of auxiliary lemmas. Lemma 5.4 Let n ∈ N and let S = {y0, ...yn} where {yi}n in R such that diam S = yn − y0 ≥ 1. Let h be a function on S . Then there exist k ∈ {0, ..., n − 1} and i ∈ {0, ..., n − k} such that yi+k − yi ≤ 1 and i=0 is a strictly increasing sequence of points ∆nh[S ] ≤ 2n ∆kh[yi, ..., yi+k]/ diam S . (5.6) Furthermore, either i + k + 1 ≤ n and (5.7) Proof. We proceed by induction on n. Let n = 1, and let S = {y0, y1} where y0 < y1 and y1 − y0 ≥ 1. Then for every function h on S the following inequality yi+k+1 − yi ≥ 1, or i ≥ 1 and yi+k − yi−1 ≥ 1 . ∆1h[y0, y1] = h(y1) − h(y0) y1 − y0 ≤ h(y1) + h(y0) y1 − y0 ≤ 2 max{h(y0),h(y1)} y1 − y0 holds. Let us pick i ∈ {0, 1} such that h(yi) = max{h(y0),h(y1)}. Then ∆1h[y0, y1] ≤ 2 max{∆0h[y0],∆0h[y1]} y1 − y0 64 = 2∆0h[yi]/(y1 − y0) proving (5.6) for n = 1 with k = 0. It is also clear that i + k + 1 ≤ n and yi+k+1 − yi ≥ 1 if i = 0, and i ≥ 1 and yi+k − yi−1 ≥ 1 if i = 1 proving (5.7) and the lemma for n = 1. For the induction step, we fix n ≥ 1 and suppose the lemma holds for n; we then prove it for Let {xi}n+1 i=0 be a strictly increasing sequence in R such that xn+1 − x0 ≥ 1, and let h : S → R be a ∆n+1h[S ] = ∆n+1h[x0, ..., xn+1] = (∆nh[x1, ..., xn+1] − ∆nh[x0, ..., xn]) /(xn+1 − x0) n + 1. function on the set S = {x0, ...xn+1}. Then, thanks to (2.3), so that ∆n+1h[S ] ≤ (∆nh[x1, ..., xn+1] + ∆nh[x0, ..., xn])/ diam S . (5.8) Let If diam S 1 = xn − x0 ≥ 1, then, by the induction hypothesis, there exists k1 ∈ {0, ..., n − 1} and i1 ∈ {0, ..., n − k1} such that xi1+k1 − xi1 ≤ 1 and S 1 = {x0, ..., xn} and S 2 = {x1, ..., xn+1} . ∆nh[S 1] ≤ 2n ∆k1h[xi1, ..., xi1+k1]/ diam S 1 ≤ 2n ∆k1h[xi1, ..., xi1+k1] . (5.9) Moreover, i1 + k1 + 1 ≤ n and either (5.10) Clearly, if diam S 1 ≤ 1, then the inequality xi1+k1 − xi1 ≤ 1 and (5.9) hold provided k1 = n and i1 = 0. It is also clear that in this case xi1+k1+1 − xi1 ≥ 1, or i1 ≥ 1 and xi1+k1 − xi1−1 ≥ 1 . i1 + k1 + 1 ≤ n + 1 and xi1+k1+1 − xi1 = xn+1 − x0 ≥ 1 . This observation and (5.10) imply the following statement: k ∈ {0, ..., n} and i ∈ {0, ..., n + 1− k}, and either i + k + 1 ≤ n + 1 and xi+k+1 − xi ≥ 1, or i ≥ 1 and xi+k − xi−1 ≥ 1 (5.11) provided i = i1 and k = k1. In the same way we prove the existence of k2 ∈ {0, ..., n} and i2 ∈ {0, ..., n + 1 − k2} such that xi2+k2 − xi2 ≤ 1, ∆nh[S 2] ≤ 2n ∆k2h[xi2, ..., xi2+k2], (5.12) and (5.11) holds provided i = i2 and k = k2. Let us pick (cid:96) ∈ {1, 2} such that ∆k(cid:96)h[xi(cid:96), ..., xi(cid:96)+k(cid:96)] = max{∆k1h[xi1, ..., xi1+k1],∆k2h[xi2, ..., xi2+k2]} . Inequalities (5.8), (5.9) and (5.12) tell us that ∆n+1h[S ] ≤ 2n (∆k1h[xi1, ..., xi1+k1] + ∆k2h[xi2, ..., xi2+k2])/ diam S ≤ 2n+1∆k(cid:96)h[xi(cid:96), ..., xi(cid:96)+k(cid:96)] . diam S Thus, ∆n+1h[S ] ≤ 2n+1 ∆kh[xi, ..., xi+k]/ diam S provided i = i(cid:96) and k = k(cid:96). Furthermore, k ∈ {0, ..., n}, i ∈ {0, ..., n + 1− k}, and the statement (5.11) holds for these i and k. This proves the lemma for n + 1 completing the proof. (cid:3) The next two lemmas are variants of results proven in [22]. 65 Lemma 5.5 Let m ∈ N, p ∈ [1,∞) and k ∈ {0, ..., m − 1}. Let I ⊂ R be a bounded interval, and let z0, ..., zm−1 be m distinct points in I. Then for every function F ∈ Lm p (R) and every x ∈ I the following inequality (cid:90) I(m−k)p−1 m−1(cid:88) F(m)(s)p ds + I( j−k)p ∆ jF[z0, ..., z j]p (5.13) I j=k  F(k)(x)p ≤ C(m)p holds. Proof. Since FI ∈ C j[I] for every j ∈ {k, ..., m − 1}, property (2.6) implies the existence of a point y j ∈ I such that (5.14) (5.15) ∆ jF[z0, ..., z j] = 1 j! F( j)(y j). y(cid:90) The Newton-Leibniz formula tells us that F( j)(y) = F( j)(y j) + Hence, y j F( j)(y) ≤ F( j)(y j) + We apply this inequality to y = x and j = k, and obtain that Inequality (5.15) implies that F(k)(x) ≤ F(k)(yk) + (cid:90) I F(k+1)(s) ≤ F(k+1)(yk+1) + F(k+2)(t) dt I I so that F( j+1)(s) ds for every y ∈ I . (cid:90) y ∈ I. F( j+1)(s) ds, (cid:90) F(k+1)(s) ds . Repeating this inequality m − k − 1 times, we get I m−1(cid:88) j=k F(k)(x) ≤ F(k)(x) ≤ F(k)(yk) + IF(k+1)(yk+1) + I (cid:90)  (cid:90) I j−k F( j)(y j) + Im−k−1 I1−1/p I j−k F( j)(y j) + Im−k−1 m−1(cid:88) j=k I I F(k)(x) ≤ 66 Hence, by the Holder inequality, for every s ∈ I, (cid:90) F(k+2)(t) dt. F(m)(s) ds. 1/p F(m)(s)p ds so that F(k)(x)p ≤ (m − k + 1)p−1  m−1(cid:88) j=k I( j−k)p F( j)(y j)p + I(m−k)p−1  . F(m)(s)p ds (cid:90) I This inequality together with (5.14) imply (5.13) proving the lemma. Lemma 5.6 Let Z = {z0, ..., zm} be an (m + 1)-point subset of R, and let g be a function on Z. Then for every k = 0, ..., m − 1, and every S ⊂ Z with #S = k + 1 the following inequality (cid:3) ∆kg[S ] ≤ C(m) (diam Z) j−k · ∆ jg[z0, ..., z j] holds. j=k Proof. Let I = [min Z, max Z]. Then I ⊃ Z and I = diam Z. Let LZ[g] be the Lagrange polynomial of degree at most m which agrees with g at Z. We know that see (2.5). We also know that there exists ξ ∈ I such that ∆mg[Z] = (LZ[g])(m), (5.16) m(cid:88) 1 m! 1 k! (5.17) We apply Lemma 5.5 to the function F = LZ[g], points {z0, ..., zm} and p = 1. This lemma and (LZ[g])(k)(ξ). ∆kg[S ] = (5.17) tell us that ∆kg[S ] ≤ C(m) (LZ[g])(m)(s) ds + I j−k ∆ j(LZ[g])[z0, ..., z j] m−1(cid:88) I(m−k)−1 I (cid:90) Im−k ∆mg[Z] + j=k m−1(cid:88) j=k I j−k ∆ jg[z0, ..., z j]  This inequality and (5.16) imply that ∆kg[S ] ≤ C(m) proving the lemma. (cid:3) Integrating both sides of inequality (5.13) on I (with respect to x), we obtain the following Lemma 5.7 In the settings of Lemma 5.5, the following inequality F(k)(x)p dx ≤ C(m)p F(m)(s)p ds + I( j−k)p+1 ∆ jF[z0, ..., z j]p  .  (cid:90) I holds. (cid:90) I(m−k)p m−1(cid:88) I j=k 67 Lemma 5.8 Let p ∈ [1,∞) and m, n ∈ N, m < n. Let S = {y0, ..., yn} where {yi}n increasing sequence of points in R. Suppose that there exists (cid:96) ∈ N, m ≤ (cid:96) ≤ n, such that i=0 is a strictly Then for every function g defined on S and every k ∈ {0, ..., m − 1} the following inequality y(cid:96) − y0 ≤ 2 but y(cid:96)+1 − y0 > 2. m(cid:88) (cid:96)− j(cid:88) min{1, yi+m − yi}(cid:12)(cid:12)(cid:12)∆ jg[yi, ..., yi+ j] (cid:12)(cid:12)(cid:12)p ∆kg[y0, ..., yk]p ≤ C(m)p (5.18) holds. (We recall that, according to our convention (1.2), yi = +∞ provided i > n.) j=k i=0 Proof. Let m(cid:88) (cid:96)− j(cid:88) j=k i=0 A = and let V = {y0, ..., y(cid:96)}. Theorem 1.6 tells us that there exists a function G ∈ Lm GV = gV and p (R) such that , (5.19) min{1, yi+m − yi}(cid:12)(cid:12)(cid:12)∆ jg[yi, ..., yi+ j] (cid:12)(cid:12)(cid:12)p (cid:96)−m(cid:88) (yi+m − yi)∆mg[yi, ..., yi+m]p. i=0 (cid:107)G(cid:107)p Lm p (R) ≤ C(m)p (cid:96)−m(cid:88) Note that yi+m − yi ≤ y(cid:96) − y0 ≤ 2 provided 0 ≤ i ≤ (cid:96) − m, so that yi+m − yi ≤ 2 min{1, yi+m − yi}. Hence, (cid:107)G(cid:107)p p (R) ≤ C(m)p Lm min{1, yi+m − yi}∆mg[yi, ..., yi+m]p ≤ C(m)p A . We also note that, thanks to (2.6), for every there exists ¯x ∈ [y0, yk] such that i=0 ∆kg[y0, ..., yk] = (Recall that 0 ≤ k ≤ m − 1.) Let I = [y0, y(cid:96)]. We proceed by cases. The first case: y(cid:96) − y0 ≤ 1. Let 1 k! G(k)( ¯x) . Since ¯x, z0, ..., zm−1 ∈ I, property (5.21) and Lemma 5.5 imply that ∆kg[y0, ..., yk]p ≤ C(m)p G(m)(s)p ds + I( j−k)p ∆ jG[z0, ..., z j]p z j = y(cid:96)−m+1+ j, (cid:90) I(m−k)p−1 j = 0, ..., m − 1. m−1(cid:88) I j=k (5.20) (5.21) (5.22) (5.23)  . Note that I = y(cid:96) − y0 ≤ 1 and (m − k)p − 1 ≥ 0 because 0 ≤ k ≤ m − 1 and p ≥ 1. Therefore, I(m−k)p−1 ≤ 1 and I( j−k)p ≤ 1 for every j = k, ..., m − 1. 68 This observation and inequalities (5.23), (5.22) and (5.20) tell us that ∆kg[y0, ..., yk]p ≤ C(m)p ∆ jg[y(cid:96)−m+1, ..., y(cid:96)−m+1+ j]p Since y(cid:96) − y0 ≤ 1 and y(cid:96)+1 − y0 > 2, we have min{1, y(cid:96)+1 − y(cid:96)−m+1} = 1 so that  . m−1(cid:88) j=k A + m−1(cid:88) A + ≤ C(m)p {A + A} j=k ∆kg[y0, ..., yk]p ≤ C(m)p min{1, y(cid:96)+1 − y(cid:96)−m+1}∆ jg[y(cid:96)−m+1, ..., y(cid:96)−m+1+ j]p proving inequality (5.18) in the case under consideration. The second case: y(cid:96) − y0 > 1. We know that 1 < I = y(cid:96) − y0 ≤ 2. In this case inequality (5.2) and (5.21) tell us that ∆kg[y0, ..., yk] = G(k)( ¯x) ≤ 2p 1 k! dy. (5.24) (cid:16)G(k)(y)p + G(k+1)(y)p(cid:17) (cid:90) I Let Since Jν ⊂ I = [y0, y(cid:96)], we have Jν = [yν, yν+m], ν = 0, ..., (cid:96) − m. Jν ≤ I = y(cid:96) − y0 ≤ 2 for every ν = 0, ..., (cid:96) − m. (5.25) Let us apply Lemma 5.7 to the interval Jν, points zi = yν+i, i = 0, ..., m − 1, and the function (cid:90) F = G. This lemma tells us that (cid:90) m−1(cid:88) G(m)(s)p ds + Jν( j−k)p+1 ∆ jg[yν, ..., yν+ j]p G(k)(y)p dy ≤ C(m)p   ≤ C(m)p G(m)(s)p ds + Jν j=k Jν ∆ jg[yν, ..., yν+ j]p  . Since Jν ≤ 2, see (5.25), the following inequality (cid:90) m−1(cid:88) G(k)(y)p dy ≤ C(m)p G(m)(s)p ds + min{1, yν+m − yν}∆ jg[yν, ..., yν+ j]p Jν Jν(m−k)p  (cid:90)  (cid:90)  (cid:90) Jν Jν m−1(cid:88) j=k j=k m−1(cid:88) j=k+1 69 Jν Jν Jν holds for every ν = 0, ..., (cid:96) − m. (cid:90) In the same way we prove that G(k+1)(y)p dy ≤ C(m)p G(m)(s)p ds + min{1, yν+m − yν}∆ jg[yν, ..., yν+ j]p  (5.26)  (5.27) Finally, we note that I = [y0, y(cid:96)] ⊂ ∪{Jν : ν = 0, ..., (cid:96) − m}. This inclusion, inequalities (5.24), provided k ≤ m − 2 and ν = 0, ..., (cid:96) − m. (cid:90) (5.26) and (5.27) imply that ∆kg[y0, ..., yk]p ≤ C(m)p  (cid:96)−m(cid:88) ν=0 = C(m)p {A1 + A2}. Jν (cid:96)−m(cid:88) m−1(cid:88) ν=0 j=k G(m)(s)p ds + min{1, yν+m − yν}∆ jg[yν, ..., yν+ j]p  Clearly, (cid:96)−m(cid:88) m−1(cid:88) min{1, yν+m − yν}∆ jg[yν, ..., yν+ j]p = A2 = ν=0 j=k m−1(cid:88) (cid:96)−m(cid:88) j=k i=0 min{1, yi+m − yi}∆ jg[yi, ..., yi+ j]p so that A2 ≤ A, see definition (5.19). Since the covering multiplicity of the family {Jν}(cid:96)−m ν=0 is bounded by m + 1, (cid:96)−m(cid:88) (cid:90) (cid:90) A1 = G(m)(s)p ds ≤ (m + 1) G(m)(s)p ds = (m + 1)(cid:107)G(cid:107)p p (R) ≤ C(m)p A. Lm ν=0 Jν R See (5.20). Hence, ∆kg[y0, ..., yk]p ≤ C(m)p {A1 + A2} ≤ C(m)p A proving the lemma in the second case. The proof of the lemma is complete. The next lemma shows that inequality (5.5) of Assumption 5.2 has a certain self-improvement (cid:3) property. Lemma 5.9 Let {xi}n Then for every function g on S and every k ∈ {0, ..., m} the following inequality i=0, n ≥ m, be a finite strictly increasing sequence in R, and let S = {x0, ..., xn}. min{1, xi+m − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤ C(m)p min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ jg[xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p m(cid:88) n− j(cid:88) n−k(cid:88) i=0 j=k i=0 holds. (We recall our Convention 5.3: xi = −∞, if i < 0, and xi = +∞ if i > n). Proof. For k = m the lemma is obvious. Fix k ∈ {0, ..., m − 1} and set αk = and m(cid:88) n− j(cid:88) min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ jg[xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p n−k(cid:88) min{1, xi+m − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p i=0 j=k Ak = (5.28) . i=0 70 Clearly, Ak ≤ αk + Bk where Bk = n−k(cid:88) min{1, xi+k − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p . i=0 Prove that Bk ≤ C(m)p αk. We introduce a partition {I j : j = 1, 2, 3} of the set {0, ..., n − k} as follows: Let I1 = {i ∈ {0, ..., n − k} : xi+k − xi+k−m ≤ 2}, and Let I2 = {i ∈ {0, ..., n − k} : xi+k − xi+k−m > 2 and xi+m − xi > 1} I3 = {i ∈ {0, ..., n − k} : xi+k − xi+k−m > 2 and xi+m − xi ≤ 1}. (cid:12)(cid:12)(cid:12)p min{1, xi+k − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:88) i∈Iν B(ν) k = , ν = 1, 2, 3 (5.29) provided Iν (cid:44) ∅; otherwise, we set B(ν) Clearly, Bk = B(1) k + B(2) k + B(3) k = 0. k . Prove that k ≤ C(m)p αk B(ν) for every ν = 1, 2, 3. Without loss of generality, we may assume that Iν (cid:44) ∅ for each ν = 1, 2, 3. First we estimate B(1) Fix i ∈ I1 and put Convention 5.3) which contradicts the inequality xi+k − xi+k−m ≤ 2. k . We note that i + k − m ≥ 0 for each i ∈ I1; otherwise xi+k−m = −∞ (see Let and Z = {z0, ..., zm}. Note that diam Z = xi+k − xi+k−m ≤ 2 because i ∈ I1. We apply Lemma 5.6 to the set Z and set S = {xi, ..., xi+k}, and obtain that z j = xi+k−m+ j, j = 0, ..., m. = C(m)p min{1, xi+k − xi+k−m} ∆ jg[xi+k−m, ..., xi+k−m+ j]p ≤ C(m)p αk. j=0 i∈I1 71 (diam Z) j−k ∆ jg[z0, ..., z j] p m(cid:88)  m(cid:88) m(cid:88) j=k ∆kg[xi, ..., xi+k]p ≤ C(m)p ≤ C(m)p This enables us to estimate B(1) B(1) k = (cid:88) i∈I1 ≤ C(m)p k as follows: min{1, xi+k − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p m(cid:88) (cid:88) m(cid:88) min{1, xi+k − xi+k−m} (cid:88) i∈I1 j=0 ∆ jg[xi+k−m, ..., xi+k−m+ j]p ∆ jg[z0, ..., z j]p = C(m)p ∆ jg[xi+k−m, ..., xi+k−m+ j]p. j=k j=k See (5.28). Let us estimate B(2) k . Since xi+k − xi+k−m > 2 for every i ∈ I2, we have (cid:88) i∈I2 (cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p . See (5.29). We also know that xi+m − xi > 1 for each i ∈ I2 so that min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p . B(2) k = B(2) k = (cid:88) i∈I2 This equality and definition (5.28) imply that B(2) k ≤ αk. k ≤ C(m)p αk. We recall that It remains to prove that B(3) xi+k − xi+k−m > 2 and xi+m − xi ≤ 1 for every i ∈ I3. (5.30) Let Ti = [xi+k−m, xi+k], i ∈ I3. We recall that xi+k−m = −∞ if i + k − m < 0 according to our convention given in the formulation of the lemma. Thus, Ti = (−∞, xi+k] for each i ∈ I3, i < m − k. Let T = {Ti : i ∈ I3}. Note that T > 2 for each T ∈ T , see (5.30). We also note that i ≤ n−m provided i ∈ I3; in fact, otherwise i+m > n and xi+m = +∞ (according to Convention 5.3) which contradicts to inequality xi+m − xi ≤ 1. In particular, i + k ≤ n for every i ∈ I3 so that xi+k < +∞, i ∈ I3. Thus, T is a finite family of bounded from above intervals. Clearly, given i0 ∈ I3, there are at most 2m + 2 intervals Ti from T such that Ti0 ∩ Ti (cid:44) ∅. This property and Lemma 4.14 imply the existence of subfamilies {T1, ...,Tκ}, κ ≤ 2m + 3, of the family T such that: = ∅ for distinct j1, j2 ∈ {0, ..., κ}; (i) For each j ∈ {1, ..., κ} the intervals of the family T j are pairwise disjoint; (ii) T j1 ∩ T j2 (iii) T = ∪{T j : j = 0, ..., κ}. Fix j ∈ {0, ..., κ} and consider the family T j. This is a finite family of pairwise disjoint closed and bounded from above intervals in R. We know that T > 2 for each T ∈ T j. This enables us to partition T j into two families, say T (1) , with the following properties: and T (2) j j the distance between any two distinct intervals T, T(cid:48) ∈ T (ν) j is at least 2. For instance, we can produce the families T (1) Here ν = 1 or 2. in "increasing order" (recall that these intervals are disjoint) and (ii) setting T (1) j by (i) enumerating the intervals from T j to be the family of and T (2) j j 72 intervals from T j with the odd index, and T (2) index. j to be the family of intervals from T j with the even These observations enables us to make the following assumption: Without loss of generality, we may suppose that the family T = {Ti : property: i ∈ I3} has the following dist(T, T(cid:48)) > 2 for every T, T(cid:48) ∈ T , T (cid:44) T(cid:48) (5.31) We recall that Ti = [xi+k−m, xi+k], so that xi ∈ Ti, i ∈ I3. This and property (5.31) imply that . xi − x j > 2 provided i, j ∈ I3 and i (cid:44) j. Hence, [xi, xi + 2] ∩ [x j, x j + 2] = ∅, i, j ∈ I3, i (cid:44) j. (5.32) We also note that, given i ∈ I3 there exists a positive integer (cid:96)i ∈ [m, n] such that x(cid:96)i − xi ≤ 2 but x(cid:96)i+1 − xi > 2. (5.33) This is immediate from inequality xi+m − xi ≤ 1, i ∈ I3, see (5.30). (In general, it may happen that (cid:96)i = n; in this case, according to Convention 5.3, x(cid:96)i+1 = +∞.) Let us apply Lemma 5.8 to the function g, points yi = xi+ j, j = 0, ..., n− i, and the number (cid:96) = (cid:96)i. This lemma tells us that the following inequality m(cid:88) (cid:96)i− j(cid:88) j=k ν=i min{1, xν+m − xν}(cid:12)(cid:12)(cid:12)∆ jg[xν, ..., xν+ j] (cid:12)(cid:12)(cid:12)p (cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p m(cid:88) (cid:12)(cid:12)(cid:12)p = C(m)p min{1, xi+k − xi+k−m}(cid:12)(cid:12)(cid:12)∆kg[xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p = m(cid:88) (cid:88) min{1, xν+m − xν}(cid:12)(cid:12)(cid:12)∆ jg[xν, ..., xν+ j] (cid:88) (cid:96)i− j(cid:88) i∈I3 j=k i∈I3 ν=i Y j . j=k (cid:88) B(3) k = i∈I3 ≤ C(m)p ∆kg[xi, ..., xi+k]p ≤ C(m)p holds. This inequality and the property xi+k − xi+k−m > 2, i ∈ I3, imply that Note that, given j ∈ {0, ..., m} and i ∈ I3, the points xi, ..., x(cid:96)i ∈ [xi, xi + 2], see (5.33). By this property and by (5.32), for every i(cid:48), i(cid:48)(cid:48) ∈ I3, i(cid:48) (cid:44) i(cid:48)(cid:48), the families of indexes {i(cid:48), ..., (cid:96)i(cid:48)} and {i(cid:48)(cid:48), ..., (cid:96)i(cid:48)(cid:48)} are disjoint. Hence, min{1, xν+m − xν}(cid:12)(cid:12)(cid:12)∆ jg[xν, ..., xν+ j] (cid:12)(cid:12)(cid:12)p ≤ min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ jg[xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p . (cid:88) (cid:96)i− j(cid:88) i∈I3 ν=i Y j = n− j(cid:88) i=0 This inequality and (5.28) imply that m(cid:88) n− j(cid:88) k ≤ C(m)p B(3) min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ jg[xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p = C(m)p αk. j=k i=0 We have proved that B(ν) k ≤ C(m)p αk for each ν = 1, 2, 3. Hence, k ≤ C(m)p αk Bk = B(1) k + B(2) k + B(3) 73 proving that Ak ≤ αk + Bk ≤ C(m)p αk. The proof of the lemma is complete. (cid:3) p (R)-extension of the function f : E → R. We We turn to constructing of an almost optimal Wm recall that Assumption 5.2 holds for the function f . We recall that JE = {Jk = (ak, bk) : k ∈ K}, see (4.5), is the family of pairwise disjoint open intervals (bounded or unbounded) satisfying condition (4.6). We introduce a (perhaps empty) subfamily GE of JE defined by GE = {J ∈ JE : J > 4}. (5.34) Given a bounded interval J = (aJ, bJ) ∈ GE, we put nJ = (cid:98)J/2(cid:99) (5.35) where (cid:98)·(cid:99) denotes the greatest integer function. Let (cid:96)(J) = J/nJ; then 2 ≤ (cid:96)(J) ≤ 3. We associate to the interval J points n = aJ + (cid:96)(J) · n, Y(J) n = 1, ..., nJ − 1. (5.36) We set Y(J) 0 = aJ and Y(J) nJ = bJ and put nJ−1} 1 , ..., Y(J) S J = {Y(J) for every bounded interval J ∈ GE. Thus, the points Y(J) 0 , ..., Y(J) nJ divide J = (aJ, bJ) in nJ subintervals (Y(J) n , Y(J) of equal length (= (cid:96)(J)). We know that n+1 − Y(J) 2 ≤ (cid:96)(J) = Y(J) n ≤ 3 for every n = 0, ..., nJ − 1. Let J = (aJ, bJ) be an unbounded interval. (Clearly, J ∈ GE). In this case we set provided J = (aJ, +∞), and n = aJ + 2 · n, Y(J) n ∈ N, n = bJ − 2 · n, Y(J) n ∈ N, (5.37) n+1), n = 0, ..., nJ − 1, (5.38) (5.39) (5.40) (5.41) (5.42) (5.43) provided J = (−∞, bJ). In other words, we divide J in subintervals of length 2. Finally, we set S J = {Y(J) n : n ∈ N} Let for every unbounded interval J ∈ GE. J∈GE whenever GE (cid:44) ∅, and G = ∅ otherwise. Clearly, (cid:91) G = S J dist(E, G) ≥ 2. (Recall that dist(A,∅) = +∞ provided A (cid:44) ∅, so that (5.43) includes the case of G = ∅ as well). 74 (cid:101)E = E ∪ G. Let Note an important property of the set (cid:101)E which is immediate from our construction of the families S J, J ∈ GE: By f : (cid:101)E → R we denote the extension of f from E to(cid:101)E by zero; thus, dist(x,(cid:101)E) ≤ 2 for every x ∈ R. (5.45) (5.44) (cid:40) f (x), 0, f (x) = if x ∈ E, if x ∈ G. Let us show that ≤ C(m) λ. Our proof of this statement relies on a series of auxiliary lemmas. p (R)(cid:101)E and (cid:107) f(cid:107)Wm f ∈ Wm p (R)(cid:101)E Lemma 5.10 Let k ∈ {0, ..., m − 1} and let {yi}k i=0 be a strictly increasing sequence in E. Then ∆k f [y0, ..., yk] ≤ C(m) λ . Proof. We know that E contains at least m + 1 distinct points. Therefore, there exists a strictly ν=0 in E containing the points y0, ..., yk. Thus, the set Y = {y0, ..., yk} is a Property ((cid:70)7) (i), Section 2.1 (with n = m) tells us that there exist non-negative numbers αν, increasing sequence {xν}m (k + 1)-point subset of the set X = {x0, ..., xm} ⊂ E such that x0 < ... < xm. ν = 0, ..., m − k, such that α0 + ... + αm−k = 1, and (5.46) (5.47) ∆k f [Y] = αν ∆k f [xν, ..., xν+k] . Hence, ν=0 ∆k f [Y]p ≤ (m + 1)p ∆k f [xν, ..., xν+k]p. m−k(cid:88) m−k(cid:88) ν=0 m−k(cid:88) m(cid:88) m− j(cid:88) ν=0 j=k i=0 We note that for every ν ∈ {0, ..., m − k} either ν + m > m or v + k − m < 0 (because 0 ≤ k < m). Therefore, according to Convention 5.3, either xν+m = +∞ or xν+k−m = −∞ proving that min{1, xν+m − xν+k−m} = 1 for all ν = 0, ..., m − k. Thus, ∆k f [Y]p ≤ (m + 1)p min{1, xν+m − xν+k−m}∆k f [xν, ..., xν+k]p. This inequality and Lemma 5.9 (with n = m) imply that ∆k f [Y]p ≤ C(m)p min{1, xi+m − xi}∆ j f [xi, ..., xi+ j]p. Finally, applying Assumption 5.2 to the right hand side of this inequality, we get the required inequality (5.47) proving the lemma. (cid:3) 75 Lemma 5.11 For every finite strictly increasing sequence of points {yi}n ing inequality (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (cid:12)(cid:12)(cid:12)p ≤ C(m)p λp (yi+m − yi) n−m(cid:88) i=0 ⊂ (cid:101)E, n ≥ m, the follow- holds. Proof. Let and let i=0 n−m(cid:88) i=0 A = (cid:12)(cid:12)(cid:12)p (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (yi+m − yi) I1 = {i ∈ {0, ..., n − m} : yi+m − yi < 2} and Given j ∈ {1, 2}, let A j = (yi+m − yi) I2 = {i ∈ {0, ..., n − m} : yi+m − yi ≥ 2}. (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (cid:12)(cid:12)(cid:12)p (5.48) (cid:88) i∈I j provided I j (cid:44) ∅, and let A j = 0 otherwise. Clearly, A = A1 + A2. Prove that A1 ≤ 2 λp. This inequality is trivial if I1 = ∅. Let us assume that I1 (cid:44) ∅. We introduce two (perhaps empty) subfamilies of I1 by letting I1,E = {i ∈ I1 : yi ∈ E} and I1,G = {i ∈ I1 : yi ∈ G}. We recall that yi+m − yi < 2 for every i ∈ I1, and that dist(E, G) ≥ 2 (see (5.43)). These inequalities imply that and yi, ..., yi+m ∈ E for every i ∈ I1,E, yi, ..., yi+m ∈ G for every i ∈ I1,G. Recall that fG ≡ 0 and fE = f , so that A1 = 0 if I1,E = ∅. Suppose that I1,E (cid:44) ∅. In this case (5.49), (5.50) and the property fG ≡ 0 imply that (yi+m − yi)∆m f [yi, ..., yi+m]p . (yi+m − yi) (cid:88) (cid:88) A1 = i∈I1 Hence, A1 ≤ 2 (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (cid:12)(cid:12)(cid:12)p = (cid:88) i∈I1,E i∈I1,E min{1, yi+m − yi} ∆m f [yi, ..., yi+m]p because yi+m − yi < 2 for each i ∈ I1,E ⊂ I1. (cid:91) Let S = i∈I1,E {yi, ..., yi+m}. 76 (5.49) (5.50) (5.51) (5.52) i=0 is a strictly increasing sequence, one i=0. In other words, S can be represented We know that S ⊂ E, see (5.49). Furthermore, since {yi}n can consider S as a strictly increasing subsequence of {yi}n in the form S = {yiν : ν = 0, ..., (cid:96)} i=0, we have m ≤ (cid:96) ≤ n. where (cid:96) = #S and {iν}(cid:96) ν=0 is a strictly increasing subsequence of non-negative integers. Since I1,E (cid:44) ∅ and S is a subsequence of {yi}n Let xν = yiν, ν = 0, ..., (cid:96). Definition (5.52) tells us that xν+ j = yiν+ j for every ν ∈ {0, ..., (cid:96)} such that iν ∈ I1,E. This property and (5.51) imply that A1 ≤ 2 min{1, yiν+m − yiν}(cid:12)(cid:12)(cid:12)∆m f [yiν, ..., yiν+m] min{1, xν+m − xν}∆m f [xν, ..., xν+m]p . (cid:12)(cid:12)(cid:12)p = 2 (cid:88) ν: iν∈I1,E (cid:88) ν: iν∈I1,E Hence, (cid:96)(cid:88) A1 ≤ 2 min{1, xν+m − xν} ∆m f [xν, ..., xν+m]p v=0 so that, thanks to Assumption 5.2, A1 ≤ 2 λp. (5.53) Prove that A2 ≤ C(m)p λp. We may assume that the set I2 determined in (5.48) is not empty; Let i ∈ I2 and let Ti = [yi−m, yi+2m]. (We recall our Convention 5.3 concerning the values of y j otherwise, A2 = 0. whenever i (cid:60) {0, ..., n}.) Let T = {Ti : i ∈ I2}. Note that for each interval Ti0 ∈ T , there exist at most 6m + 1 intervals Ti ∈ T such that Ti ∩ Ti0 (cid:44) ∅. Lemma 4.14 tells us that we can partition T in at most κ ≤ 6m + 2 subfamilies {T1, ...,Tκ} each consisting of pairwise disjoint intervals. Using the same argument as at the end of the proof of Lemma 4.15, without loss of generality, we may assume that the family T itself consists of pairwise disjoint intervals, (5.54) i.e., Ti ∩ T j (cid:44) ∅ for every i, j ∈ I2, i (cid:44) j. In particular, this property implies that i − j > 3m for every i, j ∈ I2, i (cid:44) j. Fix i ∈ I2 and apply Lemma 5.4 to points {yi, ..., yi+m} and the function f . (Recall that yi+m−yi > 2 for all i ∈ I2.) By this lemma, there exist ki ∈ {0, ..., m − 1} and αi ∈ {0, ..., m − ki} such that yαi+ki − yαi ≤ 1 and Furthermore, ∆m f [yi, ..., yi+m] ≤ 2m ∆ki f [yαi, ..., yαi+ki]/(yi+m − yi). (5.55) either αi + ki + 1 ≤ i + m and yαi+ki+1 − yαi ≥ 1, or αi ≥ i + 1 and yαi+ki − yαi−1 ≥ 1. (5.56) 77 Since yi+m − yi > 2 and p > 1, inequality (5.55) implies that (yi+m − yi)∆m f [yi, ..., yi+m]p ≤ 2mp ∆ki f [yαi, ..., yαi+ki]p/(yi+m − yi)p−1 ≤ 2mp ∆ki f [yαi, ..., yαi+ki]p. Moreover, since yαi+ki − yαi ≤ 1 and dist(E, G) ≥ 2 (see (5.43)), either the set {yαi, ..., yαi+ki} ⊂ E or {yαi, ..., yαi+ki} ⊂ G. Since f ≡ 0 on G, we have (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (cid:12)(cid:12)(cid:12)p ≤ 2mp (yi+m − yi) ∆ki f [yαi, ..., yαi+ki]p (cid:88) i∈I2,E (cid:88) i∈I2 A2 = where (5.57) (5.58) (5.59) (5.60) I2,E = {i ∈ I2 : yαi, ..., yαi+ki ∈ E}. We recall that ki ∈ {0, ..., m − 1}, αi ∈ {0, ..., m − ki} for each i ∈ I2,E so that yαi+m − yαi+ki−m ≥ max{yαi+ki+1 − yαi, yαi+ki − yαi−1}. This inequality and property (5.56) imply that yαi+m − yαi+ki−m ≥ 1 for every i ∈ I2,E. Let H = {yαi, ..., yαi+ki} (cid:91) i∈ I2,E We know that H ⊂ E, see (5.58). In turn, assumption (5.54) tells us that Ti ∩ T j = ∅ for every and let κ = #H. i, j ∈ I2,E, i (cid:44) j, which implies the following property: {yαi, ..., yαi+ki}(cid:84){yα j, ..., yα j+k j} = ∅ for all i, j ∈ I2,E, i (cid:44) j. In particular, {αi}i∈I2,E is a strictly increasing sequence, and #I2,E ≤ #H = κ. (5.61) i=0 is a strictly increasing sequence, and H ⊂ E, we can consider the set H as a strictly i=0 whose elements lie in E. This enables us to represent Consider two cases. The first case: κ ≥ m + 1. Since {yi}n increasing subsequence of the sequence {yi}n H in the form H = {x0, ..., xκ} where x0 < ... < xκ. Furthermore, we know that for each i ∈ I2,E there exists a unique νi ∈ {0, ..., κ} such that yαi = xνi. The sequence {νi}i∈I2,E is a strictly increasing sequence such that αi + j = νi + j for every j = 0, ..., ki. See (5.58). Moreover, Since {xν}κ ν=0 is a strictly increasing subsequence of {yi}n i=0, we have xνi+ j ≥ yαi+ j and xνi− j ≤ yαi− j for every j = 0, ..., m. (5.62) Hence, ∆ki f [yαi, ..., yαi+ki] = ∆ki f [xνi, ..., xνi+ki] for all i ∈ I2,E. 78 Inequalities (5.59) and (5.62) imply that xνi+m − xνi+ki−m ≥ yαi+m − yαi+ki−m ≥ 1 so that ∆ki f [yαi, ..., yαi+ki]p = min{1, xνi+m − xνi+ki−m}∆ki f [xνi, ..., xνi+ki]p, i ∈ I2,E. This equality and (5.57) imply the following estimate of A2: A2 ≤ 2mp min{1, xνi+m − xνi+ki−m}∆ki f [xνi, ..., xνi+ki]p. (cid:88) n−k(cid:88) m(cid:88) i∈I2,E k=0 ν=0 Hence, A2 ≤ 2mp min{1, xν+m − xν+k−m}∆k f [xν, ..., xν+k]p (5.63) where n = κ − 1. We know that in the case under consideration n ≥ m. We apply Lemma 5.9 which tells us that for every k ∈ {0, ..., m} the following inequality min{1, xν+m − xν+k−m}(cid:12)(cid:12)(cid:12)∆k f [xν, ..., xν+k] (cid:12)(cid:12)(cid:12)p ≤ C(m)p n−k(cid:88) ν=0 m(cid:88) n− j(cid:88) j=k i=0 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ j f [xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p (cid:12)(cid:12)(cid:12)p ≤ λp (5.64) holds. On the other hand, by Assumption 5.2, m(cid:88) n−k(cid:88) j=k i=0 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ j f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤ C(m)p λp min{1, xν+m − xν+k−m}(cid:12)(cid:12)(cid:12)∆k f [xν, ..., xν+k] proving that n−k(cid:88) for every k ∈ {0, ..., m}. ν=0 This inequality and (5.63) imply the required inequality A2 ≤ 2mp(m + 1) C(m)p λp. The second case: κ ≤ m. Let us see that in this case A2 ≤ C(m)p λp as well. We prove this by showing that each item in the right hand side of inequality (5.57) is bounded by C(m)p λp. Let us fix i ∈ I2,E, and consider the sequence {yαi, ..., yαi+ki} ⊂ E. Definition (5.60) tells us that 0 < ki ≤ #H − 1 = κ − 1 ≤ m − 1. We apply Lemma 5.10 to the sequence {yαi, ..., yαi+ki} ⊂ E and obtain the following inequality: ∆ki f [yαi, ..., yαi+ki] ≤ C(m) λ. 79 (5.65) This and (5.57) imply that A2 ≤ 2mp C(m)p λp · #I2,E. second case. But #I2,E ≤ #H = κ ≤ m, see (5.61), proving the required inequality A2 ≤ C(m)p λp in the Thus, we have proved that in the both cases A2 ≤ C(m)p λp. Finally, A = A1 + A2 ≤ 2λ + C(m)p λ = (2 + C(m)p) λp, and the proof of the lemma is complete. (cid:3) Combining Lemma 5.11 with Theorem 1.7, we conclude that there exists a function F ∈ Lm p (R) with (5.66) = f . Since fE = f , the function F is an extension of f from E to all of R, i.e., (cid:107)F(cid:107)Lm p (R) ≤ C(m) λ such that F(cid:101)E FE = f . We denote the extension F by F = ExtE( f : Wm p (R)). (5.67) Thus, ExtE( f : Wm p (R)) = Ext(cid:101)E ( f : Lm p (R)), (5.68) see definition (4.94). Let ε = 3 and let A be a maximal ε-net in (cid:101)E; thus A is a subset of (cid:101)E having the following ((cid:70)2) dist(y, A) < 3 for every y ∈ (cid:101)E. ((cid:70)1) a − a(cid:48) ≥ 3 for every a, a(cid:48) ∈ A, a (cid:44) a(cid:48); properties: Now, properties (5.45) and ((cid:70)2) imply that (5.69) This inequality and property ((cid:70)1) enable us to represent the set A as a certain bi-infinite strongly dist(x, A) ≤ 5 for every x ∈ R. increasing sequence {ai}+∞ A = {ai}+∞ i=−∞ in(cid:101)E. Thus, i=−∞ where ai ∈ (cid:101)E and ai < ai+1 for all i ∈ Z. (5.70) (5.71) Furthermore, thanks to ((cid:70)1) and(5.69), ai → −∞ as i → −∞, and ai → +∞ as i → +∞, and (5.72) Of course, the first inequality in (5.72) is immediate from ((cid:70)1). Prove the second inequality. 3 ≤ ai+1 − ai ≤ 10 for every i ∈ Z. Let bi = (ai + ai+1)/2. Since ai, ai+1 are two consecutive points of the sequence A = {ai}+∞ and bi ∈ [ai, ai+1], we have Hence, by (5.69), dist(bi,{ai, ai+1}) ≤ 5, so that (ai+1 − ai)/2 ≤ 5 proving (5.72). dist(bi, A) = dist(bi,{ai, ai+1}). i=−∞ ⊂ (cid:101)E, 80 Lemma 5.12 For every function g ∈ Lm p (R) such that gG ≡ 0 (see (5.42)) the following inequality (cid:107)g(cid:107)p Lp(R) ≤ C(m)p holds. Proof. Fix N > m and prove that  g(ai)p i=+∞(cid:88) i=−∞ + Lm p (R) (cid:107)g(cid:107)p (cid:107)g(cid:107)p i=+∞(cid:88) i=−∞ +  . g(ai)p aN(cid:90) a−N IntN = g(t)p dt ≤ C(m)p Lm p (R) (5.73) (5.74) In the proof of this inequality we use some ideas of the work [22, p. 451]. Let i ∈ Z, −N ≤ i ≤ N − m + 1, and let Ti = [ai, ai+m−1]. We apply Lemma 5.7 to the interval Ti and the function g taking k = 0 and z j = ai+ j, j = 0, ..., m − 1. This lemma tells us that (cid:90) Ti Timp (cid:90) Ti g(t)p dt ≤ C(m)p g(m)(s)p ds + Ti jp+1 ∆ jg[ai, ..., ai+ j]p We note that aν − aµ ≥ 3 for every ν, µ ∈ {i, ..., i + m − 1}, ν (cid:44) µ, see (5.72), so that, by (2.4),  . Since ai+1 − ai ≤ 10 for every i ∈ Z (see again (5.72)), we have Ti = ai+m−1 − ai ≤ 10(m − 1). Hence,(cid:90) This inequality implies the following estimate of IntN (see (5.74)): Ti g(t)p dt ≤ C(m)p (cid:90) N−m+1(cid:88) IntN ≤ i=−N Ti ∆ jg[ai, ..., ai+ j] ≤ g(ai+ j).  (cid:90) Ti g(m)(s)p ds + g(t)p dt ≤ C(m)p j=0 g(ai+ j)p (cid:90) m−1(cid:88) N−m+1(cid:88)  ≤ C(m)p i=−N Ti g(ai)p (cid:90)  N(cid:88)  aN(cid:90) a−N  for every − N ≤ i ≤ N − m + 1.  . N−m+1(cid:88) m−1(cid:88) g(ai+ j)p i=−N j=0 g(m)(s)p ds + m−1(cid:88) j=0 m−1(cid:88) j=0 Note that given i0 ∈ Z, −N ≤ i0 ≤ N, there exist at most 2(m − 1) intervals from the family of intervals {Ti : i = −N, ..., N − m + 1} which have common points with the interval Ti0. This observation implies the following estimate of IntN: IntN ≤ C(m)p g(m)(s)p ds + g(m)(s)p ds + g(ai)p But, thanks to (5.71), IntN → (cid:107)g(cid:107)p The proof of the lemma is complete. i=−N Lp(R) as N → ∞ proving inequality (5.73). R +∞(cid:88) i=−∞  . (cid:3) 81 Lemma 5.13 The following inequality (cid:107)F(cid:107)Lp(R) ≤ C(m) λ holds. Proof. Let {ai}+∞ i=−∞ be the strictly increasing sequence defined in (5.70), and let Thus, F(ai) = f (ai) for each i ∈ L, and F(ai) = 0 for each i ∈ Z \ L. We apply Lemma 5.12 (taking g = F) and get F(cid:101)E = f , L = {i ∈ Z : ai ∈ E}. f(cid:101)E\E ≡ 0 and (cid:26)(cid:107)F(cid:107)p (cid:107)F(cid:107)p Lp(R) ≤ C(m)p + B Lm p (R) fE = f. (cid:27) We know that where (cid:88) i∈L B = (5.75) (5.76) (5.77) f (ai)p provided L (cid:44) ∅, and B = 0 otherwise. (5.78) Inequalities (5.66) and (5.77) imply that (cid:107)F(cid:107)p Lp(R) ≤ C(m)p {λp + B}. (5.79) Prove that (5.80) Indeed, this is trivial whenever L = ∅ (because B = 0.) Assume that L (cid:44) ∅. Without loss of B ≤ C(m)p λp. generality, we may also assume that L is finite. Lemma 5.10 (with k = 0) tells us that f (ai) ≤ C(m) λ for every i ∈ L, so that (5.80) holds provided #L ≤ m. where n = #L ≥ m + 1. Thus Suppose that #L ≥ m+1. For simplicity of notation, we may assume that in this case L = {0, ..., n} where {ai}n (5.72). Hence, i=0 is a strictly increasing sequence in E such that ai+1 − ai ≥ 3 for all i = 0, ..., n− 1. See B = min{1, ai+m − ai}∆0 f [ai]p. (We recall our Convention 5.3 which for the case of the sequence {ai}n provided i > n.) Therefore, by Assumption 5.2, B ≤ λp. i=0 states that ai = +∞ n(cid:88) i=0 B = f (ai)p n(cid:88) i=0 82 Thus, (5.80) holds. This inequality together with (5.79) imply (5.75) proving the lemma. (cid:3) Proof of the sufficiency part of Theorem 1.2. It is well known that (cid:107)F(cid:107)Wm m(cid:88) p (R) = (cid:107)F(k)(cid:107)Lp(R) ≤ C(m) ((cid:107)F(cid:107)Lp(R) + (cid:107)F(cid:107)Lm p (R)). (5.81) We have proved that every function f on E can be extended to a function F : R → R such that k=0 (cid:107)F(cid:107)Lm p (R) + (cid:107)F(cid:107)Lp(R) ≤ C(m) λ provided Assumption 5.2 holds. See (5.66) and Lemma 5.13. Hence we conclude that F ∈ Wm and (cid:107)F(cid:107)Wm p (R) ≤ C(m) λ. Since FE = f , we have f ∈ Wm p (R)E. Furthermore, by definitions (1.1) and (5.4), p (R) (cid:107) f(cid:107)Wm p (R)E ≤ (cid:107)F(cid:107)Wm p (R) ≤ C(m) λ = C(m)Wm,p( f : E) proving the sufficiency part of Theorem 1.2. The proof of Theorem 1.2 is complete. (cid:3) (cid:3) p (R)-functions on sequences of points. 5.3. Wm In this section we prove Theorem 1.3. Let p ∈ (1,∞) and let E = {xi}(cid:96)2 be a strictly increasing sequence of points in R where (cid:96)1, (cid:96)2 ∈ Z∪{±∞}, (cid:96)1 ≤ (cid:96)2. In this case we assume that the following version of convention (1.2) holds. Convention 5.14 xi = +∞ whenever i > (cid:96)2. quantity(cid:102)Wm,p( f : E) whenever E is a finite subset of R consisting of at most m points. In particular, this convention leads us to a certain simplification of the formula (1.6) for the Indeed, let 0 ≤ n < m and let E = {x0, ..., xn} where x0 < ... < xn. In this case i=(cid:96)1 mE = min{m, #E − 1} = n (see (1.5)), so that (cid:102)Wm,p( f : E) =  n(cid:88) k=0 n−k(cid:88) i=0 Clearly, i + m > n for each 0 ≤ i ≤ n so that, according to Convention 5.14, xi+m = +∞. Therefore min{1, xi+m − xi} = 1 for every i = 0, ..., n, proving that  1 p . min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p n−k(cid:88)  n(cid:88) 1/p (cid:12)(cid:12)(cid:12)p (cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] . k=0 i=0 (cid:102)Wm,p( f : E) = Hence, (cid:102)Wm,p( f : E) ∼ max{∆k f [xi, ..., xi+k] : k = 0, ..., n, i = 0, ..., n − k} (5.82) 83 with constants in this equivalence depending only m. We turn to the proof of the necessity part of Theorem 1.3. (Necessity.) First, consider the case of a set E ⊂ R with #E ≤ m. Let 0 ≤ n < m, and let E = {x0, ..., xn}, x0 < ... < xn. Let f be a function on E, and let F ∈ Wm p (R) be a function on R such that FE = f . Fix k ∈ {0, ..., n} and i ∈ {0, ..., n− k}. Let us apply Lemma 5.1 to the function G = F, I = R, q = p and (cid:90) S = {xi, ..., xi+k}. This lemma tells us that (cid:16)F(k)(y)p + F(k+1)(y)p(cid:17) ∆k f [xi, ..., xi+k]p ≤ 2p dy ≤ 2p+1 (cid:107)F(cid:107)p Wm p (R) proving that R ∆k f [xi, ..., xi+k] ≤ 4(cid:107)F(cid:107)Wm These inequalities and equivalence (5.82) imply that (cid:102)Wm,p( f : E) ≤ C(m)(cid:107)F(cid:107)Wm for all k ∈ {0, ..., n} and i ∈ {0, ..., n − k}. p (R) p (R). Taking the p (R) such that FE = f , we obtain the required infimum in this inequality over all function F ∈ Wm inequality (cid:102)Wm,p( f : E) ≤ C(m)(cid:107) f(cid:107)Wm p (R)E . This proves the necessity in the case under consideration. We turn to the case of a sequence E containing at least m + 1 elements. In this case mE = min{m, #E − 1} = m (see (1.5)). The necessity part of Theorem 1.2 tells us that for every function f ∈ Wm p (R)E the following (5.83) (5.84) inequality Wm,p( f : E) ≤ C(m)(cid:107) f(cid:107)Wm p (R)E Comparing (1.3) with (1.6) we conclude that (cid:102)Wm,p( f : E) ≤ Wm,p( f : E). This inequality and holds. We recall that Wm,p( f : E) is the quantity defined by (1.3). inequality (5.84) imply (5.83) completing the proof of the necessity part of Theorem 1.3. (cid:3) is a strictly increasing sequence in R. Since E contains at least m + 1 (Sufficiency.) We will consider the following two main cases. Case 1. #E > m. We recall that E = {xi}(cid:96)2 element, (cid:96)1 + m ≤ (cid:96)2. Furthermore, in this case mE = min{m, #E − 1} = m. Let f be a function on E satisfying the hypothesis of Theorem 1.3, i.e., i=(cid:96)1 This inequality and definition (1.6) enable us to make the following Assumption 5.15 The following inequality (cid:101)λ = (cid:102)Wm,p( f : E) < ∞. min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤(cid:101)λp m(cid:88) (cid:96)2−k(cid:88) holds. k=0 i=(cid:96)1 84 Our aim is to show that f ∈ Wm p (R)E p (R)E ≤ C(m)(cid:101)λ. and (cid:107) f(cid:107)Wm We prove these properties of f using a slight modification of the proof of the sufficiency part of Theorem 1.2 given in Section 5.2. More specifically, we only implement minor changes into Lemma 5.10 and Lemma 5.11 related to replacing in their proofs the constant λ with the constant(cid:101)λ, Assumption 5.2 with Assumption 5.15, and using tuples of consecutive elements of the sequence E = {xi}(cid:96)2 (rather then arbitrary finite subsequences of E which we used in Lemmas 5.10 and 5.11.) i=(cid:96)1 We begin with an analogue of Lemma 5.10. Lemma 5.16 Let k ∈ {0, ..., m−1}, ν ∈ Z, (cid:96)1 ≤ ν ≤ (cid:96)2−k, and let {yi}n Then i=0 ∆k f [y0, ..., yk] ≤ C(m)(cid:101)λ . = {xν+i}n i=0 where n = (cid:96)2−ν. (5.85) Proof. Let S = {y0, ..., yn}. If ym − y0 ≥ 2, then A = ∆k f [y0, ..., yk]p = min{1, ym − y0}∆k f [y0, ..., yk]p = min{1, xν+m − xν}∆k f [xν, ..., xν+k]p so that, thanks to Assumption 5.15, A ≤(cid:101)λp. Now let ym − y0 < 2, and let (cid:96) be a positive integer, m ≤ (cid:96) ≤ n, such that y(cid:96) − y0 ≤ 2 but y(cid:96)+1 − y0 > 2. (We recall that, according to Convention 5.14, xi = +∞ whenever i > (cid:96)2, so that yi = +∞ if i > n.) Now, Lemma 5.8 tells us that m(cid:88) m(cid:88) j=0 (cid:96)− j(cid:88) (cid:96)− j(cid:88) i=0 min{1, yi+m − yi}(cid:12)(cid:12)(cid:12)∆ j f [yi, ..., yi+ j] (cid:12)(cid:12)(cid:12)p min{1, xν+i+m − xν+i}(cid:12)(cid:12)(cid:12)∆ j f [xν+i, ..., xν+i+m] (cid:12)(cid:12)(cid:12)p . A ≤ C(m)p = C(m)p This inequality and Assumption 5.15 imply (5.85) proving the lemma. (cid:3) j=0 i=0 i=(cid:96)1 is a strictly increasing sequence of points, the set(cid:101)E = E ∪ G (see (5.44)) can be We turn to the analogue of Lemma 5.11. We recall that the set G defined by (5.42) consists of isolated points of R. (Moreover, the distance between any two distinct points of G is at least 1.) Since E = {xi}(cid:96)2 represented as a certain bi-infinite strictly increasing sequence of points: We also recall that the function f : (cid:101)E → R is defined by formula (5.46). Lemma 5.17 Let (cid:96) ∈ Z, n ∈ N, n ≥ m, and let yi = ti+(cid:96), i = 0, ..., n. Then the following inequality (cid:12)(cid:12)(cid:12)p ≤ C(m)p(cid:101)λp (cid:12)(cid:12)(cid:12)∆m f [yi, ..., yi+m] (cid:101)E = {ti}+∞ (yi+m − yi) n−m(cid:88) (5.87) (5.86) i=−∞. holds. i=0 85 i=(cid:96)1 holds for f . Proof. Repeating the proof of Lemma 5.11 for the sequence E = {xi}(cid:96)2 we show that this lemma holds under a weaker hypothesis for the function f on E. More specifically, we assume the constant λ can be replaced with the constant(cid:101)λ provided E is a sequence and Assumption 5.15 that f satisfies the condition of Assumption 5.15 rather than Assumption 5.2 (as for the case of an arbitrary closed set E ⊂ R). In other words, we prove that everywhere in the proof of Lemma 5.11 0 ≤ i ≤ n − k, and let ti, ..., ti+k ∈ (cid:101)E be k + 1 consecutive elements of the sequence(cid:101)E, see (5.86). If The validity of such a replacement relies on the following simple observation: Let 0 ≤ k ≤ m, ti, ..., ti+k belong to the sequence E = {xi}(cid:96)2 , then ti, ..., ti+k are k + 1 consecutive elements of this In particular, this observation enables us to replace the constant λ with(cid:101)λ in inequalities (5.53), sequence. In other words, there exists ν ∈ Z, (cid:96)1 ≤ ν ≤ (cid:96)2 − k, such that ti = xν+i for all i = 0, ..., k. (5.64) - (5.65), and all inequalities after (5.65) until the end of the proof of Lemma 5.11. nitions (5.52) and (5.60): since the set E = {xi}(cid:96)2 i=(cid:96)1 we can put in S = E in (5.52) and H = E in (5.60). Finally, we note the following useful simplification of the proof of Lemma 5.11 related to defi- itself is a strictly increasing sequence of points, i=(cid:96)1 After all these modifications and changes, we literally follow the proof of Lemma 5.11. This leads us to the required inequality (5.87) completing the proof of the lemma. (cid:3) Since the integer (cid:96) from the hypothesis of Lemma 5.17 is arbitrary, and the right hand side of inequality (5.87) does not depend on (cid:96), the following inequality +∞(cid:88) (cid:12)(cid:12)(cid:12)∆m f [ti, ..., ti+m] (cid:12)(cid:12)(cid:12)p ≤ C(m)p(cid:101)λp (ti+m − ti) i=−∞ holds. This inequality tells us that the function f : (cid:101)E → R satisfies the hypothesis of Theorem 1.6. By this theorem, there exists a function(cid:101)F ∈ Lm p (R) ≤ C(m)(cid:101)λ (cid:107)(cid:101)F(cid:107)Lm such that (cid:101)F(cid:101)E (cid:101)FE = f . = f . Since fE = f , the function (cid:101)F is an extension of f from E to all of R, i.e., p (R) with (5.88) The following lemma is an analogue of Lemma 5.13 for the case of sequences. Lemma 5.18 The following inequality (cid:107)(cid:101)F(cid:107)Lp(R) ≤ C(m)(cid:101)λ (5.89) holds. i=−∞ ⊂ (cid:101)E be the bi-infinite strongly increasing sequence determined by (5.70), and let B Proof. The proof relies on a slight modification of the proof of Lemma 5.13. Let {ai}+∞ We follow the proof of Lemma 5.13 and obtain an analogue of inequality (5.79) which states be the quantity defined by formula (5.78). We also recall the definition of the family of indexes L = {i ∈ Z : ai ∈ E} given in (5.76). (cid:107)(cid:101)F(cid:107)p Lp(R) ≤ C(m)p {(cid:101)λp + B}. (5.90) that 86 Thus, our task is to prove that (cid:88) i∈L B = f (ai)p ≤ C(m)p(cid:101)λp. (5.91) As in Lemma 5.13, we may assume that L (cid:44) ∅ (otherwise B = 0). Let κ = #L − 1. (Thus 0 ≤ κ ≤ +∞.) i=−∞ is strictly increasing, the family of points {ai : i ∈ L} is a subsequence of this Since {ai}+∞ sequence lying in the strictly increasing sequence E = {xi}(cid:96)2 . This enables us to consider the family {ai : i ∈ L} as a strictly increasing subsequence of {xi}(cid:96)2 , i.e., i=(cid:96)1 i=(cid:96)1 ai = xiν where ν = 0, ..., κ, and iν1 < iν2 for all 0 ≤ ν1 < ν2 ≤ κ. Note that, according to Convention 5.14, xiν+m = +∞ provided 0 ≤ ν ≤ κ and iν + m > (cid:96)2. In particular, in this case min{1, xiν+m − xiν} = 1. Let us partition the family L into the following two subfamilies: L1 = {i ∈ L : ai = xiν, xiν+m − xiν ≥ 2}, and Clearly, for every i ∈ L1, L2 = {i ∈ L : ai = xiν, xiν+m − xiν < 2}. f (ai)p = f (xiν)p = min{1, xiν+m − xiν} f (xiν)p. (cid:88) i∈L1 Hence, so that B1 = (5.92) (5.93) f (ai)p = B1 ≤ (cid:96)2(cid:88) i=iν∈L1 (cid:88) (cid:9)(cid:12)(cid:12)(cid:12)∆0 f [xiν] (cid:12)(cid:12)(cid:12)p min(cid:8)1, xiν+m − xiν min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆0 f [xi] (cid:12)(cid:12)(cid:12)p (cid:88) . f (ai)p. B2 = i∈L2 This inequality and Assumption 5.15 imply that B1 ≤(cid:101)λp. i=(cid:96)1 Let us estimate the quantity Lemma 5.16 tells us that f (ai) = ∆0 f [ai] ≤ C(m)(cid:101)λ for every i ∈ L2. B2 ≤ C(m)p (#L2)(cid:101)λp ≤ C(m)p (#L)(cid:101)λp, Hence, so that B2 ≤ C(m)p(cid:101)λp provided #L = κ + 1 ≤ m + 1. 87 Prove that B2 satisfies the same inequality whenever #L = κ + 1 > m + 1, i.e., κ > m. In particular, in this case #E = (cid:96)2 − (cid:96)1 + 1 ≥ #L > m + 1 so that (cid:96)2 − (cid:96)1 > m. Fix i ∈ L2. Thus ai = xiν for some ν ∈ {0, ..., κ}, and xiν+m − xiν < 2. We know that xiν+1 − xiν = aν+1 − aν > 2. Therefore, there exists a positive integer (cid:96)ν, iν + m ≤ (cid:96)ν < iν+1, such that x(cid:96)ν − xiν ≤ 2 but x(cid:96)ν+1 − xiν > 2. Let n = (cid:96)2−iν, and let ys = xiν+s, s = 0, ..., n. We note that n > m because iν ≥ (cid:96)1 and (cid:96)2− (cid:96)1 > m. Let us apply Lemma 5.8 to the strictly increasing sequence {ys}n s=0 and a function g(ys) = f (xiν+s) defined on the set S = {y0, ..., yn} with parameters k = 0 and (cid:96) = (cid:96)ν. This lemma tells us that j=0 j=0 s=0 s=0 m(cid:88) (cid:96)− j(cid:88) (cid:12)(cid:12)(cid:12)p min{1, ys+m − ys}(cid:12)(cid:12)(cid:12)∆ jg[ys, ..., ys+ j] m(cid:88) (cid:96)ν− j(cid:88) (cid:9)(cid:12)(cid:12)(cid:12)∆ j f [xiν+s, ..., xiν+s+ j] (cid:12)(cid:12)(cid:12)p min(cid:8)1, xiν+s+m − xiν+s iν+1−1− j(cid:88) m(cid:88) (cid:9)(cid:12)(cid:12)(cid:12)∆ j f [xiν+s, ..., xiν+s+ j] (cid:12)(cid:12)(cid:12)p min(cid:8)1, xiν+s+m − xiν+s (cid:88) iν+1−1− j(cid:88) m(cid:88) (cid:9)(cid:12)(cid:12)(cid:12)∆ j f [xiν+s, ..., xiν+s+ j] (cid:12)(cid:12)(cid:12)p min(cid:8)1, xiν+s+m − xiν+s iν+1−1− j(cid:88) (cid:88) m(cid:88) (cid:9)(cid:12)(cid:12)(cid:12)∆ j f [xiν+s, ..., xiν+s+ j] (cid:12)(cid:12)(cid:12)p min(cid:8)1, xiν+s+m − xiν+s (cid:96)2− j(cid:88) m(cid:88) min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆ j f [xi, ..., xi+ j] (cid:12)(cid:12)(cid:12)p i=iν∈L2 i=iν∈L2 s=0 j=0 s=0 s=0 j=0 j=0 . . , . Summarizing these inequalities over all i ∈ L2, we obtain that f (y0)p = ∆0g[y0]p ≤ C(m)p = C(m)p Since (cid:96)ν ≤ iν+1 − 1, we obtain that f (ai)p = f (y0)p ≤ C(m)p f (ai)p ≤ C(m)p = C(m)p (cid:88) i∈L2 B2 = Hence, clude that B2 ≤ C(m)p so that, thanks to Assumption 5.15, B2 ≤ C(m)p(cid:101)λp. i=(cid:96)1 j=0 Finally, recalling definitions of the quantities B, B1 and B2, see (5.91), (5.92), (5.93), we con- B = B1 + B2 ≤(cid:101)λp + C(m)p(cid:101)λp proving (5.91). This inequality and inequality (5.90) imply the required inequality (5.89). The proof of the lemma is complete. (cid:3) Now, inequalities (5.81), (5.88) and (5.89) imply that (cid:107)(cid:101)F(cid:107)Wm p (R) ≤ C(m) ((cid:107)(cid:101)F(cid:107)Lp(R) + (cid:107)(cid:101)F(cid:107)Lm p (R)) ≤ C(m)(cid:101)λ. 88 Since(cid:101)F ∈ Wm p (R) and(cid:101)FE = f , the function f belongs to the trace space Wm p (R) ≤ C(m)(cid:101)λ = C(m)(cid:102)Wm,p( f : E) p (R)E ≤ (cid:107)(cid:101)F(cid:107)Wm (cid:107) f(cid:107)Wm p (R)E. Furthermore, proving the sufficiency part of Theorem 1.3 in the case under consideration. Case 2. #E ≤ m. i=0 is a strictly increasing sequence, and 0 ≤ n < m. (In other We may assume that E = {xi}n words, we assume that (cid:96)1 = 0 and (cid:96)2 = n.) In this case mE = min{m, #E − 1} = n (see (1.5)), and Let f be a function on E, and let(cid:101)λ = (cid:102)Wm,p( f : E). This notation and equivalence (5.82) enable equivalence (5.82) holds with constants depending only on m. us to make the following Assumption 5.19 For every k = 0, ..., n, and every i = 0, ..., n − k, the following inequality ∆k f [xi, ..., xi+k] ≤ C(m)(cid:101)λ holds. Our aim is to prove the existence of a function(cid:101)F ∈ Wm We will do this by reduction of the problem to the Case 1. More specifically, we introduce m − n additional points xn+1, xn+2, ..., xm defined by p (R) with (cid:107)(cid:101)F(cid:107) ≤ C(m)(cid:101)λ such that(cid:101)FE = f . xk = xn + 2(k − n), (5.94) Let E = {x0, ..., xn, xn+1, ..., xm}. Clearly, E is a strictly increasing sequence in R with #E = m + 1. By ¯f : E → R we denote the extension of f from E to E by zero. Thus, Furthermore, E ⊂ E and E (cid:44) E. k = n + 1, ..., m. ¯f (xi) = f (xi) for every 0 ≤ i ≤ n, and ¯f (xi) = 0 for every n + 1 ≤ i ≤ m. (5.95) Prove that the following analogue of Assumption 5.15 (with (cid:96)1 = 0 and (cid:96)2 = m) holds for the sequence E = {xi}m i=0: (cid:102)Wm,p( ¯f : E)p = m(cid:88) m−k(cid:88) k=0 i=0 Let us prove that k=0 i=0 ∆k ¯f [xi, ..., xi+k] ≤ C(m)(cid:101)λ for every k = 0, ..., m and every i = 0, ..., m − k. 89 Let us note that, according to Convention 5.14, xi+m = +∞ provided i = 1, ..., n. Furthermore, thanks to (5.94), xm − x0 = xn + 2(m − n) − x0 ≥ 2. Hence, (cid:12)(cid:12)(cid:12)p ≤ C(m)p(cid:101)λp. min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k ¯f [xi, ..., xi+k] m(cid:88) (cid:12)(cid:12)(cid:12)∆k ¯f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p m−k(cid:88) . (cid:102)Wm,p( ¯f : E)p = (5.96) (5.97) (5.98) Assumption 5.19 tells us that (5.98) holds for every k = 0, ..., n and every i = 0, ..., n−k. It is also clear that (5.98) holds for each k > n and 0 ≤ i ≤ m − k (because in this case ∆k ¯f [xi, ..., xi+k] = 0). See (5.96). Let k ∈ {0, ..., n} and let i ∈ {n − k + 1, ..., m − k}. Then i ≤ n and i + k ≥ n + 1 so that, thanks to (5.94), xi+k − xi ≥ xn+1 − xn = 2. Therefore the diameter of the set S = {xi, ..., xi+k} is at least 2 which enables us to apply Lemma 5.4 to S and the function ¯f defined on S . This lemma tells us that there exists j ∈ {0, ..., k − 1} and ν ∈ {i, ..., i + k − j} such that xν+ j − xν ≤ 1 and ∆k ¯f [xi, ..., xi+k] ≤ 2k ∆ j ¯f [xν, ..., xν+ j]/ diam S . Since diam S ≥ 2 and 0 ≤ k ≤ m, we obtain that ∆k ¯f [xi, ..., xi+k] ≤ 2m ∆ j ¯f [xν, ..., xν+ j] . (5.99) Definition (5.94) and the inequality xν+ j − xν ≤ 1 imply that either ν ≥ n + 1 or ν + j ≤ n. If ν ≥ n +1, then ¯f (xν) = ... = ¯f (xν+ j) = 0, see (5.95), so ∆ j ¯f [xν, ..., xν+ j] = 0. Hence, ∆k ¯f [xi, ..., xi+k] = 0 proving (5.98) for this case. On the other hand, if ν + j ≤ n then ¯f (xα) = f (xα) for all α ∈ {ν, ..., ν + j}. See (5.95). This property and Assumption 5.19 imply that ∆ j ¯f [xν, ..., xν+ j] = ∆ j f [xν, ..., xν+ j] ≤ C(m)(cid:101)λ which together with (5.99) yields the required inequality (5.98) in the case under consideration. We have proved that inequality (5.98) holds for all k = 0, ..., m and all i = 0, ..., m − k. The required inequality (5.96) is immediate from (5.98) and formula (5.97). Thus, #E = m + 1 > m and an analogue of Assumption 5.15, inequality (5.96), holds for the function ¯f defined on E. This enables us to apply to E and ¯f the result of Case 1 which produces a function F ∈ Wm p (R) such that F E = ¯f and (cid:107)F (cid:107)Wm p (R) ≤ C(m)(cid:101)λ. Since ¯fE = f , we have F E = f proving that f ∈ Wm p (R)E and (cid:107) f(cid:107)Wm completes the proof of the sufficiency part of Theorem 1.3 in Case 2. p (R)E ≤ C(m)(cid:101)λ. This The proof of Theorem 1.3 is complete. (cid:3) 5.4. Wm p (R)-restrictions and local sharp maximal functions. In this section we prove Theorem 1.5. (Necessity.) Let F ∈ Wm every k = 0, ..., m, the following inequality p (R) be a function such that FE = f . Prove that for every x ∈ R and m(cid:88) k,E(x) ≤ 4 f # M[F( j)](x) (5.100) j=0 holds. (Recall that M[g] denotes the Hardy-Littlewood maximal function of a locally integrable function g on R. See (2.2).) Let 0 ≤ k ≤ m − 1, and let S = {x0, ..., xk}, x0 < ... < xk, be a (k + 1)-point subset of E such that dist(x, S ) ≤ 1. Thus, S ⊂ I = [x − 1, x + 1]. 90 Let us apply Lemma 5.1 to the function G = F, the interval I = [x − 1, x + 1] (with I = 2), q = 1 and the set S . This lemma tells us that ∆kF[S ] ≤ 2 Hence, ∆kF[S ] ≤ 4 1 I (cid:90) I I 1 I F(k)(y) dy + 4 I (cid:90) (cid:16)F(k)(y) + F(k+1)(y)(cid:17) (cid:90) dy. F(k+1)(y) dy ≤ 4{M[F(k)](x) + M[F(k+1)](x)}. This inequality and (1.7) imply (5.100) in the case under consideration. Prove (5.100) for k = m. Let x ∈ R and let S = {x0, ..., xm}, x0 < ... < xm, be an (m + 1)-point subset of E such that dist(x, S ) ≤ 1. Let I be the smallest closed interval containing S ∪ {x}. Clearly, I = diam(S ∪ {x}). This and inequality (2.11) imply that diam S diam(S ∪ {x}) ∆m f [S ] = xm − x0 I ≤ xm − x0 I ∆m f [S ] · 1 xm − x0 xm(cid:90) x0 (cid:90) I F(m)(t) dt. F(m)(t) dt ≤ 1 I This inequality and (1.8) tell us that f # m,E(x) ≤ M[F(m)](x) proving (5.100) for k = m. Now, inequality (5.100) and the Hardy-Littlewood maximal theorem imply that (cid:107) f # k,E(cid:107)Lp(R) ≤ 4(m + 1) (cid:107)M[F(k)](cid:107)Lp(R) ≤ C(m, p) (cid:107)F(k)(cid:107)Lp(R) = C(m, p)(cid:107)F(cid:107)Wm p (R). Taking the infimum in the right hand side of this inequality over all F ∈ Wm we obtain that p (R) such that FE = f , m(cid:88) k=0 m(cid:88) k=0 m(cid:88) m(cid:88) k=0 k=0 The proof of the necessity part of Theorem 1.5 is complete. (Sufficiency.) Let f be a function on E such that f # k,E ∈ Lp(R) for every k = 0, ..., m. Let (5.101) (5.102) (5.103) (cid:107) f # k,E(cid:107)Lp(R) ≤ C(m, p)(cid:107) f(cid:107)Wm p (R)E . (cid:99)Wm,p( f : E) = k,E(cid:107)Lp(R). (cid:107) f # p (R)E ≤ C(m)(cid:99)Wm,p( f : E). f ∈ Wm p (R)E and (cid:107) f(cid:107)Wm Wm,p( f : E) ≤ C(m)(cid:99)Wm,p( f : E) m(cid:88) k=0 91 Prove that We begin with the case of a set E containing at least m + 1 points. Our proof of (5.102) in this case relies on Theorem 1.2. Let us see that where Wm,p is the quantity defined by (1.3). Let n ∈ N, n ≥ m, and let S = {x0, ..., xn}, x0 < ... < xn, be a subset of E. Fix k ∈ {0, ..., m} and set n−k(cid:88) i=0 (cid:12)(cid:12)(cid:12)p min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] Ak( f : S ) = . (5.104) Let us estimate the quantity Ak( f : S ) basing on the following fact: for every i ∈ {0, ..., n− k} the following inequality min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤ 2mp m(cid:88) xi+m(cid:90) ( f # j,E)p(u) du (5.105) j=0 xi holds. We proceed by cases. Let S i = {xi, ..., xi+k}, and let ti = min{1, xi+m − xi}. Case A. diam S i = xi+k − xi ≤ ti. Let Vi = [xi, xi + ti]. Then Vi ≤ 1 and Vi ⊃ [xi, xi+k] ⊃ S i so that x ∈ Vi. dist(x, S i) ≤ 1 for every Therefore, thanks to (1.7), for every k = 0, ..., m − 1. In turn, if k = m, then ∆k f [S i] ≤ f # k,E(x), x ∈ Vi, (5.106) S i = {xi, ..., xi+m} ⊂ Vi ⊂ [xi, ..., xi+m] so that diam S i = diam(S i ∪ {x}) = xi+m − xi for every x ∈ Vi. This and (1.8) imply that ∆m f [S i] ≤ diam(S i ∪ {x}) diam S i · f # m,E(x) = f # m,E(x), x ∈ Vi, proving that inequality (5.106) holds for all k = 0, ..., m. Raising both sides of (5.106) to the power p and then integrating on Vi with respect to x, we obtain that for each k = 0, ..., m, the following inequality Vi(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p = min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p ≤ xi+m(cid:90) xi ( f # k,E)p(u) du holds. Of course, this inequality implies (5.105) in the case under consideration. Case B. diam S i = xi+k − xi > ti. Clearly, in this case ti = 1 (because xi+k − xi ≤ xi+m − xi) so that diam S i = xi+k − xi > 1. In particular, k ≥ 1. 92 Lemma 5.4 tells us that there exist j ∈ {0, ..., k− 1} and ν ∈ {i, ..., i + k− j} such that xν+ j − xν ≤ 1 and ∆k f [S i] ≤ 2k ∆ j f [xν, ..., xν+ j]/ diam S i ≤ 2m ∆ j f [xν, ..., xν+ j] . (5.107) Since xν+ j − xν ≤ 1 and xi+k − xi > 1, and [xν, xν+ j] ⊂ [xi, xi+k], there exists an interval Hi such Clearly, the set(cid:101)S = {xν, ..., xν+ j} ⊂ Hi so that dist(x,(cid:101)S ) ≤ 1 for every x ∈ Hi (because Hi = 1). that Hi = 1 and [xν, xν+ j] ⊂ Hi ⊂ [xi, xi+k]. ∆ j f [(cid:101)S ] ≤ f # ∆k f [S i] ≤ 2m f # Therefore, thanks to (1.7), proving that x ∈ Hi. x ∈ Hi, for every j,E(x) for all j,E(x) See (5.107). Raising both sides of this inequality to the power p and then integrating on Hi with respect to x, we obtain that min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p = (cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:90) (cid:12)(cid:12)(cid:12)p = Hi(cid:12)(cid:12)(cid:12)∆k f [S i] (cid:12)(cid:12)(cid:12)p xi+m(cid:90) j,E)p(u) du ≤ 2mp ( f # ≤ 2mp ( f # j,E)p(u) du Hi xi This implies (5.105) in Case B proving that this inequality holds for all k ∈ {0, ..., m} and all for every k = 0, ..., m. i ∈ {0, ..., n − k}. Inequality (5.105) enables us to show that Ak( f : S ) ≤ C(m)p(cid:99)Wm,p( f : E)p for every k = 0, ..., m. (5.108) See (5.104) and (5.101). To prove this inequality we introduce a family of closed intervals Note that each interval Ti0 T = {Ti = [xi, xi+m] : i = 0, ..., n − k}. = [xi0, xi0+m] ∈ T has at most 2m + 2 common points with each interval Ti ∈ T . In this case Lemma 4.14 tells us that there exists a positive integer κ ≤ 2m + 3 and subfamilies T(cid:96) ⊂ T , (cid:96) = 1, ..., κ, each consisting of pairwise disjoint intervals, such that T = ∪{T(cid:96) : (cid:96) = 1, ..., κ}. This property and (5.105) imply that Ak( f : S ) ≤ 2mp Ak,(cid:96)( f : S ) where Ak,(cid:96)( f : S ) = ( f # j,E)p(u) du . κ(cid:88) m(cid:88) (cid:90) (cid:96)=1 (cid:88) i:Ti∈T(cid:96) j=0 Ti 93 (cid:90) U(cid:96) m(cid:88) κ(cid:88) j=0 (cid:90) m(cid:88) m(cid:88) j=0 Ak,(cid:96)( f : S ) = j,E)p(u) du ≤ ( f # ( f # j,E)p(u) du = R j=0 (cid:107) f # j,E(cid:107)p Lp(R) so that Ak( f : S ) ≤ 2mp proving (5.108). Hence, Ak,(cid:96)( f : S ) ≤ 2mp κ (cid:96)=1 j=0 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p = m(cid:88) n−k(cid:88) k=0 i=0 Lp(R) ≤ (2m + 3) 2mp (cid:99)Wm,p( f : E)p j,E(cid:107)p (cid:107) f # m(cid:88) Ak( f : S ) ≤ C(m)p(cid:99)Wm,p( f : E)p. k=0 Let U(cid:96) = ∪{Ti : Ti ∈ T(cid:96)}, (cid:96) = 1, ..., κ. Since the intervals of each subfamily T(cid:96) are pairwise disjoint, m(cid:88) In this case Theorem 1.3 tells us that Taking the supremum in the left hand side of this inequality over all n ∈ N, n ≥ m, and all strictly increasing sequences S = {x0, ..., xn} ⊂ E, and recalling definition (1.3), we obtain the required inequality (5.103). We know that f # (5.103) imply that Wm,p( f : E) < ∞. Theorem 1.2 tells us that in this case p (R)E ≤ C(m)Wm,p( f : E). k,E ∈ Lp(R) for every k = 0, ..., m so that(cid:99)Wm,p( f : E) < ∞. This and inequality p (R)E ≤ C(m)(cid:99)Wm,p( f : E) proving (5.102) This inequality together with (5.103) imply that (cid:107) f(cid:107)Wm and the sufficiency for each set E ⊂ R with #E ≥ m + 1. It remains to prove the sufficiency for an arbitrary set E ⊂ R containing at most m points. Let 0 ≤ n < m, and let E = {x0, ..., xn}, x0 < ... < xn so that mE = min{m, #E − 1} = n, see (1.5). p (R)E and (cid:107) f(cid:107)Wm f ∈ Wm p (R)E ≤ C(m)(cid:102)Wm,p( f : E) n−k(cid:88) min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p (cid:102)Wm,p( f : E) =  n(cid:88) the above formula for(cid:102)Wm,p( f : E).) (We recall that xi+m = +∞ for all i = 0, ..., n, see Convention 5.14, so that min{1, xi+m − xi} = 1 in (cid:102)Wm,p( f : E) ≤ C(m)(cid:99)Wm,p( f : E). p (R)E ≤ C(m)(cid:99)Wm,p( f : E) proving the sufficiency for sets E ⊂ R with #E ≤ m. We literally follow the proof of inequality (5.103) and get its analog for the case under conside-  1 (cid:107) f(cid:107)Wm ration: where Hence, (cid:107) f(cid:107)Wm The proof of Theorem 1.5 is complete. k=0 i=0 p . (cid:3) 5.5. Further remarks and comments. ((cid:70)1) Let ExtE(· : Wm constructed in Section 5. We prove that this operator possesses the following property. p (R)), see (5.67), be the extension operator for the Sobolev space Wm p (R) 94 Statement 5.20 Let f be a function on E with Wm,p( f : E) < ∞ and let F = ExtE( f : Wm Then the support of F lies in a δ-neighborhood of E where δ = 3(m + 2): p (R)). supp F ⊂ [E]δ. Proof. Pick a point x ∈ R such that dist(x, E) ≥ δ and prove that F(x) = 0. Let x ∈ J = (aJ, bJ) where J is an interval from the family JE, see (4.5). (Recall that aJ, bJ ∈ E for each J = (aJ, bJ) ∈ JE and J ⊂ R \ E.) Since x ∈ (aJ, bJ) and aJ, bJ ∈ E, we have δ < dist(x, E) = min{x − aJ, bJ − x} so that bJ − aJ = x − aJ + bJ − x > 2δ > 4. Thus J > 4 proving that J ∈ GE, see (5.34). We recall that in this case we divide the interval J into nJ equal intervals with ends in points Y(J) n ∈ G ⊂ (cid:101)E for all n = 0, ..., nJ, see (5.42), (5.44). Furthermore, thanks to (5.46), n , n = 0, ..., nJ, see (5.36). The length (cid:96)(J) of each interval satisfies the inequality 2 ≤ (cid:96)(J) ≤ 3, see (5.38), (5.39) and (5.40). We also recall that Y(J) n = 1, ..., nJ − 1. k+1) for some k ∈ {0, ..., nJ − 1}. In particular, , Y(J) n ) = 0, f (Y(J) Let x ∈ (Y(J) k x − Y(J) k ≤ (cid:96)(J) ≤ 3 and Y(J) k+1 − x ≤ (cid:96)(J) ≤ 3. (5.109) Hence, We know that Y(J) k Hence, k − aJ = (x − aJ) − (x − Y(J) Y(J) k ) ≥ δ − 3. = aJ + (cid:96)(J) · k, so that k − aJ = (cid:96)(J) · k ≥ δ − 3. Y(J) k ≥ (δ − 3)/(cid:96)(J) ≥ (δ − 3)/3 ≥ m + 1. (5.110) In the same way we prove that nJ −(k +1) ≥ m +1. This inequality and (5.110) imply that the set V = {xk−m, ..., xk−m+1} ⊂ S J ⊂ G, see (5.37), (5.41) and (5.34). In particular, fV ≡ 0, see (5.109). of Propositions 3.11 and 3.12. See Section 3.1. For the set(cid:101)E and the points u, v ∈ (cid:101)E this procedure We recall the procedure of constructing of the sets Sx and points sx, x ∈ E, satisfying conditions provides sets S u, S v ⊂ (cid:101)E with #S u = #S v = m + 1. We also know that u ∈ S u, v ∈ S v, see part (i) v = Y(J) k+1. u = Y(J) k and Let Proposition 3.11. These properties together with the property (3.35) imply that S u, S v ⊂ V = {xk−m, ..., xk−m+1}. Hence, fS u ≡ 0 and fS v ≡ 0 proving that the Lagrange polynomials and Pv = LS v[ f ] = 0. Pu = LS u[ f ] = 0 95 (cid:3) Finally, since Pu = Pv = 0, the Hermite polynomials HI for the interval I = (u, v) = (Y(J) k See part (ii) of Definition 3.14 and formula (3.42). , Y(J) k+1), defined by formula (4.7), is zero as well. This and definition (4.8) tell us that FI ≡ 0 proving the required property F(x) = 0. The proof of the statement is complete. Let 0 < ε ≤ 1. We can slightly modify the construction of the extension (5.67) by changing the constant nJ from the formula (5.35) as follows: nJ = (cid:98)J/(γε)(cid:99) where γ is a certain constant depending only on m. Obvious changes in the proof of Theorem 1.2 enable us to show that after such a modification this theorem holds with the constants in equivalence (1.4) depending on m and ε. On the other hand, following the proof of Statement 5.20 one can readily prove that for γ = γ(m) big enough supp F ⊂ [E]ε. We leave the details to the interested reader. ((cid:70)2) As we have noted in Section 4.4, whenever E = {xi}(cid:96)2 Sobolev space Wm . See Remark 4.20. is a sequence of points in R, for p (R)) is an interpolating Cm−1-smooth spline of p (R)E the extension F = ExtE( f : Lm each f ∈ Lm order 2m with knots {xi}(cid:96)2 We note that the same statement holds for the extension operator ExtE(· : Wm tion: the set(cid:101)E, see (5.44), is a sequence of points in R whenever the set E is. ((cid:70)3) Definition (1.3) of the quantity Wm,p( f : E) (which controls the trace norm of f in p (R)E, see Theorem 1.2) is given in a compact form which implies that convention (1.2) holds. Wm For the reader's convenience below we present this definition in an explicit form: p (R)) for the normed p (R). This is immediate from formula (5.68) and the following obvious observa- i=(cid:96)1 i=(cid:96)1 Here the supremum is taken over all finite strictly increasing sequences {x0, ..., xn} ⊂ E with n ≥ m. (cid:96)2 − (cid:96)1 ≥ m, we can express in an explicit form the quantity(cid:102)Wm,p( f : E) defined by (1.6): in R such that (cid:96)2 < ∞ and In a similar way, given a strictly increasing sequence E = {xi}(cid:96)2 (cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p (cid:102)Wm,p( f : E) = ((cid:70)4) Let 0 ≤ n < m and let E = {x0, ..., xn} where x0 < ... < xn. Thus, in this case mE = min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p +  m(cid:88) (cid:96)2−m(cid:88) m−1(cid:88) (cid:96)2−k(cid:88)  1 i=(cid:96)2−m+1 i=(cid:96)1 i=(cid:96)1 k=0 k=0 . p min{m, #E − 1} = n, see (1.5). Equivalence (5.82) and Theorem 1.3 tell us that p (R)E ∼ max{∆k f [xi, ..., xi+k] : k = 0, ..., n, i = 0, ..., n − k} (cid:107) f(cid:107)Wm (5.111) 96 min{1, xi+m − xi}(cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p  m(cid:88) n−m(cid:88) (cid:12)(cid:12)(cid:12)∆k f [xi, ..., xi+k] (cid:12)(cid:12)(cid:12)p k=0 i=0  1 p . Wm,p( f : E) = + sup {x0,...,xn}⊂E m−1(cid:88) n−k(cid:88) k=0 i=n−m+1 with constants in this equivalence depending only on m. Slightly modifying the proof of Theorem 1.3 for the case p = ∞, we can show that for each strictly increasing sequence S = {s0, s1, ..., s(cid:96)}, (cid:96) ∈ N, and every f : S → R the following equiva- lence (cid:107) f(cid:107)Wn∞(R)S ∼ max{∆k f [si, ..., si+k] : k = 0, ..., n, i = 0, ..., (cid:96) − k} holds. The constants in this equivalence depend only on n. This equivalence and (5.111) imply that p (R)E = Wn∞(R)E provided n = #E − 1 < m, Wm (5.112) with imbedding constants depending only on m. In connection with isomorphism (5.112), we recall a classical Sobolev imbedding theorem which states that p (R) (cid:44)→ Wn∞(R) Wm but Wm p (R) (cid:44) Wn∞(R). Comparing this result with (5.112), we observe that, even though the spaces Wm p (R) and Wn∞(R) are distinct, their traces to every "small" subset of R containing at most m points, coincide with each other. 6. The Finiteness Principle for Lm∞(R) traces: multiplicative finiteness constants. 6.1. Multiplicative finiteness constants of the space Lm∞(R). Let m ∈ N. Everywhere in this section we assume that E is a closed subset of R with #E ≥ m +1. We will discuss equivalence (1.12) which states that (cid:107) f(cid:107)Lm∞(R)E ∼ sup S⊂E, #S =m+1 ∆m f [S ] (6.1) for every function f ∈ Lm∞(R)E. The constants in this equivalence depend only on m. We can interpret this equivalence as a special case of the following Finiteness Principle for the space Lm∞(R). Theorem 6.1 Let m ∈ N. There exists a constant γ = γ(m) > 0 depending only on m, such that the following holds: Let E ⊂ R be a closed set, and let f : E → R. For every subset E(cid:48) ⊂ E with at most N = m + 1 points, suppose there exists a function FE(cid:48) ∈ Lm∞(R) with the seminorm (cid:107)FE(cid:48)(cid:107)Lm∞(R) ≤ 1, such that FE(cid:48) = f on E(cid:48). Then there exists a function F ∈ Lm∞(R) with the seminorm (cid:107)F(cid:107)Lm∞(R) ≤ γ such that F = f on E. Proof. Applying equivalence (6.1) to an arbitrary set E = S with #S = m + 1, we conclude that (cid:107) f(cid:107)Lm∞(R)S ∼ ∆m f [S ]. This equivalence and the theorem's hypothesis tell us that ∆m f [S ] ≤ C1(m) for every S ⊂ E with #S = m + 1. This inequality and definition (1.11) imply that Lm,∞( f : E) ≤ C1(m), so that, by (1.12), (cid:107) f(cid:107)Lm∞(R)E ≤ C2(m)Lm,∞( f : E) ≤ C2(m) C1(m) = γ(m), and the proof of Theorem 6.1 is complete. (cid:3) 97 We refer to the number N = m + 1 as the finiteness number of the space Lm∞(R). Clearly, the value N(m) = m + 1 in the finiteness Theorem 6.1 is sharp; in other words, Theorem 6.1 is false in general if N = m + 1 is replaced by some number N < m + 1. Theorem 6.1 implies the following inequality: For every f ∈ Lm∞(R)E we have (cid:107) f(cid:107)Lm∞(R)E ≤ γ(m) sup S⊂E, #S =m+1 (cid:107) fS(cid:107)Lm∞(R)S . (6.2) (Clearly, the converse inequality is trivial and holds with γ(m) = 1.) Inequality (6.2) motivates us to call the constant γ = γ(m) the multiplicative finiteness constant for the space Lm∞(R). The following natural question arises: Question 6.2 What is the sharp value of the multiplicative finiteness constant for Lm∞(R)? We denote this sharp value of γ(m) by γ(cid:93)(Lm∞(R)). Thus, (cid:107) f(cid:107)Lm∞(R)E γ(cid:93)(Lm∞(R)) = sup sup{(cid:107) fS(cid:107)Lm∞(R)S : S ⊂ E, #S = m + 1} (6.3) where the supremum is taken over all closed sets E ⊂ R with #E ≥ m + 1, and all functions f ∈ Lm∞(R)E. The next theorem answers to Question 6.2 for m = 1, 2 and provides lower and upper bounds for γ(cid:93)(Lm∞(R)) for m > 2. These estimates show that γ(cid:93)(Lm∞(R)) grows exponentially as m → +∞. Theorem 6.3 We have: (i) γ(cid:93)(L1∞(R)) = 1 and γ(cid:93)(L2∞(R)) = 2. (ii) For every m ∈ N, m > 2, the following inequalities (cid:18) π (cid:19)m−1 hold. 2 < γ(cid:93)(Lm∞(R)) < (m − 1) 9m (6.4) The proof of this theorem relies on works [17,18,23] devoted to calculation of a certain constant K(m) related to optimal extensions of Lm∞(R)-functions. This constant is defined by (cid:107) f(cid:107)Lm∞(R)X K(m) = sup max{m!∆m f [xi, ..., xi+m] : i = 1, ..., n} (6.5) where the supremum is taken over all n ∈ N, all finite strictly increasing sequences X = {x1, ..., xm+n} ⊂ R, and all functions f ∈ Lm∞(R)X. The constant K(m) was introduced by Favard [23]. (See also [17, 18].) Favard [23] proved that K(2) = 2, and de Boor found efficient lower and upper bounds for K(m). We prove that γ(cid:93)(Lm∞(R)) = K(m), see Proposition 6.8 below. This formula and aforementioned results of Favard and de Boor imply the required lower and upper bounds for γ(cid:93)(Lm∞(R)) in Theorem 6.3. We will need a series of auxiliary lemmas. 98 Lemma 6.4 Let S ⊂ R, #S = m + 1, and let f : S → R. Then = m!∆m f [S ] . (cid:107) f(cid:107)Lm∞(R)S Proof. Let A = min S , B = max S . Let F ∈ Lm∞(R) be an arbitrary function such that FS = f . Inequality (2.13) tells us that m!∆m f [S ] = m!∆mF[S ] ≤ (cid:107)F(cid:107)Lm∞(R) . Taking the infimum in this inequality over all functions F ∈ Lm∞(R) such that FS = f , we obtain that m!∆m f [S ] ≤ (cid:107) f(cid:107)Lm∞(R)S . Let us prove the converse inequality. Let F = LS [ f ] be the Lagrange polynomial of degree at most m interpolating f on S . Then m!∆m f [S ] = L(m) S [ f ] = (cid:107)F(cid:107)Lm∞(R) . See (2.5). Since FS = f , we obtain that (cid:107) f(cid:107)Lm∞(R)S ≤ (cid:107)F(cid:107)Lm∞(R) = m!∆m f [S ] (cid:3) proving the lemma. Lemma 6.5 Let E = {x1, ..., xm+n} ⊂ R, n ∈ N, be a strictly increasing sequence, and let f be a function on E. Then max S⊂E, #S =m+1 ∆m f [S ] = max ∆m f [xi, ..., xi+m] . i=1,...,n Proof. The lemma is immediate from the property ((cid:70)7), (i), Section 2.1 (see (2.14)). (cid:3) j=−∞ be a strictly increasing sequence in R, and let f : E → R. Then Lemma 6.6 Let E = {x j}∞ sup{(cid:107) fS(cid:107)Lm∞(R)S : S ⊂ E, #S = m + 1} = sup i m!∆m f [xi, ..., xi+m] . Proof. The lemma is immediate from Lemma 6.4 and Lemma 6.5. Lemma 6.7 Let E ⊂ R be a closed set, and let f ∈ Lm∞(R)E. Then (cid:107) fE(cid:48)(cid:107)Lm∞(R)E(cid:48) . (cid:107) f(cid:107)Lm∞(R)E = sup E(cid:48)⊂E, #E(cid:48)<∞ (cid:3) Proof. We recall the following well known fact: for every closed bounded interval I ⊂ R a ball in the space Lm∞(I) is a precompact subset in the space C(I). The lemma readily follows from this statement. We leave the details of the proof to the interested reader. Proposition 6.8 For every m ∈ N the following equality γ(cid:93)(Lm∞(R)) = K(m) (cid:3) holds. 99 Proof. Lemma 6.4, Lemma 6.5 and definition (6.5) tell us that (cid:107) f(cid:107)Lm∞(R)E K(m) = sup sup{(cid:107) fS(cid:107)Lm∞(R)S : S ⊂ E, #S = m + 1} (6.6) where the supremum is taken over all finite subsets E ⊂ R and all functions f on E. Comparing (6.6) with (6.3), we conclude that K(m) ≤ γ(cid:93)(Lm∞(R)). Let us prove the converse inequality. Lemma 6.7 and (6.6) imply that for every closed set E ⊂ R and every function f ∈ Lm∞(R)E the following is true: (cid:107) f(cid:107)Lm∞(R)E = sup E(cid:48)⊂E, #E(cid:48)<∞ sup ≤ = K(m) E(cid:48)⊂E, #E(cid:48)<∞ (cid:107) fE(cid:48)(cid:107)Lm∞(R)E(cid:48) K(m) sup S⊂E(cid:48), #S =m+1 (cid:107) fS(cid:107)Lm∞(R)S . sup S⊂E, #S =m+1 (cid:107) fS(cid:107)Lm∞(R)S This inequality and definition (6.3) imply the required inequality γ(cid:93)(Lm∞(R)) ≤ K(m) proving the proposition. (cid:3) Proof of Theorem 6.1. The equality γ(cid:93)(L1∞(R)) = 1 is immediate from the well known fact that a function satisfying a Lipschitz condition on a subset of R can be extended to all of R with preservation of the Lipschitz constant. As we have mentioned above, due to a result of Favard [23]. de Boor [17, 18] proved that K(2) = 2 (cid:18) π (cid:19)m−1 2 Here and < θm ≤ K(m) ≤ Θm < (m − 1) 9m for each m > 2. θm = (cid:18) π 2 (cid:19)m+1(cid:44) ∞(cid:88) m(cid:88) j=−∞ Θm = (2m−2/m) + ((−1) j/(2 j + 1))m+1 (cid:32)m (cid:33)(cid:32)m − 1 (cid:33) i − 1 i i=1 4m−i .  (6.7) (6.8) On the other hand, Proposition 6.8 tells us that γ(cid:93)(Lm∞(R)) = K(m) which together with estimates (6.7) and (6.8) implies the statements (i) and (ii) of the theorem. The proof of Theorem 6.1 is complete. (cid:3) 6.2. Extremal functions and finiteness constants. • An extremal function for the lower bound in Theorem 6.1. Note that θ1 = 1 and θ2 = 2, so that, by part (i) of Theorem 6.3, θm = γ(cid:93)(Lm∞(R)) = K(m) for m = 1, 2. 100 In turn, Proposition 6.8 and inequalities (6.8) imply that θm ≤ γ(cid:93)(Lm∞(R)) ≤ Θm for every m ∈ N (6.9) which slightly improves the lower and upper bounds for γ(cid:93)(Lm∞(R)) given in (6.4). Since lim m→∞ θm (π/2)m+1 = 1/2, see [17], we obtain the following asymptotic lower bound for γ(cid:93)(Lm∞(R)): lim m→∞ γ(cid:93)(Lm∞(R)) (π/2)m+1 ≥ 1/2 . We also note that, the proof of the inequality θm ≤ K(m) (= γ(cid:93)(Lm∞(R))) given in [17], is constructive, i.e., this proof provides an explicit formula for a function on R and a set E ⊂ R for which this lower bound for K(m) is attained. More specifically, let E = Z = {0,±1,±2, ...} be the set of all integers, and let f : E → R be a function on E defined by f (i) = (−1)i for every i ∈ E. We let Em : R → R denote the Euler spline introduced by Schoenberg [55]: t ∈ R. (−1)i Mm+1[S i](t + (m + 1)/2), Em(t) = θm (cid:88) i∈Z Here S i = {i, ..., i + m + 1}, and Mk[S ] is the B-spline defined by formula (2.7). Figure 1: The graph of the Euler spline E2. We refer the reader to the monograph [55] for various remarkable properties of this spline. Let us list some of them: 101 even, the same true on every interval [i − 1/2, i + 1/2] ; • Em ∈ Cm−1(R) ; • If m is odd, Em is a polynomial of degree at most m on every interval [i, i + 1], i ∈ Z; if m is • Em(x + 1) = −Em(x) for all x ∈ R ; • Em(i + 1/2) = 0 for every i ∈ Z ; • E(m) m (x) = (−1)m (cid:107)Em(cid:107)Lm∞(R) sign(sin(πx)) for all x ∈ R ; • (cid:107)Em(cid:107)Lm∞(R) = θm 2m, and Em(i) = (−1)i for each i ∈ Z. In particular, EmE = f . Furthermore, one can readily see that m!∆m f [i, ..., i + m] = 2m for every i ∈ Z . (6.10) It is also proven in [17] that (cid:107)Em(cid:107)Lm∞(R) ≤ (cid:107)F(cid:107)Lm∞(R) for every function F ∈ Lm∞(R) such that FE = f . This enables us to calculate the trace norm of f in Lm∞(R): (cid:107) f(cid:107)Lm∞(R)E = (cid:107)Em(cid:107)Lm∞(R) = θm 2m . This equality, (6.10), definition (6.3) and Lemma 6.6 imply the following inequality γ(cid:93)(Lm∞(R)) ≥ (cid:107) f(cid:107)Lm∞(R)E sup{(cid:107) fS(cid:107)Lm∞(R)S : S ⊂ E, #S = m + 1} = (cid:107) f(cid:107)Lm∞(R)E supi∈Z m!∆m f [xi, ..., xi+m] = θm proving the first inequality in (6.9). • Finiteness numbers and multiplicative finiteness constants for the space Lm∞(Rn). We identify the homogeneous Sobolev space Lm∞(Rn) with the space Cm−1,1(Rn) of all Cm−1- functions on Rn whose partial derivatives of order m − 1 satisfy the Lipschitz condition on Rn. We seminorm Lm∞(Rn) by  1 2 (cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L∞(Rn) α=m . (cid:107)F(cid:107)Lm∞(Rn) = (cid:107)∇mF(cid:107)L∞(Rn) where ∇mF = DαF2 . We let Wm∞(Rn) denote the corresponding Sobolev space of all functions F ∈ Lm∞(Rn) equipped with the norm (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) max k=0,...,m (cid:107)F(cid:107)Wm∞(Rn) = ∇kF We recall the Finiteness Principle for the space Lm∞(Rn): Theorem 6.9 There exist a positive integer N and a constant γ > 0 depending only on m and n, such that the following holds: Let E ⊂ Rn be a closed set, and let f : E → R. For every subset E(cid:48) ⊂ E with at most N points, suppose there exists a function FE(cid:48) ∈ Lm∞(Rn) with (cid:107)FE(cid:48)(cid:107)Lm∞(Rn) ≤ 1, such that FE(cid:48) = f on E(cid:48). Then there exists a function F ∈ Lm∞(Rn) with (cid:107)F(cid:107)Lm∞(Rn) ≤ γ such that F = f on E. 102 See [57] for the case m = 2, and [24] for the general case m ∈ N. Note that the Finiteness Principle for the space Wm∞(Rn) holds as well. See [24]. We say that a number N ∈ N is a finiteness number (for the space Lm∞(Rn)) if there exists a constant γ > 0 such that the Finiteness Principle formulated above holds with N and γ. In this case for every function f ∈ Lm∞(Rn)E the following inequality (cid:107) f(cid:107)Lm∞(Rn)E ≤ γ (cid:107) fS(cid:107)Lm∞(Rn)S sup S⊂E, #S≤N (6.11) holds. We let N(cid:93)(Lm∞(Rn)) denote the sharp finiteness number for the space Lm∞(Rn). In the same way we introduce the notion of the finiteness number and the sharp finiteness number for the space Wm∞(Rn). As we have noted above, N(cid:93)(Lm∞(R)) = m + 1. It was proven in [57] that N(cid:93)(L2∞(Rn)) = 3 · 2n−1. We also know that See [3, 58]. This is the smallest upper bound for N(cid:93)(Lm∞(Rn)), m > 2, known to the moment. In [58] we conjecture that N(cid:93)(Lm∞(Rn)) ≤ 2(m+n n ). (cid:16)m+n−k m+1(cid:89) (cid:17) N(cid:93)(Lm∞(Rn+1)) = n−1 k . (cid:16) (cid:96)−1 (cid:17) k=1 = 0 for (cid:96) ∈ N.) (In this formula we set that (cid:16)−1−1 (cid:17) = 1 and Given a finiteness number N (for Lm∞(Rn)) we let γ(cid:93)(N; Lm∞(Rn)) denote the infimum of constants γ from inequality (6.11). We refer to γ(cid:93)(N; Lm∞(Rn)) as the multiplicative finiteness constant asso- ciated with the finiteness number N. Thus, γ(cid:93)(N; Lm∞(Rn)) = sup sup{(cid:107) fS(cid:107)Lm∞(Rn)S : S ⊂ E, #S ≤ N} (cid:107) f(cid:107)Lm∞(Rn)E where the supremum is taken over all closed sets E ⊂ Rn and all functions f ∈ Lm∞(Rn)E. Clearly, γ(cid:93)(N; Lm∞(Rn)) is a non-increasing function of N. If N = N(cid:93)(Lm∞(Rn)), we write γ(cid:93)(Lm∞(Rn)) rather than γ(cid:93)(N(cid:93)(Lm∞(Rn)); Lm∞(Rn)). We refer to γ(cid:93)(Lm∞(Rn)) as the sharp multiplicative finiteness constant (for Lm∞(Rn)). In the same fashion we introduce the constants γ(cid:93)(N; Wm∞(Rn)) and γ(cid:93)(Wm∞(Rn)). In fact we know very little about the values of the constants γ(cid:93)(N;·) and γ(cid:93)(·). Apart from the results related to γ(cid:93)(Lm∞(R)) which we present Section 6.1, there is perhaps only one other result in this direction. It is due to Fefferman and Klartag [31]: Theorem 6.10 For any positive integer N, there exists a finite set E ⊂ R2 and a function f : E → R with the following properties 1. For any W2∞-function F : R2 → R with FE = f we have that (cid:107)F(cid:107)W2∞(R2) > 1 + c0. 2. For any subset S ⊂ E with #S ≤ N, there exists an W2∞-function FS : R2 → R with FSS = f and (cid:107)FS(cid:107)W2∞(R2) ≤ 1. Here, c0 > 0 is a universal constant. 103 This result admits the following equivalent reformulation in terms of multiplicative finiteness constants for the space W2∞(R2): The following inequality inf N∈N γ(cid:93)(cid:16) N→∞ γ(cid:93)(cid:16) (cid:16)R2(cid:17)(cid:17) (cid:16)R2(cid:17)(cid:17) Since γ(cid:93)(cid:16) holds. N; W2∞ (cid:16)R2(cid:17)(cid:17) References is a non-increasing function of N, the above inequality implies that N; W2∞ > 1 lim N; W2∞ > 1. [1] J. C. Archer, E. Le Gruyer, A Constructive Proof of a Whitney Extension Theorem in One Variable, J. Approx. Theory 71 (1992) 312 -- 328. [2] I. S. Berezin, N. P. Zhidkov, Computing methods, v. 1 and 2, Pergamon Press, London, 1965. [3] E. Bierstone, P. D. Milman, Cm norms on finite sets and Cm extension criteria, Duke Math. J. 137 (2007) 1 -- 18. [4] Yu. Brudnyi, Criteria for the existence of derivatives in Lp. (Russian) Mat. Sb. (N.S.), 73 No. 115 (1967) 42 -- 64. [5] Yu. A. Brudnyi, Spaces that are definable by means of local approximations. (Russian) Trudy Moscov. Math. Obshch. 24 (1971) 69 -- 132; English transl.: Trans. Moscow Math. Soc. 24 (1974) 73 -- 139. [6] Yu. Brudnyi, P. Shvartsman, Generalizations of Whitney Extension Theorem, Intern. Math. Research Notices, No. 3, (1994) 129 -- 139. [7] Yu. Brudnyi, P. Shvartsman, The Whitney Problem of Existence of a Linear Extension Operator, J. Geom. Anal., 7, No. 4, (1997) 515 -- 574. [8] Yu. Brudnyi, P. Shvartsman, Whitney Extension Problem for Multivariate C1,ω-functions, Trans. Amer. Math. Soc. 353 No. 6, (2001) 2487 -- 2512. [9] A. P. Calder´on, Estimates for singular integral operators in terms of maximal functions, Studia Math. 44 (1972) 563 -- 582. [10] A. P. Calder´on, R. Scott, Sobolev type inequalities for p > 0, Studia Math. 62 (1978) 75 -- 92. [11] C. K. Chui and P. W. Smith, On Hm,∞-splines, SIAM J. Numer. Anal. 11 (1974) 554 -- 558. [12] C. K. Chui, P. W. Smith, and J. D. Ward, Favard's solution is the limit of Wk,p-splines, Trans. Amer. Math. Soc. 220 (1976) 299 -- 305. [13] C. K. Chui, P. W. Smith, and J. D. Ward, Quasi-Uniqueness in L∞ Extremal Problems, J. Approx. Theory 18 (1976) 213 -- 219. 104 [14] E. Cohen, T. Lyche, and R. Rosenfeld, Discrete B-Splines and Subdivision Techniques in Computer-Aided Geometric Design and Computer Graphics, Computer graphics and image processing 14 (1980) 87 -- 111. [15] C. de Boor, Best approximation properties of spline of odd degree, J. of Math. Mech., 12 (1963) 747 -- 749. [16] C. de Boor, A remark concerning perfect splines, Bull. Amer. Math. Soc. 80 (1974) 724 -- 727. [17] C. de Boor, How small can one make the derivatives of an interpolating function? J. Approx. Theory, 13 (1975) 105 -- 116. [18] C. de Boor, A smooth and local interpolant with k-th derivative, In: Numerical Solutions of Boundary Value Problems for Ordinary Differential Equations, (Proc. Sympos., Univ. Maryland, Baltimore, Md., 1974), pp. 177 -- 197. Academic Press, New York, 1975. [19] C. de Boor, On "best" interpolation, J. Approx. Theory 16 (1976) 28 -- 42. [20] C. de Boor, Splines as linear combinations of B-splines. A survey, Approximation theory, II (Proc. Internat. Sympos., Univ. Texas, Austin, Tex., 1976), pp. 1 -- 47. Academic Press, New York, 1976. [21] R. DeVore, G. Lorentz, Constructive approximation, Fundamental Principles of Mathe- matical Sciences, 303. Springer-Verlag, Berlin, 1993. [22] D. Est´evez, Explicit traces of functions from Sobolev spaces and quasi-optimal linear interpolators, Math. Inequal. Appl. 20 (2017) 441 -- 457. [23] J. Favard, Sur l'interpolation, J. Math. Pures Appl. 19 (1940) 281 -- 306. [24] C. Fefferman, A sharp form of Whitney extension theorem, Annals of Math. 161, No. 1, (2005) 509 -- 577. [25] C. Fefferman, A Generalized Sharp Whitney Theorem for Jets, Rev. Mat. Iberoamericana 21, No. 2, (2005) 577 -- 688. [26] C. Fefferman, Cm Extension by Linear Operators, Annals of Math. 166, No. 3, (2007) 779 -- 835. [27] C. Fefferman, Extension of Cm,ω-Smooth Functions by Linear Operators, Rev. Mat. Iberoamericana 25, No. 1, (2009) 1 -- 48. [28] C. Fefferman, Whitney extension problems and interpolation of data, Bulletin A.M.S. 46, No. 2, (2009) 207 -- 220. [29] C. Fefferman, The Cm norm of a function with prescribed jets I., Rev. Mat. Iberoameri- cana 26 (2010) 1075 -- 1098. [30] C. Fefferman, A. Israel, G. K. Luli, Sobolev extension by linear operators, J. Amer. Math. Soc. 27 (2014) 69 -- 145. 105 [31] C. Fefferman, B. Klartag, An example related to Whitney extension with almost minimal Cm norm, Rev. Mat. Iberoamericana 25 (2009) 423 -- 446. [32] S. Fisher, J. Jerome, The existence, characterization and essential uniqueness of solutions of L∞ extremal problems, Trans. Amer. Math. Soc. 187 (1974) 391 -- 404. [33] S. Fisher, J. Jerome, Minimum Norm Extremals in Function Spaces, Lect. Notes Math., 479 (1975). [34] A. Frolicher, A. Kriegl, Linear spaces and differentiation theory, Pure and Applied Math- ematics, J. Wiley, Chichester, 1988. [35] A. Frolicher, A. Kriegl, Differentiable extensions of functions, Differ. Geom. Appl. 3 (1993) 71 -- 90. [36] M. Golomb, Hm,p-extensions by Hm,p-splines, J. Approx. Theory 5 (1972) 238 -- 275. [37] A. Israel, A Bounded Linear Extension Operator for L2,p(R2), Annals of Math. 178 (2013) 1 -- 48. [38] A. Jakimovski, D. C. Russell, On an interpolation problem and spline functions. Gene- ral inequalities, 2 (Proc. Second Internat. Conf., Oberwolfach, 1978), pp. 205 -- 231, Birkhuser, Basel-Boston, Mass., 1980. [39] T. R. Jensen, B. Toft, Graph coloring problems, Wiley-Interscience Series in Discrete Mathematics and Optimization. A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1995. [40] J. W. Jerome, L. L. Schumaker, Characterizations of Functions with Higher Order Derivatives in Lp, Trans. Amer. Math. Soc. 143 (1969) 363 -- 371. [41] A. Jonsson, The Trace of the Zygmund Class Λ(R) to Closed Sets and Interpolating Polynomials, J. Approx. Th. 4 (1985) 1 -- 13. [42] S. Karlin, Interpolation properties of generalized perfect splines and the solutions of certain extremal problems. I, Trans. Amer. Math. Soc. 206 (1975) 25 -- 66. [43] A. Kriegl, P. W. Michor, The convenient setting of global analysis, Math. surveys and monographs, v. 53, AMS, 1997. [44] T. Kunkle, Using quasi-interpolants in a result of Favard. Approximation and computa- tion (West Lafayette, IN, 1993), 353 -- 357, Internat. Ser. Numer. Math., 119, Birkhuser Boston, Boston, MA, 1994. [45] T. Kunkle, Favard's interpolation problem in one or more variables, Constr. Approx. 18 (2002), no. 4, 467 -- 478. [46] G.K. Luli, Whitney Extension for Wk p(E) in One Dimension, (2008). Notes available at https://www.math.ucdavis.edu/ kluli/LmpNotes.pdf [47] T. Lyche, K. Morken, Making the Oslo algorithm more efficient, SIAM J. Numer. Anal. 23, No. 3, (1986) 663675. 106 [48] J. Merrien, Prolongateurs de fonctions differentiables d'une variable r´eelle, J. Math. Pures et Appl. 45 (1966) 291 -- 309. [49] A. Pinkus, On smoothest interpolants, SIAM J. Math. Anal. 19 (1988) 1431 -- 1441. [50] A. Pinkus, Uniqueness of Smoothest Interpolants, East J. Approx. 3 (1997) 377 -- 380. [51] F. Riesz, Untersuchungen uber Systeme integrierbarer Funktionen, Math. Ann. 69 (1910) 449 -- 497. [52] I. J. Schoenberg, On interpolation by spline functions and its minimal properties. 1964 On Approximation Theory (Proceedings of Conference in Oberwolfach, 1963) pp. 109 -- 129 Birkhuser, Basel. [53] I. J. Schoenberg, Spline interpolation and the higher derivatives, Proc. Nar. Acad. Sci. U.S.A. 51 (1964) 24 -- 28. [54] I. J. Schoenberg, Spline interpolation and the higher derivatives, 1969 Number Theory and Analysis, pp. 279 -- 295, Plenum, New York. [55] I. J. Schoenberg, Cardinal Spline Interpolation, CBMS, SIAM, Philadelphia, 1973. [56] I. A. Shevchuk, Extension of functions which are traces of functions belonging to Hϕ k on arbitrary subset of the line, Anal. Math., No. 3, (1984) 249 -- 273. [57] P. Shvartsman, On the traces of functions of the Zygmund class, Sib. Mat. Zh. 28 (1987) 203 -- 215; English transl. in Sib. Math. J. 28 (1987) 853 -- 863. [58] P. Shvartsman, The Whitney extension problem and Lipschitz selections of set-valued mappings in jet-spaces, Trans. Amer. Math. Soc. 360 (2008) 5529 -- 5550. [59] P. Shvartsman, Sobolev W1 p-spaces on closed subsets of Rn, Advances in Math. 220 (2009) 1842 -- 1922. [60] P. Shvartsman, Sobolev L2 p-functions on closed subsets of R2, Advances in Math. 252 (2014) 22 -- 113. [61] P. Shvartsman, Extension criteria for homogeneous Sobolev spaces of functions of one variable, arXiv:1812.00817 [62] P. Shvartsman, Sobolev functions on closed subsets of the real line, arXiv:1710.07826 [63] P. Shvartsman, Whitney-type extension theorems for jets generated by Sobolev func- tions, Advances in Math. 313 (2017) 379 -- 469. [64] P. W. Smith, Hr,∞(R) and Wr,∞(R)-splines, Trans. Amer. Math. Soc. 192 (1974) 275 -- 284. [65] P. W. Smith, Wr,p(R)-splines, J. Approx. Th. 10 (1974) 337 -- 357. [66] H. Whitney, Analytic extension of differentiable functions defined in closed sets, Trans. Amer. Math. Soc. 36 (1934) 63 -- 89. [67] H. Whitney, Differentiable functions defined in closed sets. I., Trans. Amer. Math. Soc. 36 (1934) 369 -- 387. 107
1509.01437
2
1509
2015-10-09T12:33:44
Explicit construction of non-stationary frames for $L^2$
[ "math.FA" ]
We show the existence of a family of frames of $L^2(\mathbb{R})$ which depend on a parameter $\alpha\in [0,1]$. If $\alpha=0$, we recover the usual Gabor frame, if $\alpha=1$ we obtain a frame system which is closely related to the so called DOST basis, first introduced by Stockwell and then analyzed by Battisti and Riba. If $\alpha\in (0,1)$, the frame system is associated to a so called $\alpha$-partitioning of the frequency domain. Restricting to the case $\alpha=1$, we provide a truly $n$-dimensional version of the DOST basis and an associated frame of $L^2(\mathbb{R}^d)$.
math.FA
math
EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 UBERTINO BATTISTI, MICHELE BERRA Abstract. We show the existence of a family of frames of L2pRq which depend on a parameter α P r0, 1s. If α " 0, we recover the usual Gabor frame, if α " 1 we obtain a frame system which is closely related to the so called DOST basis, first introduced in [30] and then analyzed in [4]. If α P p0, 1q, the frame system is associated to a so called α-partitioning of the frequency domain. Restricting to the case α " 1, we provide a truly n-dimensional version of the DOST basis and an associated frame of L2pRdq. 1. Introduction One of the most intriguing problems of modern Time-Frequency analysis is the construction of new efficient methods to represent signals, which can be one dimensional or, more often, multidi- mensional, such as digital images. The increasing amount of data and their complexity forces to develop optimized techniques that address the representation in a fast and efficient way. The starting point of this paper is the definition of the S-transform, first introduced by R. G. Stockwell et al. in [31] as (1.1) pS fqpb, ξq " p2πq´ 1 2 ż The expression of the S-transform (1.1) is formally very similar to the so called Short Time Fourier Transform or Gabor Transform (STFT) with Gaussian window dt. 2 e´2πi tξξfptqe´ ξ2pt´bq2 ż pSTFT fqpb, ξq " p2πq´ 1 2 e´2πi tξfptqe´ pt´bq2 2 dt. The main novelty of the S-transform is the frequency depending window. The leading idea is the heuristic fact that, in order to detect high frequencies, it is enough to consider a shorter time. Therefore, the width of the Gaussian is not fixed but depends on the frequency, shirking as far as the frequency increases. The S-transform was introduced to improve the analysis of seismic imaging, and it is now considered an important tool in geophysics, see [1]. In [21], the connection between the phase of S-transform and the instantaneous frequency - useful in several applications - has been studied. See [5, 6, 13, 18, 22, 24, 26, 36] for some applications of the S-transform to signal processing in general. From the mathematical point of view, M. W. Wong and H. Zhu in [35] introduced a generalized version of the S-transform as follows pSϕ fqpb, ξq " e´2πi tξ f ptqξ ϕpξ pt ´ bqq dt, b, ξ P R, ż R where ϕ is a general window function in L2pRq. The S-transform has a strong similarity with the STFT, actually it is also possible to show a deep link with the wavelet transform, see [17, 32]. In fact, the S-transform can be seen as an hybrid between the STFT and the wavelet transform. Representation Theory provides a very deep connection among S-transform, STFT and wavelet transform, as they all relate to the representation of the so-called affine Weyl-Heisenberg group, studied in [23]. This connection has been highlighted in the multi-dimensional case by L. Riba in [28], see also [29]. The affine Weyl-Heisenberg group is also the key to represent the α-modulation groups, [7]. 1 2 UBERTINO BATTISTI, MICHELE BERRA Our analysis focuses on the DOST (Discrete Orthonormal Stockwell Transform), a discretization of the S-Transform, first introduced by R. G. Stockwell in [30]. In [4], the DOST transform has been studied from a mathematical point of view. It is shown that the DOST is essentially the decomposition of a periodic signal f P L2pr0, 1sq in a suitable orthonormal basis defined as ď pPZ ď pPZ Dp " tDp,τu2p´1 τ"0 , (1.2) where Dp,τptq " 1? 2p´1 2p´1ÿ η"2p´1 e2πi ηpt´ τ 2p´1q, t P R, p ě 1, τ " 0, . . . , 2p´1 ´ 1, with the convention that D0ptq " 1, see Section 3 for the precise description of the basis func- tions. The DOST basis has a non stationary time-frequency localization, roughly speaking the time localization increases as the frequency increases, while the frequency localization decreases as the frequency increases. Therefore, the basis decomposition of a periodic signal is able to localize high frequencies, for example spikes. The time-frequency localization properties of the DOST basis imply that the coefficients fp,τ " pf, Dp,τqL2pr0,1sq , represent the time-frequency content of the signal f in a certain time-frequency box, which is related to a dyadic decomposition in the frequency domain. Moreover, this decomposition can be seen as a sampling of a generalized S-transform with a particular analyzing window which is essentially a box car window in the frequency domain. The DOST transform gained interest in the applied world after the FFT-fast algorithm discovered by Y. Wang ang J. Orchard, see [33, 34] for the original algorithm and [4] for a slightly different approach which shows that the fast algorithm is essentially a clever application of Plancharel The- orem. The orthogonality property is clearly very useful for several applications, nevertheless it is well known that, in order to describe signals, a certain amount of redundancy can be very effective. Therefore, it is a natural task to look for frames associated to the DOST basis. We follow the idea of the construction of Gabor frames: Gpg, α, βq " tTαkMβngu " ! ) e2πi βnpt´αkqg pt ´ αkq pk,nq PZ2 . If g is a suitable window function, for example a Gaussian, and αβ ă 1 then Gpg, α, βq is a frame, see for example [20]. It is possible to consider the Gabor frame as the frame associated to the standard Fourier basis, which is formed by all modulations with integer frequency, extended by periodicity and then localized using the translation of the analyzing window function g. Inspired by this approach, we consider the system of functions (1.3) βppq pDp,0ptqgptqq Tk ν pp,kq PZ2 , βppq " 2p´1 if p ‰ 0, βp0q " 1, ν ą 0. ) ! For simplicity, here we have not introduced a frequency parameter. The idea of the system in (1.3) is to consider the DOST basis Dp,0 and then localize it with a window function g. The main difference from the standard Gabor system is that the translation parameter k ν βppq is not uniform, but depends on the frequency parameter p. Therefore, (1.3) can be considered as a non stationary Gabor frame, using the terminology introduced in [2]. Inspired by the theory of α-modulation frames, see [14], [15], [16], [19]; in Section 3 we introduce a family of bases of L2pr0, 1sq depending on a parameter α P r0, 1s. If α " 0 we recover the standard Fourier basis, if α " 1 the DOST basis; when α P p0, 1q we show that the basis is associated to a suitable α-partitioning of the frequency domain in the sense of [16]. In Section 4, we prove the main result, see Theorem 4.13; we show that for each α P r0, 1s the localization procedure explained above produces frames of L2pRq, provided the time and frequency EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 3 parameter are small enough. The main tool is a non stationary version of the Walnut representation, see Subsection 4.2. In Section 5, we analyze the higher dimensional case. Restricting to α " 1, we consider a multidimensional partitioning of the frequency domain and the associated frames. This construction is different to the usual extension of the DOST to higher dimensions. In [33, 34], the DOST applied to two dimensional signals (digital images) was essentially the one dimensional DOST applied in the vertical and then in the horizontal direction, therefore it was not a truly bi-dimensional version of the transform. In the paper we use the following normalization of the Fourier Transform 2. Notations ż Fpfqpωq " fpωq " e´2πixωfpxqdx. In the multidimensional case, we do not write explicitly the inner product in Rd, we leave the same notation of the one dimensional case. Set and Mkfpxq " e2πikxfpxq, Tkfpxq " f px ´ kq , ż Let f, g P L2pRdq, we denote the L2-scalar product as k, x P Rd. k, x P Rd. xf, gy " f ¯gdx. We denote SpRdq the Schwartz space. Let α P r0, 1s, we define a partition of N associated to the parameter α. We use an iterative 3. α-bases of L2 pr0, 1sq approach. Let p be a non negative integer. For each α, we define βαp0q " 1, α;pu βα ppq " tiα sα;0 " 1, sα;p " iα;p ` tiα iα;0 " 0, iα;p " sα;p´1, (3.1) α;pu, p ě 1, where txu is the integer part of x. Then, set βα ppq " Iα;p " sα;p ´ iα;p width of the interval Ip. Iα;p " riα;p, sα;pq X N pth interval of the partition of N, 1a βα ppq τ " 0, . . . , βα ppq ´ 1. sα;p´1ÿ e2πi ηpt´ τ βαppqq, η"iα;p We then set for all p P N p,τptq " Bα See Figure 1 for a plot of real and imaginary part of such basis with different values of α and p. 0,0ptq " 1 B1 For p P Z´ we define Notice that if α " 0 then β0ppq " 1, for all p, therefore τ is always zero, hence p,τptq " B´p,τ , Bα τ " 0, . . . , βαp´pq ´ 1. Ť p,0ptq " e2πi pt. B0 p,0 is the ordinary Fourier basis of L2 pr0, 1sq. That is p B0 If α " 1, then i1;0 " 0, s1;0 " 1 and then i1;p " 2p´1, s1;p " 2p, β1ppq " 2p´1 for p ě 1. Therefore 4 UBERTINO BATTISTI, MICHELE BERRA (a) α " 0, p " 1 (b) α " 0, p " 5 (c) α " 0, p " 7 (d) α " 0.5, p " 1 (e) α " 0.5, p " 5 (f) α " 0.5, p " 7 (g) α " 1, p " 1 (h) α " 1, p " 5 (i) α " 1, p " 7 Figure 1. Elements of the DOST bases. Notice that α " 0 is the classical Fourier basis which has no localization in time. As α or p increases we gain time localiza- tion. The black line and the dotted-red one represents the real and imaginary part respectively. 2p´1ÿ η"2p´1 p,τptq " 1? B1 2p´1 e2πi ηpt´ τ 2p´1q, τ " 0, . . . , 2p´1 ´ 1. p,τ B1 p,τ is the so called DOST basis, see [4], [30]. while That is Ť For now on, we restrict to positive integers, all the results hold true also for negative integers via Clearly the above partitioning can be considered on R instead of Z. Let us consider the parti- simple arguments. tioning of the real line. Notice that for α " 0, trivially β0ppq " 1 " η0, @η P I0;p. EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 5 Figure 2. Growth of βα ppq with the respect of the parameter α. If α P p0, 1s then βα ppq " tiα α;pu and for η P Iα;p " riα;p, iα;p ` tiα 1 α;pu α;pu 2α`1 ď tiα p2iα;pqα ď tiα pηqα ď iα α;p iα α;p α;puq we have, for p large enough, " 1. Using Fornasier's notation, see [16], we can write Iα;p -- ηα, η P Iα;p. That is the above partitioning is an α-covering. Remark 3.1. Notice that the α-partitioning introduced above is well defined for all α P r0, 1s. The case α " 1 is not defined as a limit case, as in the usual analysis of α-modulation spaces, see [14, 16]. 1´α . The key point is that we use an iterative scheme, instead of a definition involving the function p In this way, we can get rid of the singularity which arises at α " 1. At α " 1 the growth of βppq with respect to p is exponential while for α P r0, 1q is polynomial, see Figure 2. Using the iterative definition of βppq this fact causes no problems. α Theorem 3.2. The functions p,τ pq , Bα form an orthonormal basis of L2 pr0, 1sq. Proof. Notice that for α " 0 and α " 1 the Theorem holds true in view of well known properties of Fourier basis and of results proven in [4]. p P Z, τ " 0, . . . , βα ppq ´ 1 ››Bα p,τ ›› L2pr0,1sq " 1. The proof follows closely the argument in [4]. Step 1 (cid:32) 6 Since e2πi kt ` UBERTINO BATTISTI, MICHELE BERRA ( kPZ is an orthonormal basis of L2 pr0, 1sq we can write e2πi ηpt´ τ ` L2pr0,1sq " 1 βα ppq sα;p´1ÿ sα;p´1ÿ p,τ , Bα p,τ η1"iα;p ż Bα η ´ η1 δ0 η1"iα;p η"iα;p sα;p´1ÿ sα;p´1ÿ sα;p´1ÿ η"iα;p 1 η"iα;p " 1 βα ppq βα ppq " 1 " 1. ` βαppqqe´2πi η1pt´ τ βαppqqdt Bα p1,τ1 p,τ , Bα Step 2 If p " p1 and τ " τ1 the first step implies the assertion. Moreover, if p ‰ p1 then Iα;p and Iα;p1 are disjoint, thus L2pr0,1sq " δ0 pp ´ p1q δ0 pτ ´ τ1q. ` p,τ , Bα So, we can restrict to the case p " p1. Since obtain Bα ` Bα p,τ , Bα p,τ1 L2pr0,1sq " 1 βα ppq (cid:32) p1,τ1 e2πi kt ( L2pr0,1sq " 0. kPZ is an orthonormal basis of L2 pr0, 1sq, we sα;p´1ÿ βαppq´1ÿ e2πipτ´τ1q η"iα;p βαppq η ` Let us suppose τ ´ τ1 ‰ 0, then we can write L2pr0,1sq " e2πipτ´τ1q iα;p βα ppq e2πipτ´τ1q iα;p p,τ , Bα βαppq Bα p,τ1 βαppq (3.2) " " 0. βα ppq " 1 βα ppq j"0 . βαppq e2πipτ´τ1qpiα;p`jq ´ βαppq´1ÿ e2πipτ´τ1q ¸ ¯j ¸ 1 ´ e2πipτ´τ1q βαppq 1 ´ e2πi pτ´τ1q βαppq βαppq βαppq j"0 1 1 In equation (3.2) we used well known properties of geometric series and the fact that τ´τ1 an integer number and therefore e2πipτ´τ1q{βαppq ‰ 1. Step 3 The functions Bα (cid:32) p,τ generate L2 pr0, 1sq. βαppq´1ď Notice that ( βαppq is not p,τ Ď span Bα e2πi kt kPriα;p,sα;p´1s . τ"0 Ť βαppq´1 τ"0 Therefore, to prove that the functions Bα p,τ are a basis of L2 pr0, 1sq it is sufficient to check that Bα p,τ are linear independent. That is, if βαppq´1ÿ (3.3) τ"0 ατ Bα p,τ " 0 ñ ατ " 0, τ " 0, . . . , βα ppq ´ 1. βαppq´1ÿ τ"0 EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 7 Let us consider the projection of (3.3) into the Fourier basis. We obtain the system of equation ατ e´2πi τ iα;p`j βαppq " 0, j " 0, . . . , βα ppq ´ 1. (3.4) (3.5) We can rewrite the above equation as a linear system e´2πi e´2πi iα;p βαppq iα;p`1 βαppq ... iα;p`βαppq´1 βαppq e´2πi 1 1 ... 1 e´2πi pβαppq´1q iα;p βαppq e´2πi pβαppq´1q iα;p`1 βαppq ... e´2πi pβαppq´1q iα;p`βαppq´1 βαppq . . . . . . . . . . . . ‹‹‹‹‹‚ ‹‹‹‚. 0 ‹‹‹‚" e´2πi piα;p`jq{βαppqβαppq´1 α0 α1 ... 0 ... 0 αβαppq´1 ` The square matrix in (3.5) is a Vandermonde matrix. Since the entries are all distinct the determinant of the matrix is zero and therefore the system in (3.5) has only the null solution, that is in (3.4) and (3.3), ατ mus vanish for all τ " 0, . . . , βα ppq ´ 1. (cid:3) 3.1. Localization properties of the functions Bα p,τ . Now, we investigate the time-frequency localization properties of the functions Bα p,τ . They clearly have compact support in the frequency domain and the support is precisely Iα;p. Therefore, they cannot have compact support in the time domain. Nevertheless, it is possible to determine localization properties in the spirit of Donoho-Stark Theorem, [10]. Property 3.3. For each α P r0, 1s and p, τ , the following inequality holds j"0 ¸ p,τ2dt 2 Bα 2βαppq τ βαppq` 1 βαppq´ 1 τ 2βαppq ż „ ě 0.85.  that is the L2-norm is concentrated in the interval 2βα ppq , „  2βα ppq  if τ " 0, the interval must be considered as an interval in the circle: βα ppq ´ „ βα ppq ` 1 1 τ τ 0, 1 2βα ppq Y 1 ´ 1 2βα ppq , 1 . Proof. The proof is based on Taylor expansion and Gauss summation formula. In [4] the Property is proven for the case α " 1, actually the same proof works also for general α P r0, 1s. (cid:3) 4. DOST-wave packet frame, dimension d " 1. In order to give a more flexible structure to the DOST basis, we generalize it to a redundant and non-orthogonal system and show that this leads to a L2-frame. First, we address the 1-dimensional case in full generality taking into account the phase-space tiling dependent from the parameter α defined in the first part of the paper. 4.1. Frame definition. The idea is to extend by periodicity the DOST basis and then localize them using a suitable window function. Definition 4.1. Consider (4.1) p,kptq " ϕα 1a βαppq ÿ ηPµIα;p t ´ k ν βαppq , pp, kq P Z2, where Iα;p is defined in Section 3. Then, we define the α-DOST system pp, kq P Z2, Dα pµ, ν, ϕq " , see Figure 3 for the plot of the frame elements. We did not plot the case α " 0 which is the standard Gabor frame element. e2πi ηpt´k ν βαppqqϕ ( p,kptq ϕα (cid:32) p,k 8 UBERTINO BATTISTI, MICHELE BERRA Remark 4.2. Notice that the Fourier transform of (4.1) simplifies into pϕα p,kpωq " 1a βαppq e´2πi ωk ν βαppq ÿ ηPµIα;p ϕpω ´ ηq pp, kq P Z2. This kind of system of functions has affinities with the well known family of Non-Stationary Gabor frame, see [3, 11, 12]. The function p;αpωq " Φµ ϕpω ´ ηq ηPµIα;p ÿ (cid:32) mimic a bell function of the set Iα;p as shown in Figure 4. Remark 4.3. Sometimes, when we want to highlight the dependence of the element ϕα frame parameters, we use the notation ϕα;µ,ν Remark 4.4. As we pointed out is Section 3, when α " 0 the DOST basis reduces to the Fourier one. As one may expect, when we analyze the same case for the α-DOST system above, we end up with a Gabor frame. Precisely: p,k . p,kpxq on the ( D0 pµ, ν, ϕq " Gpµ, ν, ϕq " TνkMµpϕ, pp, kq P Z2 . Indeed, recalling that βα ppq " 1 when α " 0, iα;p " p we thus obtain p,kpxq " ϕ0 e2πi ηpx´νkqϕpx ´ νkq " TνkMµpϕ. ÿ η"µp We investigate the condition under which the family Dα pµ, ν, ϕq is a frame of L2 pRq. 4.2. Walnut-like Formula. We retrieve a Walnut-like formula for our DOST system in the spirit of Gabor analysis. This representation is a very useful tool to prove the frame property. Moreover, when α " 0, we recover the Walnut representation for Gabor frames. First, we recall the Poisson formula, [20, Proposition 1.4.2]. Lemma 4.5. Suppose that for some ε ą 0 and C ą 0 we have ϕpxq ď Cp1 ` xq´d´ε and ϕpωq ď Cp1 ` ωq´d´ε. Then ÿ ÿ ϕ n γ (4.2) nPZd for γ ą 0, where d is the dimension of the base space. nPZd γd ϕpx ` γnq " e2πi nx{γ, ¯ ´ We can state the main result of this section. ÿ Lemma 4.6. Let ϕ, ψ P L2pRq, f P C8 c pRq, and xf, ϕα;µ,ν ÿ ÿ Sα;ν;µ ϕ,ψ f ptq " Then ´ ¯ (4.3) p,k p,k FtÑω S α;µ,ν ϕ,ψ f pωq " ν´1 (4.4) where p,k yψα;µ,ν ´ ptq, ¯ f pωq t P R, pp, kq P Z2. Φµ p;α ω ´ k βαppq ν p;αpωq, Ψµ pPZ kPZ Tk βαppq ν ÿ p;αpωq " Ψµ ψpω ´ ηq. ηPµIα;p EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 9 (a) α " 0.5, p " 5 (b) α " 0.5, p " 7 (c) α " 1, p " 5 (d) α " 1, p " 7 Figure 3. Plot of the elements of the α-DOST frame in time, with a Gaussian function as a window and µ " 1{2. As we did for the bases, we observe that the localization increases when α or p grows. The black line and the dotted-red one represents the real and imaginary part respectively. Proof. Take the Fourier transform of the frame operator and obtain ´ FtÑω S α;µ,ν ϕ,ψ f ¯ pωq " " p,k p,k p,kPZ2 x f , pωq ϕα;µ,ν ψα;µ,ν ÿ { y{ ÿ 1a ÿ βα ppq e´2πi k ν x fpq, ÿ 1a x fpq, ÿ ÿ βα ppq e2πi k ν x fpq, e2πi k ν kPZ p,kPZ2 " pPZ " ν´1 ν βα ppq kPZ pPZ βαppqpqΦµ p;αpqy βαppqpqΦµ p;αpqy 1a βα ppq e´2πi k ν 1a ¸ βαppq e2πi k ν βαppq ωΨµ p;αpωq ¸ βαppq ω p;αpωq Ψµ βαppqpqΦµ p;αpqye2πi k ν βαppq ω p;αpωq Ψµ 10 UBERTINO BATTISTI, MICHELE BERRA (a) α " 0.5, p " 5 (b) α " 0.5, p " 7 (c) α " 1 p " 5 (d) α " 1, p " 7 Figure 4. Plot of the elements of the DOST-frame in frequency, with a Gaussian function as a window and µ " 1{2, ν " 1{5. Notice that as p grows, the window translates and approximates a bell function of the set Iα;p. ` p;α, n " k and γ " βαppq Then, (4.2) with ϕ " f Φµ pωq " ν´1 S α;µ,ν ϕ,ϕ f FxÑω f Φµ kPZ ν p;α ´ f ÿ ÿ ÿ pPZ pPZ ÿ ÿ ÿ kPZ " ν´1 ¯ yields βα ppq ω ´ k βα ppq ¯ Φµ p;α ν ν f pωq Φµ p;α ω ´ k ω ´ k ´ p;αpωq Ψµ ω ´ k βαppq ν βαppq ν p;αpωq Ψµ p;αpωq, Ψµ (cid:3) " ν´1 as desired. Tk βαppq ν pPZ kPZ Remark 4.7. We stress that the equalities (4.3) and (4.4) are, at this stage, formal. The frame property we prove in Theorem 4.11 implies that, under suitable assumption on the function ϕ, Sα;µ,ν is a linear operator from L2pRq to itself, therefore (4.3) and (4.4) are equalities in L2pRq. 4.3. Painless frame expansion. In order to have a clear scheme of the frame construction we first consider the painless case. Definition 4.8. Let ϕ P SpRq, such that supp ϕ Ă r´L, Ls and 0 ď ϕ ď 1. Given µ P R, µ ą 0, we say that pϕ, µq is admissible if ϕ,ϕ EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 11 ÿ ηPµIα;p (4.5) p;αpωq " Φµ ϕpω ´ ηq satisfies the following properties 1. 0 ď Φµ 2. For every ω P R tp : Φµ 3. There exists C3 ą 0 such that for every ω P Rd, Φµ p;αpωq ď C1, for some C1 ą 0. p;αpωq ą 0u ď C2. p;αpωq ą C3, for some p P Z. Using this definition we can immediately obtain some important properties of Φµ p;α. (cid:12)(cid:12) ď p2L ` µq βα ppq , Lemma 4.9. Let pϕ, µq be admissible, then ÿ (4.6) (cid:12)(cid:12)supp Φµ p;α @p P Z, 3 ď C 2 Φµ p;αpωq2 ď C 2 1 C2, @ω P R, pPZ where C1, C2, C3 are defined above (cf. Definition 4.8). (4.7) Hence Proof. The first property comes directly from the definition of Φµ p;α (cf. (4.5)) and the fact that supp ϕ Ă r´L, Ls. First, we observe that ω ´ η P supp ϕ if ´L ď ω ´ η ď L; thus, writing η " µpiα;p ` tq, this translates as Since L is fixed and βα ppq ě 1, uniformly in p, we can write (cid:12)(cid:12)supp Φµ p;α µiα;p ´ L ď ω ď µsα;p ` L. (cid:12)(cid:12) ď µpsα;p ´ iα;pq ` 2L " µβα ppq ` 2L. (cid:12)(cid:12)supp Φµ (cid:12)(cid:12) ď pµ ` 2Lqβα ppq , p;α as desired. (cid:12)(cid:12)Φµ Given ω P R, we know that there are at most C2 indexes for which Φµ (cid:12)(cid:12) ď C1, thus the upper bound of (4.7) holds true. p;α On the other hand, for the same ω, there exists at least one p such that Φµ p;αpωq ą 0. Moreover, p;αpωq ě C3. (cid:3) Remark 4.10. An explicit example of an admissible function is the compact version of a Gaussian; precisely ϕpωq " χεpωq 1? , where χε is a bounded smooth function such that e´πω2 2 ω P r´1, 1s , ω P r´1 ´ ε, 1 ` εs . (cid:32) ( p,kpxq ϕα pp,kqP Z2 In [27], similar functions are considered. Theorem 4.11. Consider pϕ, µq being admissible. Then, there exist positive lattice parameters ν, µ such that the α-DOST system (cf. (4.1)) is a frame for L2 pRq. Precisely, there exist A, B ą 0 such that for all f P L2 pRq (4.8) p,kpxqy2 ď B}f}2 2. xf, ϕα A}f}2 2 ď Dα pµ, ν, ϕq " ÿ p,k # χεpωq " 1, 0, Proof. The frame property (4.8) is equivalent to Hence, we evaluate A}f}2 2 ď xSµ,ν ϕ,ϕf, fy ď B}f}2 2. ϕ,ϕ fq, fy, xS α;µ,ν ϕ,ϕ f, fy " xFpS α;µ,ν 12 UBERTINO BATTISTI, MICHELE BERRA where F is the Fourier transform as above. The Walnut representation formula (4.4) yields ` xF S α;µ,ν ϕ,ϕ f , fy " 1 ν " 1 ν ÿ kPZ f Φµ ÿ x pPZ ν βαppq f T k T k p;α, fy ÿ p;αΦµ kPZzt0u ν βαppqΦµ p;αΦµ p;α, fy ν βαppq f T k T k ν βαppqΦµ p;αΦµ p;α, fy. (4.9) We notice that by (4.7) Indeed, We claim that for ν small enough, ÿ ÿ x pPZ x pPZ ` 1 ν ÿ (cid:12)(cid:12)Φµ 1 ν x ÿ pPZ x pPZ ÿ 1 ν kPZzt0u ÿ 1 ν x pPZ f Φµ p;αΦµ p;α, fy -- }f}2 2. (cid:12)(cid:12)2 f , fy -- x f , fy " }f}2 2. p;α ν βαppq f T k T k ν βαppqΦµ p;αΦµ p;α, fy " 0. Recall that supp Φµ p;αpωq ď p2L ` µqβα ppq (cf. (4.6)), thus for ν ă p2L ` µq´1 and k ‰ 0 supp T k ν βαppqΦµ p;α X supp Φµ p;α " H, as desired. (cid:3) 4.4. Conjugate Filter. We can define a way to represent functions even without a canonical dual frame; namely, we construct a conjugate filter for the window function ϕ. This technique is a powerful tool for numerical implementations, see [27]. Set ÿ pPZ p;αpωq Φµ (cid:12)(cid:12)Φµ p;αpωq(cid:12)(cid:12)2 , p;αpωq " ν Ωµ;ν ÿ p;α forms almost a partition of unity p;αpωqΦµ Ωµ;ν p;αpωq " ν. (4.10) then, the product between Ωµ;ν p;α and Φµ (4.11) pPZ We have the following corollary: Corollary 4.12. Consider the functions Ωα;ν p;µ defined in (4.10), then set Then, for any f P L2 pRq (4.12) pΨα p,kpωq " 1a βαppq e´2πi ωk ν ÿ fptq " xf, Ψα p,kyϕα p,kptq. p,k βαppq Ωα;ν p;µpωq EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 13 Proof. We notice immediately that supp Ωµ;ν p;αpωq " supp Φµ p;αpωq. We take the Fourier transform of (4.12), thusÿ p,k " " (4.13) p,kypϕα ÿ p,kpxq xf, Ψα ´ ÿ ÿ pPZ βα ppq´ 1 ÿ kPZ βαppq´1 ÿ kPZ F pPZ ÿ xf, Ψα (4.14) p,kypϕα By (4.11), equation (4.14) becomes ÿ p,k Then, by Poisson formula (cf. (4.2)), (4.13) turns into ¯¸ p;αpωq Φµ 2 e´2πi ωk ν βαppq ¸ βαppq´ 1 2 x f , e´2πi ωk ν ´ ¯ f Ωµ;ν p;α k βαppq Ωµ;ν p;αpωqy e2πi ωk ν βαppq p;αpωq. Φµ ν βα ppq ¯ ÿ ´ kPZ ν´1 f Ωµ;ν p;α pPZ ¸ p;αpωq. Φµ βα ppq ν ω ` k ÿ p,kpxq " 1 fpωq ν pPZ p;αpωqΦµ Ωµ;ν p;αpωq. p,kypϕα xf, Ψα p,kpxq " fpωq. Repeating the same procedure as in the proof of Theorem 4.11, i.e. excluding the terms with k ą 0, we can conclude that, for ν small enough, Finally, applying the inverse Fourier transform, we obtain (4.12). p,k (cid:3) 4.5. DOST frames, general construction. In this section we prove that we can build up a α- DOST frame with milder hypothesis on the window function compared with the ones of the previous section. Theorem 4.13. Consider a function ϕ P L2pRq X L1pRq. As above, set Φµ p;αpωq as (cf. (4.5)) ÿ ηPµIα;p ϕpω ´ ηq. p;αpωq " Φµ ÿ a ď Φµ p;αpωq2 ď b, @ ω P R, ϕpωq ď CN p1 ` ωqN , @ω P R Suppose that (4.15) for some N ą 2. for suitable constants a, b ą 0. Moreover, assume that the window ϕ satisfies p Then, there exists A, B ą 0 and ν0 ą 0, such that for all f P L2pRq for 0 ă ν ă ν0. A}f}2 2 ď xSµ,ν g,g f, fy ď B}f}2 2, 14 UBERTINO BATTISTI, MICHELE BERRA 4.6. Preparatory Lemmata. We need some result concerning the decay of the elements Φµ outside the sets Iα;p. Lemma 4.14. Let ϕ P L2 pRq X L1pRq such that: ϕpωq ď CN p;α p1 ` ωqN , with N ą 2. Then, (4.16) Proof. From the definition of Φµ Hence, CN,µ p1 ` dpω, µIα;pqqN´1 . ÿ Φµ p;αpωq ď ÿ ηPIα;p p;α we obtain: ϕpω ´ µηq ď ÿ (cid:12)(cid:12)Φµ p;αpωq(cid:12)(cid:12) ď p;αpωq(cid:12)(cid:12) ď p1 ` dpω, µIα;pqq´N (cid:12)(cid:12)Φµ p;αpωq(cid:12)(cid:12) ď (cid:12)(cid:12)Φµ ÿ ηPIα;p CN p1 ` ω ´ µηqN . ηPIα;p CN p1 ` dpω, µIα;pqqN . p1 ` ω ´ µηqN ÿ CNp1 ` ω ´ µηq´N ď CN pµj ` 1q´N , jPZ If ω P µIα;p, then dpω, µIα;pq " 0. Thus (4.17) ηPIα;p which is clearly convergent for N ą 2. If ω R µIα;p, then dpω, µIα;pq " δ and (cid:12)(cid:12)Φµ p;αpωq(cid:12)(cid:12) ď 2CN p1 ` δqN ď 2CN p1 ` δqN " 2CN p1 ` δqN " 2CN p1 ` δqN " 2CN p1 ` δqN for N ą 2, which completes the proof. Lemma 4.15. Let ω P R and N ą 1. Thenÿ βαppqÿ ż `8 ζ"0 ż `8 1 ` 1 ` 1 ` p1 ` δq 1 ` p1 ` δq µ 0 0 µ p1 ` δqN p1 ` µζ ` δqN ¸ p1 ` µx ` δqN dx p1 ` δqN ż `8 1 1`δ xqN dx p1 ` µ p1 ` xqN dx ď p1 ` δqN´1 CN,µ C 1 0 (cid:3) 1 p1 ` dpω, µIα;pqqN ď C pPZ and the constant is independent on ω. Proof. For any ω P R, there exists only one p such that ω P µIα; ¯p. Then, for j ě 2, dpω, µIα;p`jq ě j´1ÿ k"1 βα pp ` kq ě pj ´ 1q, EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 15 since βα ppq ě 1. The same argument works for j ă ´2. Obviously Hence, we can bound our sum asÿ 1ÿ j"´1 1 p1 ` dpω, µIα;p`jqqN ď 3. ÿ 1 p1 ` dpω, µIα;pqqN ď pPZ CN jN jPZzt0u and the constants do not depend on the particular choice of ω, as desired. (cid:3) ÿ ÿ p;α as in Theorem 4.13 above, then βαppq ν ω ´ k lim νÑ0 p;α kPZzt0u ››››Φµ ›››› 8 p;αpωq Φµ " 0. Lemma 4.16. Let Φµ pPZ Proof. Let us define for all p P Z p;αpωq " χµIα;ppωqΦµ p;αpωq 1Φµ and We can write ÿ (4.18) (4.19) (4.20) (4.21) pPZ ď p;αpωq " Φµ p;αpωq ´ 1Φµ p;αpωq. 2Φµ ›››› kPZzt0u ÿ ÿ ÿ ÿ ÿ pPZ pPZ pPZ ` ` ` p;α p;α ››››Φµ βαppq ››››1Φµ ω ´ k ÿ ν ››››2Φµ ω ´ k ÿ ››››1Φµ ÿ ››››2Φµ ÿ ω ´ k ω ´ k ω ´ k kPZzt0u kPZzt0u p;α p;α p;α kPZzt0u pPZ kPZzt0u 1Φµ p;αpωq Φµ βαppq ν βαppq ν βαppq ν βαppq ν ›››› 8 p;αpωq 8 p;α pωq 1Φµ p;αpωq 2Φµ p;αpωq 2Φµ ›››› ›››› ›››› 8 8 . 8 For now on, let us suppose νµ ă 1; then we notice that the term in (4.18) is identically zero for each p P Z, since the supports are disjoint. Clearly this is not restrictive since we are considering the limit for ν Ñ 0. We remark that: ´ µ ω ´ k βα ppq ν , µIα;p ě k ν βα ppq , @ω P µIα;p, d (4.22) with k ´ νµ ą 0, k P Zzt0u. 16 UBERTINO BATTISTI, MICHELE BERRA We then analyze the term in (4.19). For each ω there exists a unique p " p such that ω P µIα;p, hence for each ωÿ ÿ ÿ (cid:12)(cid:12)(cid:12)(cid:12)2Φα p;µ kPZzt0u ω ´ k βαppq ν p;µ pωq 1Φα (cid:12)(cid:12)(cid:12)(cid:12) pPZ kPZzt0u ď CN,µ ď CN,µ (cid:12)(cid:12)(cid:12)(cid:12) " p;α pωq ¯¯ N´1 , µIα;p ¯ N´1 (cid:12)(cid:12)(cid:12)(cid:12)2Φµ ÿ ÿ p;α ´ ´ ÿ kPZzt0u kPZzt0u kPZzt0u ν βαppq ν 1Φµ ω ´ k ´ CN,µ ω ´ k βαppq 1 ` d ¯ ´ CN k βαppq ν ´ µ 1 ` 1´N k ´ µ ν . (4.23) ď CN,µ CN We made use of the decay proven in Lemma 4.14 and of (4.22). Since N ´ 1 ą 1 and the inequality (4.23) does not depend on ω, we obtain βαppq ν ´ µ ››››2Φµ ď CN,µ CN p;α pωq ω ´ k ÿ ÿ ÿ k ν ›››› 1´N 1Φµ p;α pPZ kPZzt0u ÿ 8 kPZzt0u " CN,µ CN νN´1 1 pk ´ νµq1´N . kPZzt0u (4.24) Since the sum is convergent uniformly with respect to ν, the term in (4.24) goes to 0 as ν Ñ 0 with the rate νN´1. In order to consider the term in (4.20), notice that for all ω P R, there exists a unique ¯p such that ω P µIα; ¯p. Moreover, arguing as above, for all p P Zzt¯pu there exist a unique kp P Zzt0u such that ω ´ kp βαppq ν ÿ P Iα;p. Hence, we can write (4.20) as βαppq ν p;α pPZzt ¯pu ›››› p;αpωq 2Φµ . 8 ››››1Φµ kp ν ω ´ kp ´ µ Finally, by (4.16), we can write ÿ ››››1Φµ p;α ω ´ kp dpω, µIα;pq ě βαppq ν p;αpωq 2Φµ ›››› 8 (4.25) By (4.22) pPZzt ¯pu (4.26) if ω ´ kp βα ppq ν P µIα;p. βα ppq , ÿ ÿ ÿ pPZzt ¯pu pPZzt ¯pu pPZzt ¯pu ď ď ď ď νε pPZzt ¯pu C 2 N,µ p1 ` dpω, µIα;pqqN´1 C 2 N,µ p1 ` dpω, µIα;pqqN´1´ε ÿ p1 ` dpω, µIα;pqqN´1´ε C 2 N,µ C p1 ` dpω, µIα;pqqN´1´ε . 1 ´ p1 ` dpω, µIα;pqqε ´ ¯ kp 1 ` ν ´ µ 1 ¯ βα ppq ε The sum in (4.26) converges uniformly with respect to ω by Lemma 4.15, therefore the whole sum goes to zero as ν does. EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 17 The last term to be taken into account is (4.21), which includes the "tails" of our window function. Fix p P Z, by definition if either ω P µIα;p or ω ´ k βαppq ν p;α 2Φµ " k : 0 ă d ω ´ k βαppq ν βα ppq ω ´ k P µIα;p, then p;αpωq " 0. 2Φµ , µIα;p ν ă 1 2ν βα ppq * . Define the set p " F ω Notice that only a finite number of element k belong to this set; precisely, when µν ă 1, then So, for all ω, we can split the term in (4.21) as follows: ÿ ÿ pPZ kPZzt0u (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;α p ď 2, F ω ω ´ k βαppq ν p;αpωq 2Φµ uniformly in ω and p. (cid:12)(cid:12)(cid:12)(cid:12) ď ÿ ÿ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ ÿ ÿ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ pPZ ` kRF ω p;α p kPF ω pPZ p ω ´ k ω ´ k p;α βαppq ν βαppq ν p;αpωq 2Φµ p;αpωq 2Φµ (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) . We show that both terms are bounded by a constant times νε with ε ą small enough independent on ω. We notice that if k R F ω p , then βα ppq , µIα;p d ω ´ k ω ´ j d ν where k is the closest index to k inside F ω into j P Zzt0u such that ÿ ÿ pPZ kRF ω p (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;α ω ´ k βαppq ν p;αpωq 2Φµ From the discussion above, and the estimate (4.16) it follows that βα ppq , ν ě k ´ k ν ν , µIα;p βα ppq . (cid:12)(cid:12)(cid:12)(cid:12) ď p ; hence, we can rearrange the (infinite) indexes k P Zzt0u βα ppq ÿ ÿ ÿ ě j (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ´ ¯ CN,µ 1 ` j ν βα ppq ´ ¯ CN,µ p1 ` dpω, µIα;pqqN´1 j ν N´1 p1 ` dpω, µIα;pqqN´1 ÿ ÿ ÿ jPZzt0u jPZzt0u (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) N´1 N´1 CN,µ CN,µ pPZ pPZ ď ď pPZ jPZzt0u CN,µ ν j CN,µ p1 ` dpω, µIα;pqqN´1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) The latter term is summable in j, Lemma 4.15 implies the summability in p as well, therefore the whole term goes to zero as νN´1. The last part follows from the observation below: if there exists k P F ω βα ppq . βα ppq ω ´ k p , then , µIα;p ν ă 1 2ν Since µ is fixed and ν Ñ 0, we can assume µν ă 1; thus, there exists δ such that µν " 1 ´ δ. We split our analysis in two cases, first we consider βα ppq 0 ă d ω ´ k d , µIα;p ν ď δ 2ν βα ppq . 18 UBERTINO BATTISTI, MICHELE BERRA Let s P µIα;p such that dpω, sq " dpω, µIα;pq then by triangular inequality dpω, µIα;pq " dpω, sq ě d ě k βα ppq ´ ν ´ d ω ´ k βα ppq ν , s βα ppq ω, ω ´ k ` µ ν βα ppq . δ 2ν Hence dpω, µIα;pq ą k βα ppq ν If k ě 1, it is clear that ´ βα ppq ν 1 ´ δ 2 " βα ppq ν k ´ 1 ` δ 2 dpω, µIα;pq ą δ 2ν βα ppq ¯ ε ď CN,µνε p1 ` dpω, µIα;pqqN´1´ε . 1 βαppq ν 2 1 ` δ , µIα;p ą δ 2ν βα ppq Then, 2Φµ p;αpωq ď CN,µ p1 ` dpω, µIα;pqqN´1´ε ´ On the other hand, when d ω ´ k βα ppq ν we have ω ´ k (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;α βαppq ν (cid:12)(cid:12)(cid:12)(cid:12) ď ď ´ ` ´ CN,µ ω ´ k βαppq ¯¯ N´1 , µIα;p 1 ` d CN,µ 2ν βppq 1 ` δ ν N´1 (cid:12)(cid:12)(cid:12)(cid:12) ď F ω p ÿ pPZ ω ´ k βαppq ν p;αpωq 2Φµ CN νε p1 ` dpω, µIα;pqqN´1´ε which goes to zero as νN´1. Using the inequalities, it is clear that in both cases ÿ ÿ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;α (4.27) pPZ kPF ω p since the sum is uniformly bounded with respect to ω, the above inequalities imply that (4.27) goes to zero as νε. The quantity p1 ` dpω, µIα;pqqN´1´ε is summable if N ´ 1 ´ ε ą 1. This is granted by the fact that N ą 2 and that we can chose ε small enough. (cid:3) 4.7. Proof of the main result. xS α;µ,ν ϕ,ϕ f, fy " xF " 1 ν ď 1 ν ď 1 ν " 1 ν Similarly, ` ÿ ÿ ÿ ν ν 2 p;α p;α Φµ kPZ kPZzt0u Tk βαppq S α;µ,ν ϕ,ϕ f p;αpωq Φµ βαppq ν x pPZ ` 1 ÿ ν f pωq ω ´ k ω ´ k βαppq ω ´ k ÿ ÿ ν , fy " 1 x Tk βαppq ν pPZ p;α pωq , f pωqy p;α pωq Φµ f pωq Φµ ´ ¯ ÿ ÿ f pωq x (cid:12)(cid:12)Φµ p;α pωq(cid:12)(cid:12)2 f pωq , f pωqy pPZ x ÿ ÿ pPZ ` 1 ÿ ››Φµ ν pPZ kPZzt0u p;α pωq ÿ ÿ $&%ÿ ››Φµ kPZzt0u p;α pωq $&% inf ÿ ››››Φµ ››2 ››››Φµ 8 }f}2 p;αpωq ››››Φµ Φµ ››2 ω ´ k 8 ` ››››Φµ ÿ ÿ p;α pωq(cid:12)(cid:12)2 ´ (cid:12)(cid:12)Φµ ÿ p;α pωq(cid:12)(cid:12)2 ď b. (cid:12)(cid:12)Φµ pPZ ` 1 ν kPZzt0u kPZzt0u a ď ωPR pPZ pPZ pPZ pPZ pPZ pPZ p;α p;α p;α ν xS α;µ,ν ϕ,ϕ f, fy ě 1 By hypothesis, we know that βαppq ν p;αpωq, f pωqy Φµ ´ ¯ f pωq ´ f pωq ν xTk βαppq ›››Tk βαppq ν 8 ›››› ›››› 8 βαppq ν p;αpωq Φµ ω ´ k βαppq ν p;αpωq Φµ , f pωqy ›››› 8 2 2 ››› f ››› ¯››› ,.-}f}2 ,.-}f}2 ›››› 2 2 8 EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 19 Proof. As we did in the painless case, we can represent the frame operator as (cf. (4.9)) ´ ¯ Φµ p;α βαppq ν p;αpωq, f pωqy Φµ ω ´ k Lemma 4.16 implies that lim νÑ0 βαppq ν Hence, there exists ν0 ą 0 such that for all 0 ă ν ă ν0 βαppq ν ω ´ k kPZzt0u pPZ p;α p;α Then, the action of the frame operator can be bounded as follows p;αpωq Φµ " 0. ›››› 8 ›››› 8 2 ď a ÿ ÿ ››››Φµ ››››Φµ ÿ ÿ ÿ ››Φµ ››2 p;α pωq 8 ` pPZ ÿ b ` a }f}2 (cid:12)(cid:12)Φµ p;α pωq(cid:12)(cid:12)2 ´ $&%ÿ ` $&% inf kPZzt0u pPZ 2 2 ωPR pPZ (cid:12)(cid:12)xS α;µ,ν ϕ,ϕ f, fy(cid:12)(cid:12) ď 1 (cid:12)(cid:12)xS α;µ,ν ϕ,ϕ f, fy(cid:12)(cid:12) ě 1 ď 1 ν ν ν p;αpωq Φµ ω ´ k ÿ p;α ››››Φµ ››››Φµ ÿ p;α ω ´ k pPZ kPZzt0u ÿ pPZ kPZzt0u βαppq ν ›››› 8 ,.-}f}2 ,.- ě a ›››› 2 8 p;αpωq Φµ ω ´ k βαppq ν p;αpωq Φµ 2ν }f}2 2 , as desired. (cid:3) Φµ p;αpωq2 " 1 2 e´πpµpiα; ¯p`tq´ηq2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 p ηPµIα;p (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ÿ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)βαp ¯pq´1ÿ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)βαp ¯pq´1ÿ j"0 20 UBERTINO BATTISTI, MICHELE BERRA 4.8. Existence of DOST frames. We show that the Gaussian satisfies the hypotheses of the main Theorem 4.13. Theorem 4.17. Consider ϕ " 1? p;αpωq as in (4.5). Then, ÿ e´πx2 a ď 2 and set Φµ Φµ p;αpωq2 ď b,@ ω P R., (4.28) for suitable constants a, b ą 0. Moreover, ϕpωq ď p CN @ω P R, p1 ` ωqN , (4.29) for any N ě 0. Proof. The polynomial decay claimed in (4.29) is trivial, since ϕ is a Schwartz function. For the lower bound in (4.28), we argue as follows: for any ω P R there exists only one p such that ω P µIα;p. For the sake of simplicity, we assume p ą 0, the negative case follows with the same argument. Therefore, ÿ Φµ p;αpωq2 ě Φµ p;αpωq2. Since ω P µIα, ¯p, then ω " µpiα; ¯p ` tq, 0 ď t ď βαp¯pq. Hence (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 , e´πµ2pt´jq2 because of the positivity of the Gaussian. The maximum value is reached when t´ j is small. Our construction implies that there exists j such that t ´ j ď 1, thus e´2πµ2 , sup j"0,...,βαp ¯pq´1 " 1 2 e´πpµpiα; ¯p`tq´µpiα; ¯p`jqq2 j"0 ě 1 2 " 1 2 e´πµ2pt´jq2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)e´πµ2(cid:12)(cid:12)(cid:12)2 " 1 Φµ p;αpωq2 ě 1 ÿ ÿ p;αpωq2 ď Φµ ÿ We rewrite the sum above as follows:ÿ 2 2 p p which is independent on p and ω, as desired. Due to the fact that the Gaussian is positive, it follows that also Φµ p;αpωq2. Φµ Finally, since ϕ belongs to the Wiener Space, we can write p j p;αpωq " Φµ ÿ ϕpω ´ µjq. ¸ 2 ϕpω ´ µjq ď ` 1 }ϕ}W 2 ă 8, 1 µ ÿ pPZ Φµ p;αpωq2 ď esssup j where }}W is the Wiener norm. We have used a well known property of Wiener space, see e.g. (cid:3) [20][Lemma 6.1.2]. p;α is positive as well. Hence EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 21 5. Higher Dimensions We consider here the case α " 1 and an arbitrary dimension. We define a (parabolic) phase space tiling and, for suitable window functions, we provide a frame of L2pRdq. We follow the ideas of wave atoms proposed in [8, 9] and subsequently adapted to the Gaussian case by [27]. For the sake of simplicity, we enlighten the notation used before by suppressing the parameter α. As for the dimension d " 1, we begin with the painless case using a smooth and compactly supported window function. Moreover, we can define an explicit conjugate frame that leads to a reconstruction formula. Then we generalize the construction. 5.1. Phase space partition. Define the Cartesian coronae Cp as follows: C0 " r´1, 1sd, Cp " " ω " pω1, . . . , ωdq P Rd : max 1ďiďd * ωs P rβ ppq , β pp ` 1qq p ě 1. , Each corona is further partitioned in (open) boxes of side β ppq " 2p´1, precisely Xp,(cid:96) " dź ` " β ppq (cid:96)s, β ppq p(cid:96)s ` 1q ¯ , s"1 ls ` 1 R C0, i.e. the centers are outside the inner corona. where (cid:96)s " ´2,´1, 0, 1 and maxs"1...d The indexes (cid:96)s label every possible box inside the corona. It can be easily checked that the number č of such boxes (or multi-indexes) is 2dp2d ´ 1q, for every p ě 1. We also define, according to Section 3, 2 p " Xp,(cid:96) I (cid:96) Zd, (5.1) (5.2) see Figure 5. We generalize now the 1-dimensional DOST system (cf. (4.1)). Definition 5.1. Given the set I (cid:96) p (cf. (5.1)), consider e2πi ηpx´k ν βppqqϕ ϕγpxq " β ppq´ d 2 , x ´ k ν β ppq ÿ ηPµI (cid:96) p where γ " pp, k, (cid:96)q P pN, Zd, Jq, J contains the admissible indexes (cid:96) described above. Then, we define the (multi-dimensional) DOST-system M D pµ, ν, ϕq " tϕγpxquγ . p in frequency. See Figure 6 as an example of frame element both in time and frequency. Figure 7 shows how the elements are localized in the set I (cid:96) 5.2. Painless frame expansion. As for the case of dimension d " 1, we start with compactly supported window functions. We adapt the definition of admissible window. Definition 5.2. Let ϕ P SpRdq such that supp ϕ Ă r´L, Lsd, for some µ ą 0 and 0 ď ϕ ď 1. We say that pϕ, µq is admissible if, given (5.3) p;(cid:96)pωq " Φµ ϕpω ´ ηq ÿ ηPµI (cid:96) p it satisfies the following properties p;(cid:96)pωq ď C1, for some C1 ą 0. 1. 0 ď Φµ 2. For every ω P Rd tp, (cid:96) : Φµ 3. There exists C3 ą 0 such that for every ω P Rd, Φµ p;(cid:96)pωq ą 0u ď C2. p;(cid:96)pωq ą C3, for some pp, (cid:96)q P N J. 22 UBERTINO BATTISTI, MICHELE BERRA β pp ` 1q β ppq I (cid:96) p . . . . . . . . . . . . . . . . Xp,(cid:96) β ppq β pp ` 1q Figure 5. 2D phase space partitioning, and the set I (cid:96) p. Using this definition we can immediately obtain some important properties of Φµ p;(cid:96). Lemma 5.3. Let pϕ, µq be admissible, then supp Φµ p;(cid:96) Diam (5.4) ´ ¯ 3 ď C 2 (5.5) where C1, C2, C3 are defined above (cf. 4.8) and Diam denotes the maximum distance between points of the set, i.e. the diameter. ` ¯ Rd pxq " ´ Sν;µ ϕ,ψf ¯ @p P N, 1 C2, ÿ ď CL,µβ ppq , p,(cid:96) Φµ p;(cid:96)pωq2 ď C 2 ` Rd , set ÿ . For each f P L2 ÿ ` xf, ϕγyψγpxq, x P Rd, γ P ÿ γ N, Zd, J . Φµ p;(cid:96) ω ´ k βαppq ν ´ ¯ f pωq ÿ p,;(cid:96)pωq " Ψµ ψpω ´ ηq. ηPµI (cid:96) p S α;µ,ν ϕ,ψ f pωq " 1 νd Tk βαppq ν p,(cid:96) k Lemma 5.4. Let ϕ, ψ in L2 ´ FxÑω Then (5.6) where p,;(cid:96)pωq, Ψµ The proofs of these lemmata are analogous to Lemmata 4.6-4.9. Theorem 5.5. Consider pϕ, µq being admissible. Then there exists ν ą 0 such that the DOST- system is a frame for L2 (5.7) Rd . Precisely, there exist A, B ą 0 such that for all f P L2 Rd M D pµ, ν, ϕq ÿ A}f}2 2 ď xf, ϕγpxqy2 ď B}f}2 2. ` ` The proof is analogous to the one made in dimension d " 1. γ EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 23 Corollary 5.6. Consider the functions Ωµ;ν 2 e´2πi ωk ν 2p´1 Ωµ;ν γ " pp, k, (cid:96)q P pN, Zd, Jq, ω P Rd. ÿ pPZ p;(cid:96)pωq Φµ p;(cid:96)pωq(cid:12)(cid:12)(cid:12)2 , (cid:12)(cid:12)(cid:12)Φµ p;(cid:96) pωq " νd Ωµ;ν ÿ p;(cid:96) pωqΦµ Ωµ;ν p;(cid:96)pωq " νd. p,(cid:96) p;(cid:96) defined above, set p;(cid:96) pωq, ÿ xf, Ψγyϕγpxq. fpxq " γ 5.3. Conjugate Filter. Set (5.8) then (5.9) pΨγ pωq " β ppq´ d ` Rd Then, for any f P L2 (5.10) ization constant.ÿ x f ,pΨγypϕγpxq ÿ ÿ β ppq´ d ÿ " γ γ " 1 νd p,(cid:96) kPZd Proof. We briefly review the proof of Corollary 4.12. Nothing changes here a part from the normal- β ppq´ d p;(cid:96) pωqy ¯ p;(cid:96) pωqy βppq Ωµ;ν βppq Ωµ;ν 2 x fpωq, e´2πi ωk ν ´ β ppq´d x fpωq, e´2πi ωk ν 2 e´2πi ωk ν βppq Φµ p;(cid:96)pωq ¸ e´2πi ωk ν βppq p;(cid:96)pωq. Φµ ÿ fpωqν´d Ωµ;ν p;(cid:96) Φµ p;(cid:96)pωq " fpωq, Then, by Poisson formula (cf. (4.2)) and (5.9), supposing ν0 small enough, we have for all 0 ă ν ă ν0 as desired. p,(cid:96) (cid:3) 5.4. DOST frames, general construction. We state and prove that we can build up a DOST ÿ frame with similar hypothesis to the ones given in Section 4. Theorem 5.7. Consider a function ϕ P L2pRdq X L1 p;(cid:96)pωq (cf. (5.3)) . Let Φµ ` ϕpω ´ ηq. Rd p;(cid:96)pωq " Φµ ÿ ηPµI (cid:96) p Φµ p;(cid:96)pωq2 ď b, @ω P R, Suppose that µ ą 0 is chosen so that a ď for suitable constants a, b ą 0. Moreover, assume that the window ϕ satisfies p ϕpωq ď CN p1 ` ωqN , ` Rd Then there exists A, B ą 0 and ν0 ą 0, such that for all f P L2 g,g f, fy ď B}f}2 2 ď xSµ,ν A}f}2 2 (5.11) for some N ą 2d. for some 0 ă ν ă ν0. 24 UBERTINO BATTISTI, MICHELE BERRA 5.5. Preparatory Lemmata. We recall the same result proved in dimension d " 1. Lemma 5.8. Let ϕ P L2 X L1pRdq such that (5.11) holds true. Then Rd ` (5.12) p;(cid:96)pωq ď Φµ Lemma 5.9. Let ω P Rd and N ą d. Thenÿ ` 1 ` d ››››Φµ and the constant is independent on ω. Lemma 5.10. Let Φµ p;(cid:96) as before then ÿ ÿ p,(cid:96) p;(cid:96) lim νÑ0 kPZdzt0u p,(cid:96) ` 1 ` d ` 1 ω, µI (cid:96) p ` CN ω, µI (cid:96) p N´d . N ď C ω ´ k β ppq ν ›››› 8 p;(cid:96)pωq Φµ " 0. Proof. As we did in Lemma 4.16, set 1Φµ p;(cid:96)pωq " χµI (cid:96) ppωqΦµ p;(cid:96)pωq and Then, we split the norm as follows: ÿ ÿ kPZdzt0u kPZdzt0u ›››› p;(cid:96)pωq ´ 1Φµ p;(cid:96)pωq. p;(cid:96) 2Φµ p;(cid:96)pωq " Φµ ››››Φµ β ppq ››››1Φµ ω ´ k ÿ ν ››››2Φµ ω ´ k ÿ ››››1Φµ ω ´ k ÿ ››››2Φµ ω ´ k ÿ ω ´ k kPZdzt0u kPZdzt0u p;(cid:96) p;(cid:96) p;(cid:96) p;(cid:96) p;(cid:96)pωq Φµ β ppq ν β ppq ν β ppq ν β ppq ν ›››› 8 p;(cid:96)pωq 1Φµ 8 p;(cid:96) pωq 1Φµ p;(cid:96)pωq 2Φµ p;(cid:96)pωq 2Φµ ›››› ›››› ›››› 8 8 . 8 p,(cid:96) ÿ ÿ ÿ ÿ p,(cid:96) p,(cid:96) ` ` ` kPZdzt0u p,(cid:96) p,(cid:96) " (5.13) (5.14) (5.15) (5.16) Notice that if νµ ă 1, then the term in (5.13) is identically zero for each p P N, since the supports are disjoint. In order to analyze the term in (5.14), notice that for each ω P Rd there exist unique ¯p, ¯l such that ω P I ¯(cid:96) ¯p. Notice that we have ω ´ kβ ppq (5.17) For each ω P Rd, by (5.17) we have d ν p;(cid:96) pωq 2Φµ p;(cid:96) ω ´ k (cid:12)(cid:12)(cid:12)(cid:12)1Φµ 1 ν (cid:12)(cid:12)(cid:12)(cid:12) ď ą k ` , µI (cid:96) p β ppq ν dµ ´ ? ` C 2 N 1 ` d ω, µI (cid:96) p β ppq , N´d @ω P I (cid:96) p. dµ k d´N . ´ ? 1 ν ÿ ÿ EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 Our hypotheses grant that N ´ d ą d, then the above remarks implies that β p¯pq ω ´ k ν ´ ? (cid:12)(cid:12)(cid:12)(cid:12) ď (cid:12)(cid:12)(cid:12)(cid:12)1Φµ p;(cid:96) pωq 2Φµ β ppq ν ω ´ k kPZdzt0u ď C dµ ¯p;¯(cid:96) p,(cid:96) p;(cid:96) k (cid:12)(cid:12)(cid:12)(cid:12)1Φµ ¯p;¯(cid:96) pωq 2Φµ ÿ kPZdzt0u 1 ν d´N 25 (cid:12)(cid:12)(cid:12)(cid:12) where the latter tends to 0 as ν Ñ 0 and the constants are uniformly bounded with the respect of ω. Hence the term in (5.14) has a limit vanishing as ν approaches zero. ¯p, ¯l such that ω P I ¯(cid:96) w ´ k βppq The term in (5.15), goes to zero as ν goes to zero as well. We notice that for each ω, there exist , there exists kp,l such that ¯p. For the same reason for each p, l P ( ¯(cid:96) ν P I (cid:96) p. Therefore,ÿ Then, applying equation (5.17), (5.12) and Lemma 5.9 as in (4.26) we can conclude that the term in (5.18) goes to zero as ν goes to zero independently on ω. ››››1Φµ ››››1Φµ p;(cid:96) p;(cid:96) ÿ ÿ kPZdzt0u p‰ ¯p,(cid:96)‰¯(cid:96) ω ´ k ω ´ kp,l (5.18) p,(cid:96) " " We consider now the term in (5.16), set k : 0 ă d F ω p,(cid:96) ď 2d, One can split the term in (5.16) as follows: p,(cid:96) " F ω then ÿ ω ´ k β ppq ν p;(cid:96)pωq 2Φµ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;(cid:96) kPZdzt0u ω ´ kβ ppq ν (cid:12)(cid:12)(cid:12)(cid:12) ď ÿ p,(cid:96) kRF ω ` Then, for each ω P Rd,ÿ ÿ ÿ This yields ÿ kPF ω and p,(cid:96) kRF ω p,(cid:96) ω ´ k p;(cid:96) (cid:12)(cid:12)(cid:12)(cid:12)2Φµ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ ω ´ k (cid:12)(cid:12)(cid:12)(cid:12)2Φµ p;(cid:96) p;(cid:96) β ppq ν β ppq ν β ppq ν ω ´ k p;(cid:96)pωq 2Φµ p;(cid:96)pωq 2Φµ p;(cid:96)pωq 2Φµ pPZ kPZdzt0u d´N , dµ ´ ? 1 ν ď C ` Nzt¯pu , Jz p;(cid:96)pωq 2Φµ (cid:32) ›››› 8 p;(cid:96)pωq 2Φµ ›››› β ppq ν β ppq ν . 8 * β ppq kPF ω p,(cid:96) p;(cid:96) p;(cid:96) , µI (cid:96) p ă 1 2ν (cid:12)(cid:12)(cid:12)(cid:12)2Φµ ω ´ k ÿ (cid:12)(cid:12)(cid:12)(cid:12)2Φµ (cid:12)(cid:12)(cid:12)(cid:12) ď ` (cid:12)(cid:12)(cid:12)(cid:12) ď ` (cid:12)(cid:12)(cid:12)(cid:12)ď 1 ` d ` ` CN ν´ε ω, µI (cid:96) p 1 ` d ÿ ` ` ` ` 1 ` d 1 ` d pPZ νd´N CN ω, µI (cid:96) p N´d uniformly in ω. N´d´ε . CN νε ω, µI (cid:96) p N´d´ε N´d´ε (cid:12)(cid:12)(cid:12)(cid:12) 2Φµ p;(cid:96)pωq β ppq ν ω ´ k β ppq ν p;(cid:96)pωq 2Φµ (cid:12)(cid:12)(cid:12)(cid:12) . is summable which goes to zero as ν goes to zero as desired. We stress that if N ´ d´ ε ą d which is granted by Lemma 5.9. Since the bounds are all unifrom with the respect (cid:3) to ω, we can conclude that the terms in (5.16) goes to zero as ν does. ω, µI (cid:96) p 26 UBERTINO BATTISTI, MICHELE BERRA 5.6. Proof of the main result. Proof. From the d " 1 case, we get ν sup ωPRd xS α;µ,ν " " ϕ,ϕ f, fy ď 1 ϕ,ϕ f, fy ě 1 xS α;µ,ν ÿ p;(cid:96) pωq(cid:12)(cid:12)(cid:12)2 (cid:12)(cid:12)(cid:12)Φµ ››››Φµ ÿ ÿ H0 " inf ωPRd p,(cid:96) ν kPZdzt0u p,(cid:96) p;(cid:96) kě1 " 1 H ν ν * * }f}2 2 }f}2 2 H0pωq ` H ν H0pωq ´ H ν kě1pωq kě1pωq ω ´ k β ppq ν ›››› 8 . p;(cid:96)pωq Φµ with By hypothesis, we know that Lemma 5.10 implies that limνÑ0 H ν Then, for all ν P p0, ν0q, the action of the frame operator can be bounded as follows @0 ă ν ă ν0. a ď H0pωq ď b. ( kě1 " 0. Hence, there exists ν0 ą 0 such that kě1 ă a (cid:32) H ν 2 , " }H0}8 ` H ν kě1 H0 ´ H ν ` * }f}2 2 ď 1 b ` a 2ν }f}2 2 ě a }f}2 2 . }f}2 kě1 ν ν 2 2 inf ωPRd (cid:12)(cid:12)xS α;µ,ν ϕ,ϕ f, fy(cid:12)(cid:12) ď 1 (cid:12)(cid:12)xS α;µ,ν ϕ,ϕ f, fy(cid:12)(cid:12) ě 1 ν (cid:3) Remark 5.11. As we did in dimension d " 1 (cf. Theorem 4.17), we can show that the normalized Gaussian fulfills the hypothesis of Theorem 5.7. 6. Conclusions We constructed a frequency-adapted frame which covers Gabor and Stockwell-related frames. Our setting includes also general α-phase-space partitioning. This approach appears natural to describe α-Modulation spaces and this will be subject of a future work. In [4], the author prove that the DOST basis is able to diagonalize the S-transform with a suit- able window function which is essentially a boxcar window in the frequency domain and that the evaluation of the DOST-coefficients turns out to be the evaluation of the S-transform with this particular window in a suitable lattice. The natural question is if the α-DOST bases introduced in this paper have the same property, clearly not with respect to the S-transform, but in relation to [16] another transform, similar to the flexible Gabor-wavelet transform (or α-transform), see e.g. for the the definition. The n-dimensional case considered in Section 5 is restricted to the case α " 1, hence a suitable phase-space partitioning is yet to be defined for α P r0, 1q and will be part of our future studies. This issue has been already analyzed in the two dimensional case by N. Morten in [25] using the theory of Brushlets. From a computational stand point, we aim to implement and compare our results with existent methods. We are interested in testing in various applications such as medical and seismic imaging and also general image processing. As pointed out in the introduction, we remark that our approach consider the n-dimensional case in a peculiar way: instead of applying the one dimensional DOST in each direction sequentially, we provide a native n-dimensional setting. This approach is similar to the Wavepackets and Curvelets one, see [8, 27]. A natural question arises: is the density of our frames comparable with the Gabor case? For instance, is it true that if the volume of the lattice is strictly lower than 1, the Gaussian leads to a EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 27 (a) Time view, α " 1, p " 3 (b) Frequency view, α " 1, p " 3 (c) Time view, α " 1, p " 5 (d) Frequency view, α " 1, p " 5 Figure 6. Time and frequency outlook of two window functions. We observe heavy decay in time while in frequency we localize around a certain frequency. The window ϕ is in both cases a normalized Gaussian. frame? And is this condition independent on α? Acknowledgments. We thank Elena Cordero, Fabio Nicola, Luigi Riba and Anita Tabacco for the useful discussions on the subject. We are also graceful to Maarten V. de Hoop and Man Wah Wong for the opportunity to talk about our projects in international meetings. The authors were partially supported by the Gruppo Nazionale per l'Analisi Matematica, la Proba- bilit`a e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). The first author is partially supported by the Research Project FIR (Futuro in Ricerca) 2013 Geometrical and qualitative aspects of PDE's. [1] Special issue: Time-frequency applications, in The Leading Edge, vol. 34, 2015. [2] P. Balazs, M. Dorfler, F. Jaillet, N. Holighaus, and G. Velasco, Theory, implementation and applications of nonstationary Gabor frames, J. Comput. Appl. Math., 236 (2011), pp. 1481 -- 1496. References 28 UBERTINO BATTISTI, MICHELE BERRA Figure 7. Two window functions for α " 1 and p " 5, 7. We see how the normal- ization β ppq´1 and the width of I (cid:96) p affects the shape of the windows. The window ϕ is, in both cases, a normalized Gaussian. [3] , Theory, implementation and applications of nonstationary Gabor frames, J. Comput. Appl. Math., 236 (2011), pp. 1481 -- 1496. [4] U. Battisti and L. Riba, Window-dependent bases for efficient representations of the Stockwell transform, Applied and Computational Harmonic Analysis, (2015), p. to appear. [5] M. Biswal and P. K. Dash, Detection and characterization of multiple power quality disturbances with a fast S-transform and decision tree based classifier, Digit. Signal Process., 23 (2013), pp. 1071 -- 1083. [6] R. S. Chora´s, Time-Frequency Analysis of Image Based on Stockwell Transform, in Image Processing and Communications Challenges 5, vol. 233 of Advances in Intelligent Systems and Computing, Springer International Publishing, 2014, pp. 91 -- 97. [7] S. Dahlke, M. Fornasier, H. Rauhut, G. Steidl, and G. Teschke, Generalized coorbit theory, Banach frames, and the relation to α-modulation spaces, Proc. Lond. Math. Soc. (3), 96 (2008), pp. 464 -- 506. [8] L. Demanet and L. Ying, Wave atoms and sparsity of oscillatory patterns, Appl. Comput. Harmon. Anal., 23 (2007), pp. 368 -- 387. [9] [10] D. L. Donoho and P. B. Stark, Uncertainty principles and signal recovery, SIAM J. Appl. Math., 49 (1989), , Wave atoms and time upscaling of wave equations, Numer. Math., 113 (2009), pp. 1 -- 71. pp. 906 -- 931. [11] M. Dorfler and E. Matusiak, Nonstationary Gabor frames -- existence and construction, Int. J. Wavelets Multiresolut. Inf. Process., 12 (2014), pp. 1450032, 18. [12] , Nonstationary Gabor frames - approximately dual frames and reconstruction errors, Adv. Comput. Math., 41 (2015), pp. 293 -- 316. [13] S. Drabycz, R. Stockwell, and J. R. Mitchell, Image Texture Characterization Using the Discrete Or- thonormal S-Transform, 22 (2009), pp. 696 -- 708. [14] H. G. Feichtinger and M. Fornasier, Flexible Gabor-wavelet atomic decompositions for L2-sobolev spaces, Ann. Mat. Pura Appl., 185 (2006), pp. 105 -- 131. [15] H. G. Feichtinger and P. Grobner, Banach spaces of distributions defined by decomposition methods. I, Math. Nachr., 123 (1985), pp. 97 -- 120. [16] M. Fornasier, Banach frames for α-modulation spaces, Applied and Computational Harmonic Analysis, 22 (2007), pp. 157 -- 175. [17] P. C. Gibson, M. P. Lamoureux, and G. F. Margrave, Letter to the editor: Stockwell and wavelet transforms, J. Fourier Anal. Appl., 12 (2006), pp. 713 -- 721. [18] B. G. Goodyear, H. Zhu, R. A. Brown, and J. R. Mithcell, Removal of phase artifacts from fMRI using a Stockwell transform filter improves brain activity detection, Magnetic Resonance in Medicine, 51 (2004), pp. 16 -- 21. [19] P. Grobner, Banachraeume glatter Funktionen und Zerlegungsmethoden, ProQuest LLC, Ann Arbor, MI, 1992. Thesis (Dr.natw.) -- Technische Universitaet Wien (Austria). [20] K. Grochenig, Foundations of time-frequency analysis, Applied and Numerical Harmonic Analysis, Birkhauser Boston, Inc., Boston, MA, 2001. EXPLICIT CONSTRUCTION OF NON-STATIONARY FRAMES FOR L2 29 [21] Q. Guo, S. Molahajloo, and M. W. Wong, Phases of modified Stockwell transforms and instantaneous frequencies, J. Math. Phys., 51 (2010), pp. 052101, 11. [22] M. Jaya Bharata Reddy, R. Krishnan Raghupathy, K. P. Venkatesh, and D. K. Mohanta, Power quality analysis using Discrete Orthogonal S-transform (DOST), Digit. Signal Process., 23 (2013), pp. 616 -- 626. [23] C. Kalisa and B. Torr´esani, N-dimensional affine Weyl-Heisenberg wavelets, Annales de l'Institut Henri Poincar´e (A) Physique Th´eorique, 59 (1993), pp. 201 -- 236. [24] J. Ladan, An Analysis of Stockwell Transforms, with Applications to Image Processing, Thesis University of Waterloo, Ontario, Canada, (2014). [25] N. Morten, Orthonormal bases for α-modulation spaces, Collect. Math., 2 (2010), pp. 173 -- 190. [26] N. Ortigosa, O. Cano, G. Ayala, A. Galbis, and C. Fern´andez, Atrial fibrillation subtypes classification using the General Fourier-family Transform, Med. Eng. Phys., 36 (2014), pp. 554 -- 560. [27] J. Qian and L. Ying, Fast multiscale Gaussian wavepacket transforms and multiscale Gaussian beams for the wave equation, Multiscale Model. Simul., 8 (2010), pp. 1803 -- 1837. [28] L. Riba, Multi-Dimensional Stockwell Transforms and Applications, PhD thesis, Universit`a degli Studi di Torino, Italy, 2014. [29] L. Riba and M. W. Wong, Continuous inversion formulas for multi-dimensional modified Stockwell transforms, Integral Transforms Spec. Funct., 26 (2015), pp. 9 -- 19. [30] R. Stockwell, A basis for efficient representation of the S-transform, Digital Signal Processing, 17 (2007), pp. 371 -- 393. [31] R. G. Stockwell, L. Mansinha, and R. P. Lowe, Localization of the complex spectrum: the S transform, IEEE Transactions on Signal Processing, 44 (1996), pp. 998 -- 1001. [32] S. Ventosa, C. Simon, M. Schimmel, J. J. Danobeitia, and A. M`anuel, The S-transform from a wavelet point of view, IEEE Trans. Signal Process., 56 (2008), pp. 2771 -- 2780. [33] Y. Wang, Efficient Stockwell Transform with Applications to Image Processing, PhD thesis, University of Wa- terloo, Ontario, Canada, 2011. [34] Y. Wang and J. Orchard, Fast discrete orthonormal Stockwell transform, SIAM Journal on Scientific Com- puting, 31 (2009), pp. 4000 -- 4012. [35] M. W. Wong and H. Zhu, A characterization of Stockwell spectra, in Modern trends in pseudo-differential operators, vol. 172 of Oper. Theory Adv. Appl., Birkhauser, Basel, 2007, pp. 251 -- 257. [36] H. Zhu, B. Goodyear, M. Lauzon, R. Brown, G. Mayer, A. Law, L. Mansinha, and J. Mitchell, A new local multiscale Fourier analysis for medical imaging, Med. Phys., 30 (2003), pp. 1134 -- 41. Dipartimento di Scienze Matematiche, Politecnico di Torino, corso Duca degli Abruzzi 24, 10129 Torino, Italy E-mail address: [email protected] Universit`a di Torino, Dipartimento di Matematica, via Carlo Alberto 10, 10123 Torino, Italy E-mail address: [email protected]
1011.6607
1
1011
2010-11-30T16:55:19
Operator Aczel inequality
[ "math.FA", "math.OA" ]
We establish several operator versions of the classical Aczel inequality. One of operator versions deals with the weighted operator geometric mean and another is related to the positive sesquilinear forms. Some applications including the unital positive linear maps on $C^*$-algebras and the unitarily invariant norms on matrices are presented.
math.FA
math
OPERATOR ACZEL INEQUALITY MOHAMMAD SAL MOSLEHIAN Abstract. We establish several operator versions of the classical Aczel in- equality. One of operator versions deals with the weighted operator geometric mean and another is related to the positive sesquilinear forms. Some applica- tions including the unital positive linear maps on C ∗-algebras and the unitarily invariant norms on matrices are presented. 1. Introduction Let B(H ) denote the algebra of all bounded linear operators acting on a com- plex Hilbert space (H , h·, ·i) and I is the identity operator. In the case where dim H = n, we identify B(H ) with the full matrix algebra Mn(C) of all n × n matrices with entries in the complex field C. An operator A ∈ B(H ) is called positive (positive-semidefinite for matrices) if hAξ, ξi ≥ 0 holds for every ξ ∈ H and then we write A ≥ 0. For A, B ∈ B(H), we say A ≤ B if B − A ≥ 0. Let f be a continuous real valued function defined on an interval J. The function f is called operator decreasing if B ≤ A implies f (A) ≤ f (B) for all A, B with spectra in J. A function f is said to be operator concave on J if λf (A) + (1 − λ)f (B) ≤ f (λA + (1 − λ)B) for all A, B ∈ B(H ) with spectra in J and all λ ∈ [0, 1]. By a C ∗-algebra we mean a closed ∗-subalgebra A of B(H ) for some Hilbert space H . Any finite dimensional C ∗-algebra is isometrically ∗-isomorphic to a direct sum of finitely many full matrix algebras. A C ∗-algebra is called unital if it has an identity. A map taking the identity to identity is called unital. A map Φ : A → B between C ∗-algebras is called positive if it takes positive operators to positive ones, in particular, a positive linear map from A into C is called a positive linear functional. A map Φ : A → B is called 2-positive if the map Φ2 : M2(A ) → M2(B) defined by Φ2([aij]) = [Φ(aij)] is positive, where M2(A ) 2010 Mathematics Subject Classification. Primary 47A63; Secondary 15A60, 46L05, 26D15. Key words and phrases. Aczel inequality; operator geometric mean; operator concave; oper- ator decreasing; positive linear functional; C ∗-algebra. 1 2 M.S. MOSLEHIAN is the C ∗-algebra all 2 × 2 matrices with entries in A . An operator A is called a contraction if kAk ≤ 1. We refer the reader to [11] for undefined notions on operator theory and to [8] for more information on operator inequalities. In 1956, Acz´el [3] proved that if ai, bi (1 ≤ i ≤ n) are positive real numbers such that a2 i > 0 or b2 i=2 b2 i > 0, then i=2 a2 1 −Pn a1b1 − 1 −Pn ≥ a2 1 − aibi!2 n Xi=2 a2 i! b2 1 − n Xi=2 b2 i! . n Xi=2 The Acz´el inequality has some applications in mathematical analysis and in the theory of functional equations in non-Euclidean geometry. During the last decades several interesting generalization of this significant inequality were ob- tained. Popoviciu [12] extended Acz´el's inequality by showing that aibi!p n Xi=2 a1b1 − 1 −Pn 1 − ≥ ap 1 −Pn ap i! bp 1 − n Xi=2 bp i! , n Xi=2 if p ≥ 1 and ap i > 0. Acz´el's inequality and Popovi- ciu's inequality was sharpened by Wu [15], see also [14]. A variant of Acz´el's i > 0 or bp i=2 ap i=2 bp inequality in inner product spaces was given by Dragomir [5] by establishing that if a, b are real numbers and x, y are vectors of an inner product space such that a2 − kxk2 > 0 or b2 − kyk2 > 0, then (a2 − kxk2)(b2 − kyk2) ≤ (ab − Rehx, yi)2, see also [6]. Cho, Mati´c and Pecari´c [4] generalized Acz´el's inequality for linear isotonic functionals and convex functions. Several Acz´el type inequalities involv- ing norms in Banach spaces were presented by Mercer [9]. Also, Sun [13] gave an Acz´el -- Chebyshev type inequality for positive linear functionals. To find operator versions of Hua's inequality (see [10]), Fujii [7, Theorem 3] obtained an Acz´el op- erator inequality by showing that if Φ : A → B is a contractive 2-positive unital linear map between unital C ∗-algebras, A, B ∈ A are contraction and Φ(B ∗A) is normal with the polar decomposition Φ(B ∗A) = UΦ(B ∗A), then 1 − Φ(B ∗A) ≥ (1 − Φ(A∗A))♯U ∗(1 − Φ(B ∗B))U . In this paper we establish several operator versions of the classical Aczel inequal- ity. One of operator versions deals with the weighted operator geometric mean A♯tB := A1/2(A−1/2BA−1/2)tA1/2 (t ∈ [0, 1]) and another is related to the positive sesquilinear forms. Some applications including the unital positive linear maps on C ∗-algebras and the unitarily invariant norms on matrices are presented. Recall that a unitarily invariant norm · has the property U XV = X, where OPERATOR ACZEL INEQUALITY 3 U and V are unitaries and X ∈ Mn(C). For more information on the theory of the unitarily invariant norms the reader is referred to [1]. 2. Operator Aczel inequality via geometric mean We start this section with a lemma about a parameterized operator power mean mt satisfying AmtB ≤ (1−t)A+tB for any two positive invertible operators A, B. The power means A♯r,tB := A1/2(1−t+t(A−1/2BA−1/2)r)1/rA1/2 (r ∈ [−1, 1]\{0}) and A♯0,tB := A♯tB (t ∈ [0, 1]) are such parameterized operator power means. Clearly if AB = BA, then A♯tB = A1−tBt; see [8, Chapter V]. Lemma 2.1. Suppose that mt is a parameterized operator power mean not greater than the weighted arithmetic mean. If J is an interval of (0, ∞) and f : J → (0, ∞) is operator decreasing and operator concave on J and A, B ∈ B(H ) are positive invertible operators with spectra contained in J , then f (AmtB) ≥ f (A)mtf (B) (2.1) Proof. It follows from AmtB ≤ (1 − t)A + tB that f (AmtB) ≥ f (1 − t)A + tB) (since f is operator decreasing) ≥ (1 − t)f (A) + tf (B) (since f is operator concave) (2.2) ≥ f (A)mtf (B) (by the property of mt) . (cid:3) Theorem 2.2. Let J be an interval of (0, ∞), let f : J → (0, ∞) be operator decreasing and operator concave on J , 1 q = 1, p, q > 1 and let A, B ∈ B(H ) p + 1 be positive invertible operators with spectra contained in J . Then f (Ap♯1/qBq) ≥ f (Ap)♯1/qf (Bq) hf (Ap♯1/qBq)ξ, ξi ≥ hf (Ap)ξ, ξi 1 p hf (Bq)ξ, ξi 1 q . (2.3) (2.4) for any vector ξ ∈ H . Proof. Lemma 2.1 yields inequality (2.4). Let ξ ∈ H be an arbitrary vector. It follows from (2.2) that hf (Ap♯1/qBq)ξ, ξi ≥ 1 p hf (Ap)ξ, ξi + 1 q hf (Bq)ξ, ξi ≥ hf (Ap)ξ, ξi 1 p hf (Bq)ξ, ξi 1 q (weighted arithmetic-geometric mean inequality). 4 M.S. MOSLEHIAN (cid:3) Remark 2.3. The Holder -- McCarthy inequality asserts that if C ∈ B(H ) is a positive operator, then hC rξ, ξi ≤ hCξ, ξir for all 0 < r < 1 and all unit vectors ξ ∈ H ; cf. [8, Theorem 1.4]. It follows from (2.4) that hf (Ap♯1/qBq)ξ, ξi ≥ hf (Ap) 1 p ξ, ξihf (Bq) 1 q ξ, ξi . Thus kf (Ap♯1/qBq)1/2ξk ≥ kf (Ap) 1 2p ξk kf (Bq) 1 2q ξk . Corollary 2.4. Let 1 p + 1 q = 1, p, q > 1 and A, B ∈ B(H ) be commuting positive invertible operators with spectra contained in (0, 1). Then 1 − kABξk2 ≥ h(cid:0)1 − kAp/2ξk2(cid:1) for any unit vector ξ ∈ H . 1 p (cid:0)1 − kBq/2k2(cid:1) 1 q . Proof. Apply Theorem 2.2 to the function f (t) = 1 − t on (0, 1) and note that Ap♯1/qBq = AB. (cid:3) Corollary 2.5. Let J be an interval of (0, ∞), let f : J → (0, ∞) be operator decreasing and operator concave on J and A, B ∈ B(H ) be positive invertible commuting operators with spectra contained in J . Then f (AB) ≥ (f (Ap))1/p (f (Bq))1/q . Corollary 2.6. If f is a decreasing concave function on an interval J and ai, bi (1 ≤ i ≤ n) are positive numbers in J , then n Xi=1 f (aibi) ≥ n Xi=1 f (bq i )!1/q . f (ap i )!1/p n Xi=1 Apply (2.4) to the positive operators A(x1, · · · , xn) = (a1x1, · · · , anxn) and B(x1, · · · , xn) = (b1x1, · · · , bnxn) acting on the Hilbert space H = Cn and ξ = (1, 1, · · · , 1). 3. Aczel's inequality via positive linear functionals In this section we present a new version of Aczel's inequality through positive linear functionals. The first result is a generalization of the main result of [5]. The proof differs from that the main result of [5]. It is an Acz´el type inequality for sesquilinear forms. OPERATOR ACZEL INEQUALITY 5 Theorem 3.1. Let φ(., .) be a positive sesquilinear form on a linear space X , let x, y ∈ X such that φ(x, x) ≤ M 2 M1, M2 and let L : C → R be a function fulfilling L(z) ≤ z for all z ∈ C. Then 2 for some positive numbers 1 or φ(y, y) ≤ M 2 (cid:0)M1M2 − L(φ(x, y))(cid:1)2 ≥(cid:0)M 2 1 − φ(x, x)(cid:1)(cid:0)M 2 2 − φ(y, y)(cid:1) . Proof. We may assume that both φ(x, x) ≤ M 2 the Cauchy -- Schwarz inequality implies that φ(x, y) ≤ M1M2. We have 1 and φ(y, y) ≤ M 2 2 hold. Then (M1M2 − L(φ(x, y)))2 ≥ (M1M2 − φ(x, y))2 φ(x, x) + φ(y, y) ≥(cid:16)M1M2 −pφ(x, x)φ(y, y)(cid:17)2 (cid:19)2 ≥(cid:18)M1M2 − ≥(cid:18) M 2 ≥(cid:0)M 2 1 − φ(x, x)(cid:1)(cid:0)M 2 2 − φ(y, y)(cid:1) 2 − φ(x, x) − φ(y, y) 1 + M 2 2 2 as desired. (Cauchy -- Schwarz inequality) (arithmetic-geometric mean ineq.) (cid:19)2 (arithmetic-geometric mean ineq.) (cid:3) The next result is a consequence of Theorem 3.1, but we present a different proof for it. Theorem 3.2. Suppose that Φ : A → B is a unital positive linear map between unital C ∗-algebras of operators acting on a Hilbert space H , A, B ∈ A such that Φ(A∗A) or Φ(B ∗B) is a contraction. Then (cid:0)1 − hΦ(B ∗A)ξ, ξi(cid:1)2 for all unit vectors ξ ∈ H . ≥(cid:0)1 − hΦ(A∗A)ξ, ξi(cid:1)(cid:0)1 − hΦ(B ∗B)ξ, ξi(cid:1) Proof. Without loss of generality, assume that Φ(A∗A) is a contraction. Let ξ ∈ H be a unit vector. Hence hΦ(A∗A)ξ, ξi ≤ 1. Let us consider the quadratic polynomial P (t) = (1 − hΦ(A∗A)ξ, ξi)t2 − 2(1 − hΦ(B ∗A)ξ, ξi)t + (1 − hΦ(B ∗B)ξ, ξi) , where t ∈ R. It follows from the Cauchy -- Schwarz inequality applied to the sesquilinear form hA, Bi = hΦ(B ∗A)ξ, ξi that P (1) = −hΦ(A∗A)ξ, ξi + 2hΦ(B ∗A)ξ, ξi − hΦ(B ∗B)ξ, ξi ≤ 0 . 6 M.S. MOSLEHIAN Clearly, limt→∞ P (t) = ∞. Hence the equation P (t) = 0 has a root in R. Thus (cid:0)1 − hΦ(B ∗A)ξ, ξi(cid:1)2 as desired. −(cid:0)1 − hΦ(A∗A)ξ, ξi(cid:1)(cid:0)1 − hΦ(B ∗B)ξ, ξi(cid:1) ≥ 0 (cid:3) The next result is immediately deduced from Theorem 3.1. Corollary 3.3. Let ψ be a positive linear functional on a C ∗-algebra A , let A, B ∈ A such that ψ(A∗A) ≤ M 2 M1, M2 and let L : C → R be a function fulfilling L(z) ≤ z for all z ∈ C. Then 2 for some positive numbers 1 or ψ(B ∗B) ≤ M 2 Corollary 3.4 (Acz´el's Inequality). If ai, bi (1 ≤ i ≤ n) are positive numbers ≥(cid:0)M 2 1 − ψ(A∗A)(cid:1)(cid:0)M 2 2 − ψ(B ∗B)(cid:1) . (3.1) such that Pn (cid:0)M1M2 − L(ψ(B ∗A))(cid:1)2 i < 1 or Pn aibi!2 1 − Xi=1 i=2 a2 n i=2 b2 i < 1, then n n a2 ≥ 1 − i! 1 − Xi=1 Apply Corollary 3.3 to the n × n matrices A =   Xi=1 a1 0 b2 i! . . . . 0 an   and B = , positive linear functional tr(·) on Mn(C) and L(z) = z. b1 0 . . . 0 bn     If we assume that A and B are normal contractions of a unital C ∗-algebra A , AB = BA and consider the positive sesquilinear form φ(C, D) = ψ(D∗C) (C, D ∈ A ), where ψ is a pure state (or, equivalently, a non-zero complex homomorphism) on the commutative C ∗-algebra generated by three elements A, B and the identity I of A , then we get from (3.1) that in which Re denotes the real part. Hence (cid:0)1 − Re(ψ(B ∗A))(cid:1)2 ψ(cid:0)1 − Re(B ∗A)(cid:1)2 (cid:0)1 − Re(B ∗A)(cid:1)2 ≥(cid:0)1 − ψ(A∗A)(cid:1)(cid:0)1 − ψ(B ∗B)(cid:1) , ≥ ψ(cid:0)(1 − A∗A)(1 − B ∗B)(cid:1) , ≥ (1 − A∗A)(1 − B ∗B) . whence OPERATOR ACZEL INEQUALITY 7 The same assertion is valid with the imaginary part Im instead of Re. We proved therefore the following result. Corollary 3.5. Let A, B be commuting normal contractions of a unital C ∗- algebra A . Then and (cid:0)1 − Re(B ∗A)(cid:1)2 (cid:0)1 − Im(B ∗A)(cid:1)2 ≥ (1 − A∗A)(1 − B ∗B) ≥ (1 − A∗A)(1 − B ∗B) . Corollary 3.6. Let ψ be a positive linear functional on Mn(C), let A, B ∈ Mn(C) such that ψ(A) ≤ M 2 2 for some positive numbers M1, M2. Then 1 or ψ(B) ≤ M 2 (cid:0)M1M2 − ψ(A♯B))(cid:1)2 ≥(cid:0)M 2 1 − ψ(A)(cid:1)(cid:0)M 2 2 − ψ(B)(cid:1) . 1 and ψ(B) ≤ M 2 Proof. We may assume that ψ(A) ≤ M 2 2 . The positive linear functional ψ on Mn(C) can be characterized by ψ(C) = hCZ, Zi, where Z ≥ 0 and h·, ·i denotes the canonical inner product on Mn(C) defined by hX, Y i = tr(Y ∗X). It follows from (3.1) with L(z) = z and elements A1/2 and (A−1/2BA−1/2)1/2A1/2 that (cid:0)M1M2 − h(A−1/2BA−1/2)1/2A1/2Z, A1/2Zi(cid:1)2 1 − h(A−1/2BA−1/2)1/2A1/2Z, (A−1/2BA−1/2)1/2A1/2Zi(cid:1) ≥(cid:0)M 2 ×(cid:0)M 2 2 − hA1/2Z, A1/2Zi(cid:1) , or equivalently (cid:0)M1M2 − hA1/2(A−1/2BA−1/2)1/2A1/2Z, Zi(cid:1)2 1 − hA1/2(A−1/2BA−1/2)1/2(A−1/2BA−1/2)1/2A1/2Z, Zi(cid:1) ≥(cid:0)M 2 2 − hA1/2A1/2Z, Zi(cid:1) , 1 − ψ(A)(cid:1)(cid:0)M 2 ×(cid:0)M 2 ≥(cid:0)M 2 (cid:0)M1M2 − ψ(A♯B))(cid:1)2 2 − ψ(B)(cid:1) . whence Note that 0 ≤ ψ(A♯B) = ψ(A♯B). (cid:3) Finally, we present an Acz´el type inequality involving unitarily invariant norms. Note that AXB ≤ kAk X kBk (3.2) 8 M.S. MOSLEHIAN for all X, A, B. The arithmetic-geometric mean inequality states that A∗XB ≤ 1 2 AA∗X + XBB ∗ . (3.3) Proposition 3.7. Let · be a unitarily invariant norm on Mn(C) and let X, A, B ∈ Mn(C) such that kAk2 X ≤ 1 or kBk2 X ≤ 1. Then (1 − A∗XB)2 ≥ (1 − kAk2 X)(1 − kBk2 X) Proof. ≥ (1 − A∗XB)2 ≥(cid:18)1 − (1 − AA∗X) + (1 − XBB ∗) 1 2 AA∗X + XBB ∗ (cid:19)2 2 (by (3.3)) (triangle Ineq.) ≥ (1 − AA∗X)(1 − XBB ∗) (arithmetic-geometric mean ineq.) ≥ (1 − kAk2 X)(1 − kBk2 X) (by (3.2)) . (cid:3) References 1. R. Bhatia, Matrix Analysis, Springer-Verlag, New York, 1997. 2. R. Bhatia and C. Davis, More matrix forms of the arithmeticgeometric mean inequality, SIAM J. Matrix Anal. 14 (1993) 132-136. 3. J. Acz´el, Some general methods in the theory of functional equations in one variable, New applications of functional equations. (Russian) Uspehi Mat. Nauk (N.S.) 11 (1956), no. 3(69), 3 -- 68. 4. Y.J. Cho, M. Mati´c and J.E. Pecari´c, Improvements of some inequalities of Acz´el's type, J. Math. Anal. Appl. 259 (2001), no. 1, 226 -- 240. 5. S.S. Dragomir, A generalization of J. Aczl's inequality in inner product spaces, Acta Math. Hungar. 65 (1994), no. 2, 141 -- 148. 6. S.S. Dragomir and B. Mond, Some inequalities of Aczl type for gramians in inner product spaces, Nonlinear Funct. Anal. Appl. 6 (2001), pp. 411-424. 7. J.I. Fujii, Operator inequalities for Schwarz and Hua, Sci. Math. 2 (1999), no. 3, 263 -- 268. 8. T. Furuta, J. Mi´ci´c Hot, J.E. Pecari´c and Y. Seo, Mond -- Pecari´c Method in Operator In- equalities, Element, Zagreb, 2005. 9. A.M. Mercer, Extensions of popovicius inequality using a general method, J. Inequal. Pure Appl. Math. 4 (1) (2003) Article 11. 10. M.S. Moslehian, Operator extensions of Hua's inequality, Linear Algebra Appl. 430 (2009), no. 4, 1131 -- 1139. 11. J.G. Murphy, C ∗-Algebras and Operator Theory , Academic Press, San Diego, 1990. 12. T. Popoviciu, On an inequality, Gaz. Mat. Fiz. Ser. A 11 (64) (1959), 451-461. OPERATOR ACZEL INEQUALITY 9 13. X.H. Sun, Acz´el -- Chebyshev type inequality for positive linear functions, J. Math. Anal. Appl. 245 (2000), 393-403. 14. S. Wu and L. Debnath, A new generalization of Acz´el's inequality and its applications to an improvement of Bellman's inequality, Appl. Math. Lett. 21 (2008), no. 6, 588 -- 593. 15. S. Wu, Some improvements of Acz´el's inequality and Popoviciu's inequality, Comput. Math. Appl. 56 (2008), no. 5, 1196 -- 1205. Department of Pure Mathematics, Center of Excellence in Analysis on Al- gebraic Structures (CEAAS), Ferdowsi University of Mashhad, P. O. Box 1159, Mashhad 91775, Iran. E-mail address: [email protected] and [email protected] URL: http://profsite.um.ac.ir/~moslehian/
1702.06049
1
1702
2017-02-20T16:19:07
Some natural subspaces and quotient spaces of $L^1$
[ "math.FA" ]
We show that the space $\text{Lip}_0(\mathbb R^n)$ is the dual space of $L^{1}({\mathbb R}^{n}; {\mathbb R}^{n})/N$ where $N$ is the subspace of $L^{1}({\mathbb R}^{n}; {\mathbb R}^{n})$ consisting of vector fields whose divergence vanishes. We prove that although the quotient space $L^{1}({\mathbb R}^{n}; {\mathbb R}^{n})/N$ is weakly sequentially complete, the subspace $N$ is not nicely placed - in other words, its unit ball is not closed for the topology $\tau_m$ of local convergence in measure. We prove that if $\Omega$ is a bounded open star-shaped subset of $\mathbb {R}^n$ and $X$ is a closed subspace of $L^1(\Omega)$ consisting of continuous functions, then the unit ball of $X$ is compact for the compact-open topology on $\Omega$. It follows in particular that such spaces $X$, when they have Grothendieck's approximation property, have unconditional finite-dimensional decompositions and are isomorphic to weak*-closed subspaces of $l^1$. Numerous examples are provided where such results apply.
math.FA
math
SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 GILLES GODEFROY AND NICOLAS LERNER Abstract. We show that the space Lip0(Rn) is the dual space of L1(Rn; Rn)/N where N is the subspace of L1(Rn; Rn) consisting of vector fields whose divergence vanishes. We prove that although the quotient space L1(Rn; Rn)/N is weakly sequentially complete, the subspace N is not nicely placed - in other words, its unit ball is not closed for the topology τm of local convergence in measure. We prove that if Ω is a bounded open star-shaped subset of Rn and X is a closed subspace of L1(Ω) consisting of continuous functions, then the unit ball of X is compact for the compact-open topology on Ω. It follows in particular that such spaces X, when they have Grothendieck's approximation property, have unconditional finite-dimensional decompositions and are isomorphic to weak*-closed subspaces of l1. Numerous examples are provided where such results apply. 1. Introduction Among the wealth of important discoveries due to Uffe Haagerup, one can single out what is now universally called Haagerup's approximation property, a funda- mental concept in operator algebras and their various applications. The present work investigates approximation properties on a much lesser scale, and the tools we use are familiar to every functional analyst: among them, dilation operators on star-shaped domains and Grothendieck's approximation property. Our purpose is to analyse some natural subspaces (and quotient spaces) of L1. We are therefore outside the reflexive world, where the lack of compactness can hurt some proofs and where some natural operators become unbounded. This leads us to weaken the topologies, thus to enter the realm of non-locally convex spaces and to use the topol- ogy τm of convergence in measure. Such tools will allow us to provide satisfactory results on subspaces of L1 which satisfy quite weak assumptions: for instance, we show (Corollary 10) that if Ω is a star-shaped bounded open subset of Rn, if X is a closed subspace of L1(Ω) consisting of continuous functions and stable under the di- lation operators (Tρ), and if X has Grothendieck's approximation property, then X is isomorphic to a weak-star closed subspace of l1. Hence such a space has a "some- what discrete" structure. It turns out that these assumptions are satisfied by many classical spaces. Moreover these spaces X have unconditional finite dimensional de- compositions. We therefore apply a rule of thumb which has been discovered by Nigel Kalton and some of his co-authors: homogeneity of a Banach space X implies unconditionality on X. We now outline the content of this note. Let Ω be an open subset of Rn, equipped with the Lebesgue measure denoted m. A closed subspace X of L1(Ω) is called nicely placed if its unit ball is closed for the topology τm of local convergence in measure (see Chapter IV in [17] or [20]). It is known that the quotient space L1/X is L-complemented in its bidual (and thus weakly sequentially complete) when X is Date: February 21, 2017. 1 2 GILLES GODEFROY AND NICOLAS LERNER nicely placed, and the same conclusion holds when we consider integrable functions with values in a finite-dimensional normed space (see e.g. p.200 in [17]), in particular integrable vector fields on Rn. Our first result is somewhat negative: we show that the free space F (Rn) over Rn is isometric to the quotient of the space (L1(Rn))n = L1(Rn; Rn) of integrable vector fields on Rn by the space N of divergence-free vector fields, and we show that although F (Rn) shares many properties of spaces which are L-complemented in their bidual, the space N is not nicely placed. This discards a natural conjecture, but leads to several questions. In the second (independent) part of our paper, we show that if Ω is star-shaped and bounded, homogeneous subspaces X of L1(Ω) consisting of continuous functions are very special examples of nicely placed subspaces: their unit ball BX is actually τm-compact locally convex. This bears strong consequences on the structure of such spaces. 2. Divergence-free vector fields and the space F (Rn) We first provide a representation result for the predual of the space of Lipschitz functions on the space Rn. We recall the usual notation (1) Lip0(Rn) = {f : Rn → R, such that f (0) = 0 and sup x6=y f (x) − f (y) x − y < +∞.} Functions in Lip0(Rn) are the continuous functions from Rn into R such that f (0) = 0 with a distribution gradient in L∞(Rn). The vector space Lip0(Rn) is a Banach space, with norm kf kLip0(Rn) = sup x6=y f (x) − f (y) x − y = k∇f kL∞(Rn). Lip0(Rn) is the dual space of the Banach space denoted F (Rn), also known as the Lipschitz-free space over Rn. Let us recall that it has been recently shown by N. Weaver ([26]) that the free space over an arbitrary metric space M is strongly unique isometric predual of its dual Lip0(M). In particular, any Banach space whose dual is isometric to Lip0(Rn) coincide with F (Rn). Following [24], we represent the space Lip0(Rn) as a closed subspace of (L∞(Rn))n = L∞(Rn; Rn) (in fact the closed L∞ currents), and then we check that this closed subspace is exactly the orthogonal space to a subspace N of the predual L1(Rn; Rn). Thus the free space is identical with the quotient space L1(Rn; Rn)/N. This approach relies on de Rham's theorem on closed currents and an integration by parts. However, some technicalities are needed since derivatives must be taken in the distribution sense (see Remark 6). It should be noted that free spaces over convex open subsets of Rn are similarly represented in the recent work [3], without using Weaver's work - which requests a slightly different approach. We begin with two simple lemmas. Lemma 1. Let X = L1(Rn; Rn) be the Banach space of integrable vector fields and let N be the subspace of X made of vector fields with null distribution divergence: (2) N = {(fj)1≤j≤n ∈ X, X1≤j≤n ∂fj ∂xj = 0}. Then N is a closed subspace of X. SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 3 N.B. It is convenient to note the elements F = (fj)1≤j≤n ∈ L1(Rn; Rn) as vector fields F = X1≤j≤n fj ∂ ∂xj . The distribution divergence of F is then defined by div F = P1≤j≤n Proof. Let(cid:0)Fk = P1≤j≤n fk,j∂j(cid:1)k≥1 be a sequence of vector fields of N, converging in X with limit F = P1≤j≤n fj∂j (this means that for all j ∈ {1, . . . , n}, limk fj,k = fj in L1(Rn)). Let φ ∈ C ∞ c (Rn): we have ∂fj ∂xj . h ∂fk,j ∂xj , φiD ′(Rn),D(Rn) = −hfk,j, ∂φ ∂xj iD ′(Rn),D(Rn) = −ZRn fk,j(x) ∂φ ∂xj (x)dx, and consequently lim h k ∂fk,j ∂xj , φiD ′(Rn),D(Rn) = −ZRn fj(x) ∂φ ∂xj (x)dx = h ∂fj ∂xj , φiD ′(Rn),D(Rn), which implies 0 = limkh X1≤j≤n {z =0 ∂fk,j ∂xj thus div F = 0, proving the sought result. , φiD ′(Rn),D(Rn) = hP1≤j≤n } ∂fj ∂xj , φiD ′(Rn),D(Rn), and (cid:3) Lemma 2. The space Lip0(Rn) is isomorphic to the closed L∞(Rn) currents, i.e. to the subspace (3) Cn = {(uj)1≤j≤n ∈ (L∞(Rn))n, such that ∂uj ∂xk = ∂uk ∂xj for 1 ≤ j < k ≤ n}. More precisely, the mapping is an isomorphism of Banach spaces. Lip0(Rn) ∋ a 7→ da ∈ Cn, N.B. As in Lemma 1, we can prove that Cn is a closed subspace of the Banach space (L∞(Rn))n. All the derivatives are taken in the distribution sense. It is convenient to note the elements of Cn as u = P1≤j≤n ujdxj, so that for a ∈ Lip0(Rn), we have da = X1≤j≤n ∂a ∂xj dxj. Proof. From the definition of Lip0(Rn), we see that da is a L∞(Rn) current and also that da is closed since, in the distribution sense, we have ∂2a ∂xj∂xk = ∂2a ∂xk∂xj , meaning that the linear mapping given in the lemma is well-defined from Lip0(Rn) into Cn. This mapping is also isometric (and thus one-to-one) since kdakCn = k∇akL∞(Rn) = kakLip0(Rn). 4 GILLES GODEFROY AND NICOLAS LERNER For concluding the proof, we need only to prove that this mapping is onto: in fact thanks to de Rham's theorem on closed currents (see [5] or [19]), if u ∈ Cn, there exists a distribution w on Rn such that dw = u. As a result, the distribution w has a gradient in Lp loc for any p ∈ (1, +∞) and the Sobolev embedding theorem implies that (taking p > n) w is a (Hölder) continuous function. We can take now a(x) = w(x) − w(0), and we find that a belongs to Lip0(Rn) and satisfies da = u. (cid:3) We now state and prove a representation result for F (Rn). Proposition 3. Let X = L1(Rn; Rn) be the Banach space of integrable vector fields and let N be the closed subspace of X made of vector fields with null distribution divergence as defined by (2). Then, the free space F (Rn) over Rn is isometric to X/N and we have Lip0(Rn) = (X/N)∗. Proof. Note that it suffices to prove the last equation Lip0(Rn) = (X/N)∗ since by Weaver's result the isometric predual is unique. The case n = 1 is easy since, in that case N = {0}, so that X/N = L1(R); thanks to Lemma 2, we have also Lip0(R) = C1 = L∞(R), proving our claim which reduces to (cid:0)L1(R)(cid:1)∗ = L∞(R). Let us now assume that n ≥ 2. We start with a lemma. Lemma 4. We define X × Lip0(Rn) ∋ (f, a) 7→ Φ(f, a) = ZRn X1≤j≤n fj ∂a ∂xj dx ∈ R. The mapping Φ is bilinear continuous. Moreover for a ∈ Lip0(Rn) and f ∈ N (given by (2)),we have Φ(f, a) = 0. Proof of the lemma. The bilinearity and continuity of Φ are obvious. Let ρ ∈ C ∞ c (Rn; R+), supported in the unit ball, even with integral 1; we set for ǫ > 0, ρǫ(x) = ǫ−nρ(x/ǫ) and we define aǫ(x) = (a ∗ ρǫ)(x) = Z a(y)ρǫ(x − y)dy. We note that a ∈ C ∞ and daǫ = da ∗ ρǫ, which is thus bounded in (L∞(Rn))n by kdakL∞(Rn) and converges a.e. towards da, thanks to Lebesgue differentiation SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 5 Theorem1 (of course, no convergence in L∞ is expected). We have thus (4) ZRn X1≤j≤n fj ∂a ∂xj dx = lim ǫ ZRn X1≤j≤n fj ∂aǫ ∂xj dx = lim k→+∞ZRn X1≤j≤n ǫ (cid:16) lim fj(x) ∂aǫ ∂xj (x)χ0(x/k)dx(cid:17), where χ0 is a C ∞ in x ≤ 1. We note that c function, valued in [0, 1], equal to 1 on x ≤ 1/2 and supported (5) Z fj(x) ∂aǫ ∂xj = hfj(x), ∂ ∂fj ∂xj and since f ∈ N, we find = −h (x)χ0(x/k)dx = hfj(x), χ0(x/k) ∂aǫ ∂xj (x)iD ′,D ∂xj(cid:8)aǫ(x)χ0(x/k)(cid:9)iD ′,D − hfj(x), aǫ(x)(∂jχ0)(x/k)k−1iD ′,D (x), aǫ(x)χ0(x/k)iD ′,D −Z aǫ(x)fj(x)(∂jχ0)(x/k)k−1dx, (6) ZRn X1≤j≤n fj ∂a ∂xj dx = − lim ǫ (cid:16) lim k→+∞ZRn aǫ(x)(cid:0) X1≤j≤n fj(x)(∂jχ0)(x/k)k−1(cid:1)dx(cid:17). On the other hand, the term (∂jχ0)(x/k) is vanishing outside of {x, k/2 < x < k}, so that (7) (cid:12)(cid:12)(cid:12)(cid:12) ZRn aǫ(x)fj(x)(∂jχ0)(x/k)k−1dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ Z k 2 ≤x≤k aǫ(x) − aǫ(0)fj(x)dx k−1k∂jχ0kL∞(Rn) +ZRn aǫ(0)fj(x)dx k−1k∂jχ0kL∞(Rn). Since kdaǫkL∞(Rn) ≤ kdakL∞(Rn) = L < +∞, we obtain aǫ(x) − aǫ(0) ≤ Lx, so that the first term in the right-hand side of (7) is bounded above by Z k 2 ≤x≤k Lxfj(x)dx k−1k∂jχ0kL∞(Rn) ≤ Zx≥k/2 fj(x)dxLk∂jχ0kL∞(Rn), which is independent of ǫ and goes to 0 when k goes to +∞ since each fj belongs to L1(Rn). Moreover, we have a(0) = 0 and thus aǫ(0) = Z (cid:0)a(y) − a(0)(cid:1)ρ(−y/ǫ)ǫ−ndy, 1For u ∈ L∞(Rn), we have (u ∗ ρǫ)(x) − u(x) = = (cid:12)(cid:12)(cid:12)(cid:12) Z (cid:0)u(y) − u(x)(cid:1)ρǫ(x − y)dy(cid:12)(cid:12)(cid:12)(cid:12) ≤ u(y) − u(x)dy BnkρkL∞(Rn). 1 ǫnBn Zy−x≤ǫ {z →0, a.e. in x (Lebesgue's differentiation theorem) } 6 GILLES GODEFROY AND NICOLAS LERNER so that aǫ(0) ≤ LZ yρ(y/ǫ)ǫ−ndy ≤ ǫC0, C0 = Z zρ(z)dz, and we obtain that the second term in the right-hand side of (7) is bounded above by ZRn ǫC0fj(x)dx k−1k∂jχ0kL∞(Rn), which goes to 0 when k → +∞ since each fj belongs to L1(Rn). Finally the right- hand side of (7) goes to 0 when k → +∞ and this implies that the left-hand side of (6) is zero, which is the sought result. This lemma implies that the mapping Φ defined on X/N × Lip0(Rn) by where p : X → X/N is the canonical surjection, is well-defined and is a continuous bilinear mapping. (cid:3) Φ(cid:0)p(f ), a(cid:1) = Φ(f, a), • Going back to the proof of Theorem 3, we see that Φ induces a continuous linear mapping L from Lip0(Rn) into (X/N)∗ defined by Lip0(Rn) ∋ a 7→ L(a) ∈ (X/N)∗, (L(a))(p(f )) = Φ(f, a). We check first that L is one-to-one. Lemma 5. Let a ∈ Lip0(Rn) such that for all f ∈ X, Φ(f, a) = 0. Then we have a = 0. Proof of the lemma. Let χk ∈ C ∞ )1≤j≤n (which belongs to X) (χk ∂a ∂xj c (Rn; R+), χk = 1 on x ≤ k. We have for f = 0 = Φ(f, a) = ZRn χk(x) X1≤j≤n(cid:0) ∂a ∂xj (x)(cid:1)2 dx, which implies that da = 0 on x ≤ k for any k and thus da = 0, inducing a = 0. (cid:3) • Finally, let us prove that L is onto. Let ξ ∈ (X/N)∗; since ξ ◦ p ∈ X ∗ = (cid:0)L∞(Rn)(cid:1)n , we find (uj)1≤j≤n ∈ (cid:0)L∞(Rn)(cid:1)n hξ, p(f )i(X/N )∗,X/N = Z X1≤j≤n such that ∀f ∈ N, Z X1≤j≤n fjujdx = 0. ujfjdx, Let j, k be given in {1, . . . , n}. We have h − ∂uj ∂xk ∂uk ∂xj with fj = − ∂ϕ ∂xk and we get , ϕiD ′,D = huj, − ∂ϕ ∂xk iD ′,D + huk, ∂ϕ ∂xj iD ′,D = Z (cid:0)ukfk + ujfj(cid:1)dx, fk = ∂ϕ ∂xj . The vector field (fj∂j + fk∂k) is L1 with null divergence , ∂uj ∂xk − ∂uk ∂xj = 0, proving that the current u = P1≤j≤n ujdxj is closed and thus belongs to Cn (see (3)). Lemma 2 implies that there exists a ∈ Lip0(Rn) such that da = u, proving that L is onto. (cid:3) SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 7 Remark 6. The proof of Lemma 4 is giving a little bit more than the statement of this lemma: in fact Formula (5) holds without the assumption f ∈ N and we obtain from the sequel of the proof that, for (f, a) ∈ X × Lip0(Rn), and ρ, χ0 as in Lemma 4, ZRn X1≤j≤n fj ∂a ∂xj dx = − lim ǫ→0(cid:16) lim k→+∞ h X1≤j≤n ∂fj ∂xj , (a ∗ ρǫ)(x)χ0(x/k)iD ′,D(cid:17), a formula which can be written for the L1(Rn) vector field F = P1≤j≤n fj∂xj as (8) F (a)dx = − lim ǫ→0(cid:16) lim k→+∞ hdiv F, (a ∗ ρǫ)(x)χ0(x/k)iD ′,D(cid:17). ZRn c (Rn), the above formula follows from a standard integration When a belongs to C 1 by parts and the right-hand side of (8) is −hdiv F, aiD ′(1),C1 , although in the more general case tackled here, we have to pay attention to the fact that div F could be a distribution of order 1 which is not defined a priori on Lipschitz continuous functions. c Note that for n = 2, we have F (R2) = L1(R2; R2)/(cid:0)∇⊥L2(R2) ∩ L1(R2; R2)(cid:1), where ∇⊥ denotes the orthogonal gradient defined by ∇⊥ψ = (∂x2ψ, −∂x1ψ), and for n = 3, F (R3) = L1(R3; R3)/(cid:0)curl L3/2(R3; R3) ∩ L1(R3; R3)(cid:1). As shown in [4], the Lipschitz-free space Lip0(Rn) over Rn is weakly sequentially complete. Since it is now represented as a quotient space of L1, it is natural to wonder whether the kernel N of the quotient map is "nicely placed" (see [8]), in other words if its unit ball is closed in L1 for the topology τm of local convergence in measure. Indeed, the quotient of L1 by any such space enjoys a strong form of weak sequential completeness ([25]). But one has: Proposition 7. Let n > 1, let X = L1(Rn; Rn) be the Banach space of integrable vector fields and let N be the subspace of X of vector fields with null distribution divergence: N = {(fj)1≤j≤n ∈ X, X1≤j≤n ∂fj ∂xj = 0}. Then N is not nicely placed, that is, its unit ball is not closed for the topology τm of local convergence in measure. Proof. First observe that the space N is translation invariant, and thus is stable un- der convolution with integrable functions. If N is nicely placed, it follows from Bo- clé's differentiation lemma ([1]) that if a measure-valued vector field X ∈ (M(Rn))n is divergence-free, then its absolutely continuous part is divergence-free as well. Indeed, Boclé's Lemma shows that if (ck) is an approximation of identity in the convolution algebra L1(Rn) and µ is a singular measure, then (µ ∗ ck) converges to 0 in quasi-norm k . kp for all 0 < p < 1, and it follows that a nicely placed translation-invariant space of measures is stable under the Radon-Nikodym projec- tion (see the proof of Lemma 1.5 in [11]). Let us provide an example of an unstable divergence-free vector field in the case n = 2. 8 GILLES GODEFROY AND NICOLAS LERNER Let χ ∈ C 1 c (R2) be arbitrary, and H = 1R+. We consider the function ψ ∈ L2(R2) defined by The field ψ(x1, x2) = χ(x1, x2)H(x1) X = ∇⊥ψ = ∂ψ ∂x2 ∂ ∂x1 − ∂ψ ∂x1 ∂ ∂x2 , is divergence-free. Moreover X = ∂χ ∂x2 H(x1) ∂ ∂x1 −(cid:16) ∂χ ∂x1 H(x1) + χδ0(x1)(cid:17) ∂ = ∂χ ∂x2 H(x1) ∂χ ∂x1 H(x1) ∂ ∂x2 ∂x2 ∂ ∂x1 − V {z −χδ0(x1) W {z ∂ ∂x2 , } } The field V takes its values into L1, the field W is singular and div(V + W ) = 0, div W = − ∂χ ∂x2 δ0(x1) 6= 0, as soon as ∂χ (0, x2) does not vanish identically. This concludes the proof, and ∂x2 actually shows that the vector field V 6∈ N is the limit of the k . k1-bounded sequence (X ∗ ck) ⊂ N for the topology of local convergence in measure. (cid:3) k) = P z∗∗(x∗ z∗∗(weak∗ −P x∗ We recall that a Banach space Z has property (X) if every z∗∗ ∈ Z ∗∗ such that k) for every weakly unconditionally convergent series (x∗ k) actually belongs to Z (see p. 147 in [17]). In other words, property (X) means that elements of Z are those elements of Z ∗∗ which are somehow σ-additive. If Z has (X), then Z is strongly unique isometric predual for every equivalent norm, and is weakly sequentially complete ([14]). Moreover, every space which is L-complemented in its bidual has (X) ([25]). However, the following question seems to be open. Problem: Assume n > 1. Does the Banach space F (Rn) enjoy Property (X) ? 3. Closed subspaces of L1 consisting of continuous functions on a star-shaped domain. When X is a nicely placed subspace of L1, a distinguished subspace of X ∗ is a candidate for being the natural predual of X. We denote (see Definition IV.3.8 in [17]): X ♯ = {x∗ ∈ X ∗; x∗ is τm − continuous on BX}. The following proposition is valid in any separable L1-space, and requests no topology on the measure space. Proposition 8. Let X be a closed subspace of L1(m). Let (Tn) be a sequence of bounded linear operators from L1(m) to itself such that limn kTn(f ) − f k1 = 0 for every f ∈ L1. We assume that: (1) Tn(X) ⊂ X for every n ≥ 1, and the restriction of Tn to X is a weakly compact operator. SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 9 (2) Tn is (τm − τm)-continuous on k . k1-bounded subsets of L1 for every n ≥ 1. Then the space X is nicely placed and is isometric to the dual (X ♯)∗ of the space X ♯. Proof. Let (fk) be a sequence in BX , which τm-converges to g ∈ L1. By (2), for every n the sequence (Tn(fk))k is τm-convergent to Tn(g). Since by (1) this sequence is weakly relatively compact in L1, we have lim k kTn(fk − g)k1 = 0 and thus Tn(g) ∈ X for every n. But since (Tn) is an approximating sequence it follows that g ∈ X and thus X is nicely placed. Note now that for any h ∈ L∞ and any n, the restriction of T ∗ n (h) to X is τm- continuous on the unit ball of X, that is, belongs to X ♯. If follows that X ♯ separates X, and thus by Theorem 1.3 in [13] the space X ♯ is an isometric predual of X, and moreover it is an M-ideal in its bidual X ∗. (cid:3) The following theorem is the main application of Proposition 8. If Ω is a star- shaped open subset of Rn, ρ ∈ (0, 1) and f is any function defined on Ω, we denote Tρ(f )(x) = f (ρx) for every x ∈ Ω. We equip Ω with the topology induced by Rn and with the Lebesgue measure. We denote by τK the compact-open topology on the space C(Ω), that is, the topology of uniform convergence on compact subsets of Ω. With this notation, the following holds: Theorem 9. Let Ω be a star-shaped bounded open subset of Rn, and let X be a closed vector subspace of L1(Ω). We assume that X ⊂ C(Ω), and that Tρ(X) ⊂ X for every ρ ∈ (0, 1). Then the closed unit ball BX = {f ∈ X; kf k1 ≤ 1} of X is τK-compact and the topologies τK and τm coincide on BX. Proof. For showing this, pick any ρ ∈ (0, 1). We denote by K = Ω the closure of Ω in Rn, which is compact since Ω is bounded. The set Tρ(BX) is weakly relatively compact in L1(Ω), hence Tρ2(BX) is weakly relatively compact in C(K), and thus pointwise (on K) relatively compact in C(K). Since Tρ2(BX ) is also weakly rela- tively compact in L1(Ω), by Lebesgue's dominated convergence theorem it is k . k1- relatively compact in L1(Ω), and therefore Tρ3(BX ) is k . k∞-relatively compact in C(K). Since ρ ∈ (0, 1) was arbitrary, it follows that BX is relatively compact in C(Ω) for the compact-open topology τK. If we let T(n−1)/n = Tn for convenience, we can apply Proposition 8 with the same notation, and conclude that BX is τm-closed in L1. Note now that any τK-convergent sequence in BX is τm-convergent, and it follows that its limit belongs to BX since BX is τm-closed in L1. Therefore BX is τK-compact. Finally, compactness shows that the topologies τK and τm coincide on BX. (cid:3) The motivation for this result is that it implies that the unit ball BX of X is τm- compact locally convex, since τK is locally convex. Such subspaces of L1 have been previously studied in some detail ([8], [9]). It can be shown in particular that, under the mild assumption that they enjoy Grothendieck's approximation property, they yield to a satisfactory unconditional decomposition. The precise statement is given below, in the special case considered in Theorem 9. Note that it has been shown by W. B. Johnson and M. Zippin [21] that every quotient of c0 is isomorphic to a 10 GILLES GODEFROY AND NICOLAS LERNER subspace of c0, hence (2) below actually improves on (1). Observe that (2) implies that X is arbitrarily close to weak*-closed subspaces of l1. Corollary 10. Let Ω be a star-shaped bounded open subset of Rn, and let X be a closed vector subspace of L1(Ω). We assume that X ⊂ C(Ω), and that Tρ(X) ⊂ X for every ρ ∈ (0, 1). Then X = (X ♯)∗ isometrically, where X ♯ denotes the subspace of X ∗ consisting of the linear forms which are τK-continuous on BX. Moreover: (1) for any ǫ > 0, there exists a subspace Eǫ of c0 such that dBM (X ♯, Eǫ) < 1 + ǫ. (2) If X has Grothendieck's approximation property, then for any ǫ > 0 there is a quotient space Yǫ of c0 such that dBM (X ♯, Yǫ) < 1 + ǫ. Moreover there exists a sequence of finite rank operators (Ai) on X ♯ such that (a) supN,ǫi=1 kPN (b) for every f ∈ X, one has f = P∞ i=1 ǫiAik < 1 + ǫ. i=1 A∗ i (f ), where the series is norm-convergent. Proof. Since our assumptions imply that the unit ball BX of X is τm-compact locally convex, X = (X ♯)∗ isometrically and (1) follows from Proposition 2.1 in [8]. If X has the approximation property, it actually has the unconditional metric approxi- mation property (UMAP ) and moreover its natural predual X ♯ is arbitrarily close to quotients of c0 by Theorem 3.3 in [8]. Since X = (X ♯)∗ has (UMAP ) and X ♯ is an M-ideal in its bidual, and thus in particular a strict u-ideal, we may apply Theorem 9.2 in [7] which shows in particular that X ♯ has (UMAP ). Now Theorem 3.8 in [2] shows the existence of a sequence (Ai) of finite rank operators satisfying (a) and a weaker version of (b) where norm- convergence is replaced by weak*-convergence. But for every f ∈ X, the series i (f ) is weakly unconditionally convergent, hence norm-convergent since X (cid:3) does not contain c0. This concludes the proof. P∞ i=1 A∗ Remark 11. The proof allows to state some more results. Indeed the unit ball BX is τm-closed in L1(Ω) and thus by [10] the quotient space L1/X is weakly sequentially complete. Moreover, since this unit ball is even τm-compact locally convex, the space X satisfies by [22] the following extension result: if X ⊂ Y separable, any continuous linear operator from X to a C(K)-space Z extends to a continuous linear operator from Y to Z. Examples. 1) Proposition 8 trivially applies to any reflexive subspace X of L1, by taking Tn = IdL1 for all n. Note that in such a space X the τm-topology coincide with the norm topology, hence X ♯ = X ∗. This provides examples of spaces X such that (1) and (2) hold true, but Tn does not induce a compact operator on X ∗- take any infinite dimensional reflexive space X. 2) The subspace Har(Ω) of L1(Ω) which consists of harmonic functions satisfies the assumptions of Theorem 9. This is also the case for the Bergman space L1 a(Ω) of integrable holomorphic functions on Ω star-shaped open bounded in Cn ≃ R2n,. Actually, it is clear that many spaces of holomorphic functions on unit balls of Cn provide examples where Theorem 9 and Corollary 10 apply. 3) More generally, if G : Ω → (L∞, weak∗) is a continuous function, the space XG = {f ∈ L1(Ω); f (x) = ZΩ f (ω)G(x)(ω)dm(ω) f or all x ∈ Ω} SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 11 is k . k1-closed and consists of continuous functions. It follows that the space X(Gi) = Ti∈I EGi is k . k1-closed and consists of continuous functions for an ar- bitrary collection of continuous maps (Gi). When this space is moreover stable under the dilation operators Tρ, Theorem 9 applies. 4) If ∆ denotes the Laplace operator, a function f is called biharmonic if ∆2(f ) = ∆ ◦ ∆(f ) = 0. The space of biharmonic functions on Ω satisfies the assumptions of Theorem 9. positive real numbers such that inf i(λi+1 − λi) > 0 and Pi≥1 λ−1 5) The sequence (xk(t) = 2k(t2k ))k≥1 in L1([−1, 1]) is equivalent to the unit vector basis of l1 ([15]) and thus its closed linear span is contained in C((−1, 1)) and is isomorphic to l1. More generally, let Λ = (λi)i≥1 be an increasing sequence of i < +∞. The Müntz space M1(Λ) is the closed linear span of the sequence (tλi)i≥1 in L1([0, 1]). Then M1(Λ) is contained in C([0, 1)) (see [16]) and Theorem 9 and Corollary 10 apply to the space M1(Λ). Note that in [12] this result is shown using analyticity of the elements of M1(Λ) on (0, 1] but actually continuity suffices as shown above. Also, the point 0 does not belong to the interior of the unit interval but the reader will check that this causes no inconvenience in the above proofs. We refer to [6] for precise recent results on the geometry of Müntz spaces. Let us also mention that Müntz spaces of functions on the cube [0, 1]n have been investigated (see [18] and subsequent works), and Theorem 9 apply to such spaces as well. 6) Theorem 9 and Corollary 10 have an Lp-version for p > 1, and actually this version is rather easier since there is no need to enter the "Kalton zone" 0 ≤ p < 1 in this case. Recall that Theorem 4.4 in [23] states in particular that if 1 < p < +∞, a subspace X of Lp whose unit ball is k . k1-compact is arbitrarily close in Banach- Mazur distance to subspaces of lp. Along the lines of the above proofs, it follows that if 1 < p < +∞ and X ⊂ Lp(Ω) is a closed subspace which consists of continuous functions and such that Tρ(X) ⊂ X for every ρ ∈ (0, 1), then for every ǫ > 0, there is a subspace Eǫ ⊂ lp such that dBM (X, Eǫ) < 1 + ǫ. Note that in this case, there is no need to assume any approximation property. 7) It is interesting to compare the dilation operators with the approximation schemes from harmonic analysis. Let T be the unit circle equipped with the Haar measure, and (σn) the sequence of Fejér kernels. If we let Tn(f ) = f ∗ σn, then of course lim kTn(f ) − f k1 = 0 for every f ∈ L1(T). Any translation invariant subspace X = L1 Λ(T) satisfies Tn(X) ⊂ X, and weak compactness is obvious since the Tn's are finite rank operators. However, condition (2) of Proposition 8 fails. Actually, no non- zero weakly compact operator on L1 satisfies (2), since the existence of 0 6= F ∈ L∞ whose restriction to BL1 is τm-continuous would follow and there is no such F . We now consider two specific examples of translation-invariant subspaces of L1(T). Let X = L1 N(T) = H 1(D) be the classical Hardy space on the unit disc, seen as a subspace of L1(T). Then X is nicely placed and X = (X ♯)∗ ([10]) but the unit ball BX is not τm-compact. Actually, every infinite-dimensional translation- invariant subspace L1 Λ(T) contains an isomorphic copy of l2 (since Λ contains an infinite Sidon set) and thus fails to have a τm-relatively compact unit ball. The topology τm is strictly finer than the weak* topology associated with X ♯ = V MO on BX . However, weak* convergent sequences admit subsequences whose Cesaro 12 GILLES GODEFROY AND NICOLAS LERNER means are τm-convergent to the same limit (Corollary 4.3 in [11]). The operators Tn are (τm − τm)- continuous on BX , but not on X (by the argument from Example 3.6(b) in [17]). Let Λ = Sn≥1{k.2n; 0 < k ≤ n} and let X = L1 It follows from the proof of Theorem III.1 in [13] that the restrictions of the operators Tn to BX are (τm − τm)-continuous, but however the space X is not nicely placed. This shows in particular that condition (2) of Proposition 8 cannot be weakened: assuming (τm − τm)-continuity of the (Tn)'s on bounded subsets of X does not suffice to reach the conclusion. Λ(T). Let us conclude this note with an open question: Problem: Let Ω be a star-shaped open subset of Rn, and let X ⊂ L1(Ω) be a Banach space which satisfies the assumptions of Theorem 9. Assume that f ∈ X vanishes on a neighbourhood of 0. Does it follow that f = 0 ? References [1] Jean Boclé, Sur la théorie ergodique, Ann. Inst. Fourier (Grenoble) 10 (1960), 1–45. MR 0139717 [2] P. G. Casazza and N. J. Kalton, Notes on approximation properties in separable Banach spaces, Geometry of Banach spaces (Strobl, 1989), London Math. Soc. Lecture Note Ser., vol. 158, Cambridge Univ. Press, Cambridge, 1990, pp. 49–63. MR 1110185 [3] M. Cúth, O. Kalenda, and P. Kaplický, Isometric representation of Lipschitz-free spaces over convex domains in finite-dimensional spaces, arXiv, 1610.03966. [4] Marek Cúth, Michal Doucha, and Przemyslaw Wojtaszczyk, On the structure of Lipschitz-free spaces, Proc. Amer. Math. Soc. 144 (2016), no. 9, 3833–3846. MR 3513542 [5] Georges de Rham, Variétés différentiables. Formes, courants, formes harmoniques, Hermann, Paris, 1973, Troisième édition revue et augmentée, Publications de l'Institut de Mathéma- tique de l'Université de Nancago, III, Actualités Scientifiques et Industrielles, No. 1222b. MR 0346830 [6] L. Gaillard and P. Lefèvre, Lacunary Müntz spaces: isomorphisms and carleson embeddings, arXiv, 1701.05807v1. [7] Gilles Godefroy and N. J. Kalton, Unconditional ideals in Banach spaces, Studia Math. 104 (1993), no. 1, 13–59. MR 1208038 [8] G. Godefroy, N. J. Kalton, and D. Li, On subspaces of L1 which embed into l1, J. Reine Angew. Math. 471 (1996), 43–75. MR 1374918 [9] , Operators between subspaces and quotients of L1, Indiana Univ. Math. J. 49 (2000), no. 1, 245–286. MR 1777031 [10] Gilles Godefroy, Sous-espaces bien disposés de L1-applications, Trans. Amer. Math. Soc. 286 (1984), no. 1, 227–249. MR 756037 [11] [12] , On Riesz subsets of abelian discrete groups, Israel J. Math. 61 (1988), no. 3, 301–331. MR 941245 , Unconditionality in spaces of smooth functions, Arch. Math. (Basel) 92 (2009), no. 5, 476–484. MR 2506948 [13] Gilles Godefroy and Daniel Li, Some natural families of M -ideals, Math. Scand. 66 (1990), no. 2, 249–263. MR 1075142 [14] Gilles Godefroy and Michel Talagrand, Classes d'espaces de Banach à prédual unique, C. R. Acad. Sci. Paris Sér. I Math. 292 (1981), no. 5, 323–325. MR 608845 [15] V. Gurariıand V. I. Macaev, Lacunary power sequences in spaces C and Lp, Izv. Akad. Nauk SSSR Ser. Mat. 30 (1966), 3–14. MR 0190703 [16] Vladimir I. Gurariy and Wolfgang Lusky, Geometry of Müntz spaces and related questions, Lecture Notes in Mathematics, vol. 1870, Springer-Verlag, Berlin, 2005. MR 2190706 SOME NATURAL SUBSPACES AND QUOTIENT SPACES OF L1 13 [17] P. Harmand, D. Werner, and W. Werner, M -ideals in Banach spaces and Banach algebras, Lecture Notes in Mathematics, vol. 1547, Springer-Verlag, Berlin, 1993. MR 1238713 [18] Simon Hellerstein, Some analytic varieties in the polydisc and the Müntz-Szasz problem in several variables, Trans. Amer. Math. Soc. 158 (1971), 285–292. MR 0285724 [19] John Horváth, Topological vector spaces and distributions. Vol. I, Addison-Wesley Publishing Co., Reading, Mass.-London-Don Mills, Ont., 1966. MR 0205028 [20] Maria A. Japón Pineda and Chris Lennard, Second dual projection characterizations of three classes of L0-closed, convex, bounded sets in L1, J. Math. Anal. Appl. 342 (2008), no. 1, 1–16. MR 2440775 [21] W. B. Johnson and M. Zippin, Subspaces and quotient spaces of (P Gn)lp and (P Gn)c0, Israel J. Math. 17 (1974), 50–55. MR 0358296 [22] N. J. Kalton, Extension of linear operators and Lipschitz maps into C(K)-spaces, New York J. Math. 13 (2007), 317–381. MR 2357718 [23] Nigel J. Kalton and Dirk Werner, Property (M ), M -ideals, and almost isometric structure of Banach spaces, J. Reine Angew. Math. 461 (1995), 137–178. MR 1324212 [24] N. Lerner, A note on Lipschitz spaces, 2016-2-17. [25] Hermann Pfitzner, Separable L-embedded Banach spaces are unique preduals, Bull. Lond. Math. Soc. 39 (2007), no. 6, 1039–1044. MR 2392827 [26] N. Weaver, On the unique predual problem for Lipschitz spaces, arXiv, 1611.01812v2. G. Godefroy and N. Lerner, Institut de Mathématiques de Jussieu, Université Pierre et Marie Curie (Paris VI), 4 Place Jussieu, 75252 Paris cedex 05, France E-mail address: [email protected] ; [email protected]
1203.5760
11
1203
2018-11-30T13:54:21
A note on the rate of convergence for a sequence of random polarizations
[ "math.FA" ]
It was shown by Burchard and Fortier that the expected $L^1$ distance between $f^*$ and $n$ random polarizations of an essentially bounded function $f$ with support in a ball of radius $L$ is bounded by $2dm(B_{2L})||f||_{\infty}n^{-1}$. The purpose of this note is to expand on that result. It is shown that the same expected $L^1$ distance is bounded by $c_nn^{-1}$ with $\limsup_{n\rightarrow \infty}c_n \leq 2^{d+1}||\nabla f||_1$ for every $f \in W_{1,1}(B_L) \cap L^{\infty}(B_L)$. Furthermore, the aforementioned expected $L^1$ distance is $O(n^{-1/q})$ for $f \in L^p(B_L)$ with $p>1$ and $\frac{1}{p} + \frac{1}{q} = 1$. An exponential lower bound is provided for the expected measure of the symmetric difference between the random polarizations of measurable sets and their Schwarz symmetrization. Finally, the rate $n^{-1}$ is shown to be, in a sense, best possible for the random polarizations of measurable sets: the expected symmetric difference between the random polarization of a ball and its corresponding Schwarz symmetrization decays at the rate $n^{-1}$.
math.FA
math
Rate of Convergence Estimates for Random Polarizations on Rd Marc Fortier ∗ January 9, 2013 Abstract It is shown in [1] that the expected L1 distance between f ∗ and n random polarizations of a function f with support in a ball of radius L is bounded above by 2dm(B2L )f ∞n−1 . Here it is shown that the d f − f ∗ L1 with bd > 2−1 expected distance is bounded below by bn some numerical constant depending only on d and, if the boundary of the sets {f > t} have measure zero for almost every t, equals n−1cn with lim supn→∞ cn ≤ L2d+1 Per(f ∗ > t)L1 . The paper concludes with the study of random polarizations of balls which shows that the aformentioned O(n−1) rate of convergence estimate is sharp for d = 1. 1 Introduction It is well known that for any function f ∈ Lp (1 ≤ p < ∞) there exists a sequence of polarizations that, when applied iteratively to f , will yield a se- quence of functions which converges in Lp to f ∗ - the symmetric decreasing rearrangement of f . In fact, one can construct explicit sequences of polar- izations that yield convergence to the symmetric decreasing rearrangement for any initial function in Lp [3]. These sequences also yield uniform conver- gence to the symmetric decreasing rearrangement when applied to any initial continuous function with compact support and also convergence in Hausdorff metric when applied to compact sets. Other convergence results have been proven. For instance, it is shown in [1] and [2] that random sequences of po- larizations can also yield almost sure convergence to the symmetric decreas- ing rearrangement. The first result on rates of convergence of polarizations It is shown in to the symmetric decreasing rearrangement appears in [1]. ∗ [email protected] 1 that paper that there exists a probability measure PL defined on the space of infinite sequences of polarizations with the following rate of convergence property: Proposition 1. [1, p.19]If f ∈ L1(Rd ) is bounded with support in BL then EPL (f σ1⋯σn − f ∗ L1 ) ≤ 2dm(B2L )f ∞n−1 . (1) It is shown in this paper that the following holds: Proposition 2. If f ∈ L1 with support in BL then d f − f ∗ L1 ≤ EPL (f σ1⋯σn − f ∗ L1 ) = n−1 cn bn with bd > 2−1 a numerical constant that depends only on d and Per {f > t}∗ dt. cn ≤ L2d+1 S ∞ lim sup n→∞ 0 (3) (2) 2 Notation and Preliminary Results 2.1 Notation In what follows, m will denote Lebesgue measure; M will denote the sigma algebra of Lebesgue measurable sets; Bt will denote the ball of radius t centered at the origin and ωd will denote the surface area of B1 . The space of reflections that do not map the origin to the origin will be denoted by Ω. σx,y will denote the unique reflection that maps x to y . Finally, R(A) will denote the outer radius of a set A and dH (⋅, ⋅) will denote the Hausdorff distance between two sets. 2.2 Rearrangements A rearrangement T is a map T ∶ M → M that is both monotone (A ⊂ B implies T (A) ⊂ T (B )) and measure preserving (m(T (A)) = m(A) for all A ∈ M). We say that a non-negative measurable function f vanishes at infinity if m(f > t) < ∞ (4) for all t > 0. If f vanishes at infinity then we can define its rearrangement T f by using the “layer cake principle” χT (f >t) (x)dt = sup{t ∶ x ∈ T (f > t)}. T f (x) = S ∞ 0 (5) 2 f σ (x) = 2.3 Polarization Let σ ∈ Ω and let f be an arbitrary function. We define the polarization of f with respect to σ as ⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩ if x ∈ X σ f (x) ∨ f (σ(x)) + f (x) ∧ f (σ(x)) if x ∈ X σ − if x ∈ X σ f (x) 0 . If A is an arbitrary set then its polarization with respect to σ is simply the polarization of χA and is denoted by Aσ : 0 ) . + ) ∪ (A ∩ X σ + ) ∩ A) ∪ (A ∩ X σ − ) ∩ Ac ) ∪ (σ(A ∩ X σ Aσ = (σ(A ∩ X σ In other words, Aσ is the same as A except that the part of A contained in X σ − whose reflection does not lie in A is replaced by its reflection in X σ + . As a result, polarization is measure preserving. It is clear from equation 6 that if f ≤ g then f σ ≤ g σ for all σ ∈ Ω and thus polarization is monotone on M i.e., polarization is a rearrangement. One can check directly that {f σ > t} = {f > t}σ for all σ ∈ Ω and, by equation 5), f σ is the rearrangement of f with respect to the polarization rearrangement. (6) (7) 2.4 Schwarz Symmetrization For any A ∈ M there exists a unique open ball centered at the origin A∗ with If f (x) the same measure as A called the Schwarz symmetrization of A. vanishes at infinity then its rearrangement with respect to the Schwarz rear- rangement is denoted by f ∗(x). It is clear that f ∗(x) is radially decreasing i.e., f ∗(x) ≤ f ∗(y ) if x ≥ y and f (x) = f (y ) if x = y . In the literature, f ∗ is also called the symmetric decreasing rearrangement of f. If f vanishes at infinity, we let rf (t) =  m(f > t) 1~d (8) κd denote the radius of the open ball {f > t}∗ . The distribution function of a function f vanishing at infinity is always right continuous and thus so is rf (t). In particular, we have (9) {f ∗ > t} = {f > t}∗ for all t ≥ 0 and thus f ∗ is right continuous. 3 (10) 2.5 Random Polarizations A probability measure PL on ∏∞ i=1 Ω can be constructed by letting PL(A1 × ⋯ × An) = nMi=1 (2Lωd)−1 S{σ(0)∶σ∈Ai } x−(d−1)dx. The following proposition motivates the choice of PL . Proposition 3. If x ∈ Rd then Tx(y ) = σx,y (0) has Jacobian  Tx(y ) x − y d−1 PL(σ1 (x) ∈ E ) = (2Lωd )−1 SE x − y −(d−1) dy for every x ∈ BL . Proof. See [1, p.18] for the proof of equation 11. Equation 12 follows directly from equation 11 and the change of variables formula for integrals. and (11) (12) 3 Rate of Convergence (13) Proof of proposition 2. It is shown in [1, p.19] that EPL (m ({f > t} △ {f ∗ > t})) − m ({f σ1 > t} △ {f ∗ > t}) equals 2 EPL (m({f > t ≥ f ∗} ∩ {f ∗(σ1 (x)) > t ≥ f (σ1 (x))})) (14) for any function f ∈ L1 with support in BL . Applying Fubini’s theorem to equation 14; setting σ(x) = y and using Proposition 3 shows that 14 equals (Lwd )−1 SA0 (t) SB0 (t) x − y −(d−1)dx dy (15) A0 (t) = {f > t, f ∗ ≤ t}, B0 (t) = {f ∗ > t, f ≤ t}. An(t) = {f σ1⋯σn > t, f ∗ ≤ t}, Bn (t) = {f ∗ > t, f σ1 ⋯σn ≤ t} with (16) If and (Lωd )−1m(An (t))−2 SAn (t) SBn (t) x − y −(d−1)dx dy = φf σ1 ⋯σn (t) 4 (17) then, by recalling equation 15, one sees that the sequence zn (t) = EPL (m ({f σ1⋯σn > t} △ {f ∗ > t})) satisfies the relation (18) n(t)αn (t) zn(t) − zn+1(t) = z 2 (19) with αn(t) = EPL (φf σ1 ⋯σn (t)m(An (t))2 ) zn (t)−2 . This implies that zn(t) = n−1 cn(t) with cn(t) = z0(t) − zn (t) αi−1 (t) 1 − αi−1(t)zi−1 (t) −1 z0 (t) cn(t)dt = n−1 cn . zn = EPL (f σ1⋯σn − f ∗ L1 ) = n−1 S ∞ 0 n−1 nQi=1 (20) (21) and Moreover, (22) (23) dt lim sup n→∞ lim sup n→∞ lim sup n→∞ cn ≤ S ∞ cn(t)dt 0 1 − αn(t)zn (t) ≤ S ∞ αn(t) 0 and similarly for the lim inf . By Jensen’s inequality, αn(t) ≥ 4−1EPL φ−1 f σ1 ⋯σn (t)−1 f σ1⋯σn (t) ≤ 2d−1LωdEPL R({f σ1⋯σn > t})d−1  . EPL φ−1 As discussed in the introduction, if τn is a sequence of polarizations that transforms any initial L1 function to its symmetric decreasing rearrangement then dH (F τ1⋯τn , F ∗) → 0 as n → ∞ for any compact set F . In particular, ωdEPL R({f σ1⋯σn > t})d−1  = Per {f > t}∗ lim (26) n→∞ (25) (24) and and lim sup n→∞ Per {f > t}∗ dt. cn ≤ L2d+1 S ∞ 0 For the lower bound on the rate of convergence, we first apply the Riesz rearrangement inequality, SAn (t) SBn (t) x − y −(d−1)dx dy ≤ SAn (t)∗ SBn (t)∗ x − y −(d−1)dx dy (28) 5 (27) (29) (30) (31) with SB1 SB1 x − y −(d−1)dxdy . and the right-hand side of 28 equals  m(An(t)) m(B1 ) (d+1)~d This implies that zn−1 − zn ≤ S ∞ en−1 (t)zn−1 (t)dt 0 EPL (m(An−1 (t))1+1~d ) en−1 (t) = cd EPL (m(An−1 (t))) for some constant cd only depending on d. By induction, one obtains n−1Mi=0 (1 − ei(t))z0 (t) dt zn ≥ S ∞ 0 n−1Mi=0 (1 − ei (t))1~n ≥ z0(t)L1 S ∞ 0 for all n ≥ 1. To complete the proof, we have cd = (2Lωd )−1m(B1 )−(d+1)~d SB1 SB1 x − y −(d−1)dxdy = (2L)−1m(B1 )−1~d γd (35) and thus ei(t) is bounded above by γdrf (t)~2L < 2−1γd . Setting bd = 1 − 2−1γd completes the proof. z0 (t) z0(t)L1 dtn (32) (33) (34) 4 Example Let x0 denote the center of a ball with x0 < L − r and radius r and let Xn denote the distance from the origin of the center of n random polarizations of x0 + Br . The functions φn(x) = (2Lωd )−1 Sy <x y n x − y −(d−1)dy (36) have the scaling property φn(λx) = λn+1φn(x). As a result, if u is any unit vector then n−1  n−1 = −ckEPL X k+1 n  − EPL X k EPL X k (37) 6 with ck = (2Lωd)−1 Sy <1(1 − y k)u − y −(d−1)dy . If zn denotes the expected value of Xn then, by using equations 37 and setting c0 = 1, one obtains the following recurrence relations for zn , (−1)n k (−1)k zn−k = △n (z0 ) = zn+1 nQk=0 n nMk=0 ck 0 and the representation (39) (38) (40) (41) tk dt k (−1)k z k+1 zn = nQk=0 n kMi=0 ci . 0 For x ≤ 2r , the symmetric difference Br △ (x + Br ) is given by ψr (x) = 2 m(Br ) − rd−1ωd S arccos(x~(2r)) sind (t) dt . 0 Equation 41 yields EPL (ψ(Xn )) ∼ ψ ′(0)zn = Per(Br )zn as n → ∞. In particular, if f is radially decreasing around x0 with support in BL , then EPL (f σ1⋯σn − f ∗ L1 ) ∼ zn Per(f ∗ > t)L1 as n → ∞. Setting d = 1, we obtain the exact representation k (−1)k S z0 ~4L nQk=0 n zn = 4L 0 = 4L S z0 ~4L (1 − t)n dt 0 n + 1 1 − (1 − z0 ~4L)n+1(cid:6) = 4L and the exact asymptotic n Per(f ∗ > t)L1 . EPL (f σ1⋯σn − f ∗ L1 ) ∼ 4L (45) Setting d = 1 in Proposition 2, we see that the rate of convergence estimate is sharp for d = 1. We can strengthen this result by analyzing the characteristic function of the random variables Xn . We have j (4Lx)j −1 (it)j EPL eitXn  = 1 + 4L S z0 ~4L (1 − x)n ∞Qj =1 j ! 0 exit(1 − x~4L)ndx = 1 + it S z0 0 7 (44) dx (46) (42) (43) (47) lim n→∞ exit e−x~4Ldx. EPL eitnXn  = 1 + it S ∞ 0 This shows that nXn converges weakly to an exponential distribution with mean 4L. Unfortunately, the constants cn are not easily computable for di- mensions d > 1, but the method of proof for d = 1 can be used to make a connection between this problem and the Hausdorff moment problem. Set- ting z0 = 1 in equation 40 gives n ) jMi=0 jMi=0 ci ≥ 0. ci = EPL (X j (−1)n △n Hausdorff has shown that if 49 holds then there exists a unique probability measure µ such that xk dµ(x) = kMi=0 S 1 0 (48) (49) (50) and ci . This gives and with We have n ) = k−1Mi=0 (ci~z0 )−1 S 1 EPL (X k 0 xk−1 (1 − xz0 )ndµ(x) EPL (eitXn ) = 1 + it S 1 0 (1 − xz0 )nψ(itxz0 )dµ(x) ψ(x) = ∞Qi=0 xk (k + 1)! kMi=0 c−1 i . (51) (52) (53) k−1Mi=0 (ci ~z0)−1 EPL ((nXn)k ) ≥ lim inf nµ(x < n−1 )e−z0 lim inf (54) n→∞ n→∞ for all k ≥ 1 but equation 37 gives EPL (Xn − Xn+1 ) ≥ c1EPL (Xn ) which implies EPL (Xn) ≤ c−1 1 n−1 ; consequently, if lim inf n→∞ nµ(x < n−1) > 0 then we have zn ∼ cn−1 for some numerical constant c depending on d. References [1] Almut Burchard and Marc Fortier, Random polarizations, Preprint arXiv:1104.4103v2 (2011), 30 Pages. 8 [2] Jean Van Schaftingen, Approximation of symmetrizations and symmetry of critical points, Topol. Methods Nonlinear Anal. 28 (2006), no. 1, 61–85. [3] , Explicit approximation of the symmetric rearrangement by po- larizations, Arch. Math. (Basel) 93 (2009), no. 2, 181–190. 9
1001.0718
5
1001
2010-11-30T23:11:04
Invariant expectations and vanishing of bounded cohomology for exact groups
[ "math.FA", "math.GR", "math.OA" ]
We study exactness of groups and establish a characterization of exact groups in terms of the existence of a continuous linear operator, called an invariant expectation, whose properties make it a weak counterpart of an invariant mean on a group. We apply this operator to show that exactness of a finitely generated group $G$ implies the vanishing of the bounded cohomology of $G$ with coefficients in a new class of modules, which are defined using the Hopf algebra structure of $\ell_1(G)$.
math.FA
math
INVARIANT EXPECTATIONS AND VANISHING OF BOUNDED COHOMOLOGY FOR EXACT GROUPS RONALD G. DOUGLAS AND PIOTR W. NOWAK Abstract. We study exactness of groups and establish a characterization of ex- act groups in terms of the existence of a continuous linear operator, called an invariant expectation, whose properties make it a weak counterpart of an invari- ant mean on a group. We apply this operator to show that exactness of a finitely generated group G implies the vanishing of the bounded cohomology of G with coefficients in a new class of modules, which are defined using the Hopf algebra structure of ℓ1(G). 1. Introduction Exactness is a weak amenability-type property of finitely generated groups. It was defined in [12] in terms of properties of the minimal tensor product of the reduced group C∗-algebra. Similar to amenability, exactness has a few equivalent definitions which are each of separate interest in different areas of mathematics. In particular, exactness is equivalent to the existence of a topologically amenable action of the group on a compact space [10] and to Yu's property A [8, 21]. For this reason exactness has many interesting applications in analysis, geometry and topology. Most notably, Yu [30] proved that groups with property A satisfy the Novikov conjecture. Amenable groups are precisely the ones which carry an invariant mean; that is, a functional on ℓ∞(G) which is positive, preserves the identity and is invariant under the natural action of G on ℓ∞(G). Invariant means allow for averaging on amenable groups, which is precisely what makes such groups so convenient to work with. In the case of exact groups there was no parallel characterization and in this article we investigate the existence of such a counterpart and some of its applications. Coarse-geometric versions of classical notions or results in group theory can sometimes be obtained by considering the problem "with coefficients in ℓ∞(G)", where the precise meaning of this phrase varies with the context. For instance, the coarse Baum-Connes conjecture for a group G is the Baum-Connes conjecture with coefficients in the algebra ℓ∞(G, K), where K denotes the compact operators on an infinite-dimensional, separable Hilbert space [29], see also [27]. Similarly, in coarse geometry the reduced crossed product ℓ∞(G) ⋊r G is an analogue of the Key words and phrases. exact group; bounded cohomology; Hochschild cohomology; convolu- tion algebra; invariant expectation. During this research the first author was partially supported by NSF grant DMS-0600865. The second author was partially supported by NSF grant DMS-0900874. 1 2 RONALD G. DOUGLAS AND PIOTR W. NOWAK reduced group C∗-algebra of G. It is this point of view that suggests the study of the space of operators from a given space into the algebra ℓ∞(G), which, loosely speaking, plays the role of a "dual with coefficients in ℓ∞(G)". Let G be a finitely generated group. Given a left Banach G-module X we con- sider the space L(X, ℓ∞(G)) of continuous linear operators from X to ℓ∞(G). This space is naturally a bounded Banach G-module by pre- and post-composing with the actions of G on X and ℓ∞(G). Additionally, we can identify L(X, ℓ∞(G)) with the space ℓ∞(G, X∗), which is a dual space to ℓ1(G, X). Using this identification we can naturally equip L(X, ℓ∞(G)) with a weak-* topology. An invariant expectation on the group G is then an operator M from the G- module L(ℓu(G), ℓ∞(G)) into ℓ∞(G), where ℓu(G) is a certain Banach algebra as- sociated to the group, obtained as a completion of the algebraic crossed product ℓ∞(G) ⋊alg G (the precise definition is given in 2.1). The properties of the invariant expectation are similar to properties of invariant means on groups; namely, it is positive, unital in an appropriate sense and a limit of elements of a special form. The most important property of a mean, invariance, is replaced here by equivari- ance with respect to the G-actions. We refer to Definition 3.7 for details. Theorem 1.1. A finitely generated group G is exact if and only if there exists an invariant expectation on G. The existence of an invariant expectation and its relation to exactness of groups relies on the properties of a certain subspace W of L(L(ℓu(G), ℓ∞(G)), ℓ∞(G)). It is the choice of this subspace that allows us to carry out approximation arguments in a setting suitable for exactness. We apply the above characterization to compute the bounded cohomology of ex- act groups with coefficients in a new class of Banach modules. The motivation for studying bounded cohomology in the context of exactness comes from a question of N. Higson, who asked if there is a cohomological characterization of exactness. The first such characterization was proved in [3], but in terms of a cohomology theory introduced in that article. One natural direction to investigate is whether the characterization of amenability in terms of bounded cohomology, due to B.E. John- son [11], could be generalized to our setting. Here we show, using invariant expec- tations, that bounded cohomology with coefficients in a class of modules defined below, vanishes for exact groups. Characterizations of exactness via vanishing of bounded cohomology were proved later in [1] and [16], independently, using some of the ideas presented here. The space L(X, ℓ∞(G)), in addition to being a G-module, also carries a natural structure of an ℓ∞(G)-module, given by multiplying the image of an operator T by an element of ℓ∞(G). It is the existence of this second structure that is essential in our considerations. A G-submodule E ⊆ L(X, ℓ∞(G)), which additionally is an ℓ∞(G)-module in this sense, will be called a Hopf G-module, as its two module structures are defined by the actions of a Hopf-von Neumann algebra ℓ∞(G) and its predual Banach algebra ℓ1(G). By H1 b(G, E) we denote the bounded cohomology group of G in degree 1 with coefficients in a Banach G-module E. INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 3 Theorem 1.2. Let G be a finitely generated group. If G is exact then H1 for any weak-* closed Hopf G-module. b(G, E) = 0 Bounded cohomology allows the use of dimension shifting techniques: given a G-module M there is a "shifting module" ΣM such that Hn+1 b(G, ΣM). An argument due to N. Monod shows that shifting modules of weak-* closed Hopf G-modules are again weak-* closed Hopf G-modules, which allows to conclude that all the higher bounded cohomology groups with coefficients in dual Hopf G- modules also vanish for exact groups. (G, M) = Hn b The above results, after being circulated in preprint form, were followed by a rapid development of the relation between bounded cohomology and exactness. The papers [1] and [16] mentioned earlier, as well as [2] and [6], contain related results. We would like to thank Nicolas Monod for many valuable comments, as well as suggesting the name Hopf modules and for allowing us to include Proposition 5.1 in this paper. We also thank the referee for helpful comments. Contents Introduction 1. 2. Algebras, duality and topologies 2.1. The uniform convolution algebra ℓu(G) 2.2. G-duality for ℓu(G) and ℓ∞(G, C ) 2.3. Weak topologies 3. Exactness and invariant expectations 3.1. Exact groups 3.2. 4. Bounded cohomology of exact groups 4.1. Hopf G-modules 5. Concluding remarks 5.1. Dimension reduction and higher cohomology groups. 5.2. Relation between M and W. References Invariant expectations in L(L(ℓu(G), C ), C ) 1 3 4 5 5 6 6 7 14 14 16 16 17 17 2. Algebras, duality and topologies Let G be a discrete group generated by a finite set S , which is symmetric; that is, S = S −1. All Banach spaces we discuss are over R. Since in most of our arguments we will view the algebra ℓ∞(G) as an algebra of coefficients, throughout the article we will shorten the notation to C = ℓ∞(G). We denote by 1G the identity in ℓ∞(G) and by k · kC the usual supremum norm. The algebra C is equipped with a natural G-action, (1) (g ∗ f )(h) = f (g−1h) 4 RONALD G. DOUGLAS AND PIOTR W. NOWAK for g, h ∈ G and f ∈ C . For a function f : G × G → R we will denote fg = f (g, ·), so that fg is a real function on G. The set of those g ∈ G for which fg , 0 is called the support of f and denoted supp f . 2.1. The uniform convolution algebra ℓu(G). Let Cc(G, C ) = (cid:8) f : G → C : # supp f < ∞(cid:9) , where supp f denotes the support of f . By the above, Cc(G, C ) can be viewed as a linear subspace of the space of bounded functions on G ×G, each of which vanishes outside of K × G, for some finite K ⊆ G. Cc(G, C ) is a linear space and we equip it with the norm The space Cc(G, C ) is, in a natural way, a subspace of(cid:16)Lg∈G ℓ1(G)(cid:17)∞ , the infinite direct sum of copies of ℓ1(G) with the norm kηk = supg∈G kη(g)k1, where η : G → ℓ1(G), and these norms agree on Cc(G, C ). We define multiplication on Cc(G, C ) by the formula k f ku =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xg∈G . fg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C ( f ⋆ f ′)g =Xh∈G fh(h ∗ f ′ h−1g) g = g ∗ fg−1, which turns Cc(G, C ) into the algebraic crossed and an involution f ∗ product C ⋊alg G. It can be easily seen that k f ⋆ f ′ku ≤ k f kuk f ′ku for f, f ′ ∈ Cc(G, C ). Definition 2.1. The uniform convolution algebra, denoted ℓu(G), is the completion of Cc(G, C ) with respect to the norm k · ku. Note that there is a natural isometric inclusion of ℓ1(G) in ℓu(G). Indeed, con- sider f ∈ ℓ1(G) and define ξg = f (g)1G. For each element g ∈ G consider δg ∈ ℓu(G) given by (δg)h =( 1G 0 if h = g, otherwise. We have δgh = δg ⋆ δh and there is a natural action of G on ℓu(G) by isometries, also denoted by ⋆, such that (2) g ⋆ ξ = δg ⋆ ξ, for ξ ∈ ℓu(G). Observe also that δe, where e ∈ G is the identity element, is the unit in ℓu(G). INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 5 2.2. G-duality for ℓu(G) and ℓ∞(G, C ). Consider the Banach space ℓ∞(G, C ) =  f : G → C : sup g∈G k fgkC < ∞ . We denote by 1G the identity in ℓ∞(G, C ): (1G)h = 1G for every h ∈ G. This space is naturally isometrically isomorphic to ℓ∞(G × G); however, the above notation has advantages in our setting. The space ℓ∞(G, C ) is a left G- module with an action given by (3) (g ⊙ f )h = g ∗ fg−1h, where g, h ∈ G. There is a natural inclusion ℓu(G) ⊆ ℓ∞(G, C ) and the two actions ⋆ and ⊙ agree on ℓu(G). There exists a natural C -valued pairing between the elements of ℓu(G) and ℓ∞(G, C ), h·, ·iC : ℓu(G) × ℓ∞(G, C ) → C given by (4) hξ, f iC =Xg∈G ξg fg. It is well-defined since hξ, f iC ≤Xg∈G ξg fg ≤ k f kℓ∞(G,C )kξku1G. Lemma 2.2. For ξ ∈ ℓu(G) and f ∈ ℓ∞(G, C ) we have hg ⋆ ξ, f iC = g ∗ hξ, g−1 ⊙ f iC . Proof. We have hg ⋆ ξ, f iC (h) = Xk∈G = Xk∈G (g ⋆ ξk(h)) fk(h) ξg−1k(g−1h) fk(h). On the other hand, (cid:16)g ∗ hξ, g−1 ⊙ f iC(cid:17) (h) = hξ, g−1 ⊙ f iC (g−1h) = Xk∈G = Xk∈G ξk(g−1h) (cid:16)(g−1 ⊙ f )k(g−1h)(cid:17) ξk(g−1h) (cid:16) fgk(gg−1h)(cid:17) Substituting gk = k′ we see that the two expressions are equal. (cid:3) 2.3. Weak topologies. Let X be a Banach space. One of the main objects of our study will be the module L(X, C ) of bounded linear maps from X to C , with its natural operator norm, which we denote by k · kL. If not stated otherwise we consider C as a dual of ℓ1(G) with its natural weak-* topology. We will denote weak-* limits in C by w∗ − lim. The action ∗ defined in (2) of G on C is weak-*-continuous. 6 RONALD G. DOUGLAS AND PIOTR W. NOWAK The weak-* topology on L(X, C ). The space L(X, C ) is isometrically isomorphic to the space ℓ∞(G, X∗), where the isomorphism I : L(X, C ) → ℓ∞(G, X∗) is given by (I(T )g)(x) = (T (x))g for T ∈ L(X, C ), x ∈ X and g ∈ G. Thus our space L(X, C ) can be identified as the Banach space dual of ℓ1(G, X) and as such it can be naturally equipped with a weak-* topology. The following description of the weak-* topology on L(X, C ) is convenient in our setting. Proposition 2.3. Let X be a Banach space and let {Tβ} be a net in L(X, C ). The following conditions are equivalent: (a) C − limβ Tβ = T , (b) w∗ − limβ Tβ(x) = T (x) in C for every x ∈ X. The weak topology on X. Every element ξ ∈ X defines a map ξ : L(X, C ) → C by the formula ξ(T ) = T (ξ) for every T ∈ L(X, C ). This defines a natural embedding i : X → L (L(X, C ), C ) . We denote the natural norm on L (L(X, C ), C ) by k · kLL. Since the dual space X∗ of X is naturally embedded in L(X, C ) by defining T (x) = ϕ(x)1G, where ϕ ∈ X∗, we easily see that the embedding is isometric. Definition 2.4. Let X be a Banach space. The weak topology on X is the restriction to X of the weak-* topology on L(L(X, C ), C ). This gives the following descrition Proposition 2.5. A net {xβ} of elements of X converges weakly to x ∈ X if and only if for every T ∈ L(X, C ) we have T (x) = w∗ − limβ T (xβ). The weak topology of Definition 2.4 is the same as the weak topology on X in the classical sense. For our purposes it suffices to see that the weak topology in the sense of Definition 2.4 is formally stronger than the weak topology on X in the classical sense. 3. Exactness and invariant expectations 3.1. Exact groups. The term exact group originates in the theory of C∗-algebras. However, in the last decade many new characterizations were discovered and our use of this term is not restricted strictly to the C∗-algebraic definition. Originally exact groups were defined by Kirchberg and Wassermann, see [12, 13] in the study of group C∗-algebras. A C∗-algebra A is exact if given any exact sequence 0 −→ I −→ B −→ B/I −→ 0 INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 7 the sequence 0 −→ I ⊗min A −→ B ⊗min A −→ B/I ⊗min A −→ 0 remains exact. Note that the maximal tensor product always preserves short ex- act sequences in the above sense. Exactness of a C∗-algebra is weaker than its nuclearity. Indeed, A is nuclear if B ⊗min A = B ⊗max A for any C∗-algebra B. A group G is called exact if its reduced C∗-algebra C∗ r (G) is exact in the above sense. We refer to [4, 12, 13, 28] for details. Exactness of a group turned out to be equivalent to property A of Yu due to work of Guentner and Kaminker [8] and, subsequently, Ozawa [21]. Property A was introduced in [30] as a condition sufficient to enable one to embed a group (or, more generally, a metric space) coarsely into a Hilbert space. At present there are no known examples of groups which embed coarsely into the Hilbert space but do not have property A (see, however, [18]). Property A in [30] was defined in terms of a Følner-type condition which highlights the fact that it can be viewed as a weak amenability-type property. We refer to [19, 25, 28] for an introduction to property A. In [10] Higson and Roe characterized property A and exactness of a group G in terms of topologically amenable actions on the Stone- Cech compactification of G. We will use a version of the characterization from [10] as our definition of exactness. Definition 3.1. A finitely generated group G is exact if for every ε > 0 there exists an element ξ ∈ ℓu(G) such that (a) ξ is finitely supported; that is, ξg = 0 for all but finitely many g ∈ G, (b) ξ is an C -valued probability measure; that is, ξg ≥ 0 for all g ∈ G and Pg∈G ξg = 1G, and (c) ξ is ε-invariant; that is, kξ − s ⋆ ξku ≤ ε for every generator s ∈ S . Exactness has numerous consequences in the theory of C∗-algebras, index the- ory and geometric group theory. In particular, Yu proved that if G is has property A or, equivalently, is exact, then the coarse Baum-Connes conjecture holds for G [30]. This on the other hand implies the Novikov conjecture for G, the zero-in-the- spectrum conjecture and has applications to the positive scalar curvature problem. More recently exactness was related to isoperimetric inequalities on finitely gen- erated groups and quantitative invariants like decay of the heat kernel and type of asymptotic dimension [17, 20]. Exact groups constitute a very large class of groups. Most notably it includes all amenable groups, hyperbolic groups (both in [30]), linear groups [7]. We refer to [28] for a more complete list. The task of constructing a group which is not exact turns out to be a difficult one. At present only one family of examples is known, Gromov's random groups [9]. The question how to find new examples of groups which would not be exact is still open. 3.2. Invariant expectations in L(L(ℓu(G), C ), C ). A Banach space X is said to be a bounded left Banach G-module if there is a homomorphism Φ : G → L(X, X) 8 such that RONALD G. DOUGLAS AND PIOTR W. NOWAK kΦ(g)kL(X,X) < ∞. sup g∈G If X is a left G-module, then we will denote the action of g ∈ G by gx for x ∈ X. If X is a left G-module then L(X, C ) is a bounded left G-module with the left action of G given by pre- and post-composing with the actions of G: (5) for T ∈ L(X, C ), x ∈ X and g ∈ G. (g · T )(x) = g ∗(cid:16)T (g−1 x)(cid:17) , Lemma 3.2. The above action is weak-* continuous. Proof. If T = C − limβ Tβ, then w∗ − lim β g · Tβ(x) = w∗ − lim β g ∗ (Tβ(g−1x)) = g ∗ (w∗ − lim β Tβ(g−1x)) = g · T (x), where the second equality follows from weak-* continuity of the action on C . (cid:3) Observe that given f ∈ ℓ∞(G, C ) the pairing hξ, f iC for ξ ∈ ℓu(G) gives naturally an operator in L(ℓu(G), C ). Moreover, Lemma 3.3. k f kℓ∞(G,C ) = k f kL. In particular, the space ℓ∞(G, C ) is isometrically embedded in L(ℓu(G), C ). Proof. The estimate k f kL ≤ k f kℓ∞(G,C ) follows easily. To see the converse observe that for every ε > 0 there is g ∈ G such that k f kℓ∞(G,C ) ≤ k fgkC + ε. Then hδg, f iC = fg and the required inequality follows by taking ε converging to 0. (cid:3) The following lemma shows that ℓ∞(G, C ) is also a G-submodule of L(ℓu(G), C ). Lemma 3.4. The actions · and ⊙ agree on ℓ∞(G, C ) ⊆ L(ℓu(G), C ). Proof. By Lemma 2.2, for any f ∈ ℓ∞(G, C ) and ξ ∈ ℓu(G) we have (g · f )(ξ) = g ∗ (hg−1 ⋆ ξ, f iC ) = (g ⊙ f )(ξ). Taking ξ = δh for any h gives the equality. (cid:3) The action of G on the G-module L(L(ℓu(G), C ), C ) will now be denoted by • to distinguish it from the action · on L(ℓu(G), C ): (6) (g • Ξ)(T ) = g ∗(cid:16)Ξ(g−1 · T )(cid:17) , for Ξ ∈ L(L(ℓu(G), C ), C ) and T ∈ L(ℓu(G), C ). Lemma 3.5. The actions • and ⋆ agree on ℓu(G) ⊆ L(L(ℓu(G), C ), C ). INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 9 Proof. We have ([g ⋆ ξ)(T ) = T (g ⋆ ξ) = g ∗(cid:16)g−1 ∗ (T (g ⋆ ξ))(cid:17) = g ∗(cid:16)g−1 · T (ξ)(cid:17) = g ∗(cid:16)ξ(g−1 · T )(cid:17) = (g • ξ)(T ), for every T ∈ L(ℓu(G), C ), ξ ∈ ℓu(G). Consider the space (cid:3) W00 = (cid:8)ξ ∈ ℓu(G) : # supp ξ < ∞ and hξ, 1GiC = c1G for some c ∈ R(cid:9) and let W0 be the closure of W00 in the norm topology in ℓu(G). Definition 3.6. Define the subspace W ⊆ L(L(ℓu(G), C ), C ) to be the weak-* closure of W00. Clearly, W is a Banach subspace of L(L(ℓu(G), C ), C ). Moreover, it has a natural structure of a G-module. The above setup allows us to prove now the main theorem characterizing ex- actness. Amenable groups are known to be characterized by a Følner and Reiter conditions, which correspond to our Definition 3.1 of exactness (see [22, 23]). An- other standard definition of amenability is through the existence of an invariant mean on the group. The next definition provides a weak version of the invariant mean. Definition 3.7. Let G be a finitely generated group. An invariant expectation on G is a bounded linear operator M : L(ℓu(G), C ) → C which satisfies (a) M ∈ W, (b) M(1G) = 1G, and (c) M is G-invariant; that is, g • M = M for every g ∈ G. The importance of the notion of an invariant expectation is in its relation to exactness of groups described in Theorem 1.1, whose statement we recall from the introduction. Theorem 1.1. A finitely generated group G is exact if and only if there exists an invariant expectation on G. Proof. Consider a sequence {ξn} where ξn is obtained from the definition of exact- ness with ε = 1 n. Each ξn is an element of ℓu(G) and as such induces a continuous linear map ξn ∈ L(L(ℓu(G), C ), C ). We consider this last space with the weak-* topology described in the previous section. Since, by the Banach-Alaoglu theo- rem, the unit ball of the space L(L(ℓu(G), C ), C ) is compact with this topology, the sequence {ξn} has a convergent subnet {ξβ} and we define M = C − lim β ξβ, 10 RONALD G. DOUGLAS AND PIOTR W. NOWAK which is equivalent to M(T ) = w∗ − lim β ξβ(T ) for every T ∈ L(ℓu(G), C ), by Corollary 2.3. We will show that M is an invariant expectation on G. Clearly, M ∈ W and, in particular, since hξβ, 1GiC = 1G for every β, it follows that M(1G) = 1G. By lemmas 3.2 and 3.5 we have for T ∈ L(ℓu(G), C ) and any generator s ∈ S , (s • M)(T ) = s ∗(cid:16)M(s−1 · T )(cid:17) = s ∗ w∗ − lim β (cid:16)ξβ(s−1 · T )(cid:17)! s ∗(cid:16)ξβ(s−1 · T )(cid:17) (s • ξβ)(T ) ([s ⋆ ξ)(T ). = w∗ − lim β = w∗ − lim β = w∗ − lim β Thus (7) and we have (M − s • M) (T ) = w∗ − lim β (ξβ − [s ⋆ ξβ)(T ) k(ξβ − [s ⋆ ξβ)(T )kC ≤ kξβ − [s ⋆ ξβkLLkT kL ≤ kξβ − [s ⋆ ξβkukT kL ≤ εβkT kL. for every n. Since εβ tends to 0 this implies that the weak-* limit in (7) also is 0 for every T . This proves G-invariance of M. Conversely, let M be an invariant expectation on G. Since M is in W we can approximate it in the weak-* topology on the module L (L(ℓu(G), C ), C )) by finitely supported elements of W00. More precisely, there exists a net {ξβ} such that ξβ ∈ W00 and w∗ − lim β (ξβ − s • ξβ)(T ) = 0 in C for every T ∈ L(ℓu(G), C ) and s ∈ S . Since the actions • and ⋆ agree on ℓu(G), this is the same as (8) w∗ − lim β T (ξβ − s ⋆ ξβ) = 0 for every T ∈ L(ℓu(G), C ) and s ∈ S . Moreover, ξβ satisfy hξβ, 1GiC = cβ1G, the spaceLs∈S ℓu(G) with the norm where σ ∈Ls∈S ℓu(G), σ = ⊕s∈S σs. For each β consider the direct sum kσk⊕u = sup s∈S kσsku σβ = ⊕s∈S (cid:16)ξβ − s ⋆ ξβ(cid:17) . INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 11 where the net of real numbers {cβ} converges to 1. By passing to a cofinal subnet, which we will also denote by ξβ, we can assume that hξβ, 1GiC ≥ 1 2 1G. We will now ensure condition (c) of Definition 3.1 and construct a sequence ξ′ n of finitely supported elements in W00 with similar properties to those of ξβ and such that, additionally, kξn − s ⋆ ξnku tends to 0 uniformly for all s ∈ S . Consider From equation (8) we deduce that for each s ∈ S , the net {(σβ)s} converges in the weak topology on ℓu(G). Namely, for each generator s ∈ S , we have ϕ(ξβ − s ⋆ ξβ) −→ 0 for every linear functional ϕ ∈ ℓu(G)∗. Since the dual spaces satisfy the equality gives that the weak and strong closures of ∆ are the same and, in particular, 0 ∈ ∆. Thus we can approximate 0 by finite convex combinations of σβ in the norm n} such that for n is a finite convex combination of the {σβ} and with the (cid:16)Ls∈S ℓu(G)(cid:17)∗ = Ls∈S ℓu(G)∗, the net σβ converges in the weak topology on Ls∈S ℓu(G). Now Mazur's lemma applied to the closed convex hull ∆ of the {σβ} topology onLs∈S ℓu(G). This means that there exists a sequence {σ′ n(cid:1) which satisfies n converges strongly to 0 inLs∈S ℓu(G). There is a corresponding each n ∈ N the element σ′ property that σ′ sequence ξ′ n = ⊕s∈S (cid:0)ξ′ n − s ⋆ ξ′ kξβ − s ⋆ ξβku −→ 0. n ∈ ℓu(G) such that σ′ sup s∈S Since each ξ′ n is a finite convex combination of the {ξβ}, we have hξ′ C n, 1GiC = * kXi=1 kXi=1 kXi=1 ciξβi , 1G+ ciDξβi, 1GEC 1G! ci 1 2 = ≥ ≥ 1 2 1G, where ci ≥ 0 andP ci = 1. The elements ξ′ to W00. Thus the sequence ξ′ n are also finitely supported and belong n satisfies conditions (a) and (c) of Definition 3.1. 12 RONALD G. DOUGLAS AND PIOTR W. NOWAK We need to ensure condition (b) from Definition 3.1. To this end consider the sequence {ζn} defined as (ζn)g = Then ζn ∈ W00 and we have (ξ′ n)g Ph∈G (ξ′ n)h . hζn, 1GiC = 1G. Since we conclude that (9) for any h ∈ G. Xg∈G (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (ξ′ n)g ≥(cid:12)(cid:12)(cid:12)(cid:12)Xg∈G (ξ′ 1 2 1G 1 n)g(cid:17) h ∗(cid:16)Pg∈G (ξ′ ≤ 2 n)g(cid:12)(cid:12)(cid:12)(cid:12) ≥ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C n − ξ′ Proof. We have It remains to show that ζn satisfy the conditions of Definition 3.1 Clearly, (ζn)g ≥ for every n ∈ N. We only need to verify the approximate invariance. n(cid:13)(cid:13)(cid:13)u for every generator s ∈ S . 0 andPg∈G(ζn)g = 1G for every n ∈ N. It is also obvious that ζn is finitely supported Lemma 3.8. ks ⋆ ζn − ζnku ≤ 4(cid:13)(cid:13)(cid:13)s ⋆ ξ′ ks ⋆ ζn − ζnku =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) s ∗Pg∈G (ξ′ n)g(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)u Pg∈G (ξ′ n)g(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)u Pg∈G (ξ′ Adding a connecting term, applying the triangle inequality and (9) we obtain s ∗Pg∈G (ξ′ ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) s ⋆ ξ′ n s ⋆ ξ′ n (10) (11) n)g ξ′ n + n)g ξ′ n ξ′ n ξ′ n − (12) ≤ 2ks ⋆ ξ′ n − ξ′ − . n)g s ∗Pg∈G (ξ′ n)g(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)u Pg∈G (ξ′ ξ′ n − n)g The second summand, after cancellation, can be estimated as follows. We observe that n)g − ξ′ n s ∗Pg∈G (ξ′ n ku +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) s ∗Pg∈G (ξ′ n)g(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)u Pg∈G (ξ′ ξ′ n − n)g (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ξ′ n s ∗Pg∈G (ξ′ where the last step follow from the triangle inequality and (9). Altogether we get where the last inequality follows again from the triangle inequality. Thus the sequence ζn satisfies all three conditions of Definition 3.1 and the group (cid:3) (cid:3) INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 13 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C n)h n)h n)h −Ph∈G (ξ′ n)g(cid:17) n)g(cid:17)(cid:16)Pg∈G (ξ′ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C n)g(cid:17) s ∗(cid:16)Pg∈G (ξ′ 1 (ξ′ = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n)g Xg∈G s ∗Ph∈G (ξ′ s ∗(cid:16)Pg∈G (ξ′ = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n)h −Ph∈G (ξ′ Ph∈G (s ⋆ ξ′ n)g(cid:17) s ∗(cid:16)Pg∈G (ξ′ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n)h(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n)h −Xh∈G Xh∈G ≤ 2(cid:13)(cid:13)(cid:13)s ⋆ ξ′ n(cid:13)(cid:13)(cid:13)u, ks ⋆ ζn − ζnku ≤ 4(cid:13)(cid:13)(cid:13)s ⋆ ξ′ n(cid:13)(cid:13)(cid:13)u ≤ 4(cid:13)(cid:13)(cid:13)s ⋆ ξ′ n − ξ′ n − ξ′ (s ⋆ ξ′ (ξ′ n − ξ′ n(cid:13)(cid:13)(cid:13)u, (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)C G is exact. We will denote by M the subset of L(L(ℓu(G), C ), C ) of expectations on G, meaning elements M satisfying only conditions (a) and (b) of Definition 3.7, and by MG ⊆ M the subset of invariant expectations. We believe that, in general, MG is an infinite set. A natural question in this context is under what conditions the invariant expectation on G is unique? Remark 3.9 (Automatic positivity). The above proof establishes one additional property of the invariant expectation M constructed in the "only if" part of the proof. Namely, the restriction of M to ℓ∞(G, C ) is a positive map. Indeed, if f ∈ ℓ∞(G, C ), f ≥ 0 then hξβ, f iC ≥ 0 for every β. Since M( f ) = w∗ − lim β hξn, f iC and weak-* limits in C preserve positivity, we have M( f ) ≥ 0. However, if we start with an M that is not positive in this sense, by passing through the "if" and then the "only if" part of the proof we obtain a new invariant expectation with the positivity property. We will require one more lemma about invariant expectations for the proof of our main results. Lemma 3.10. Let G be an exact, finitely generated group and M ∈ M be a weak-* limit of a net {ξβ} of elements satisfying conditions of Definition 3.1. Let f ′ ∈ C and define f ∈ ℓ∞(G, C ) by fg = f ′ for every g ∈ G. Then M( f ) = f ′. Proof. For each h ∈ G and every β we have hξβ, f iC =Xg∈G (ξβ)g (ξβ)g fg = f ′Xg∈G = f ′, 14 RONALD G. DOUGLAS AND PIOTR W. NOWAK sinceP(ξβ)g = 1G and this property is preserved by the weak-* limit in C . 4. Bounded cohomology of exact groups (cid:3) We will now use the facts established in the previous sections to prove a vanish- ing result for bounded cohomology of exact groups. Recall that given a group G and a bounded Banach G-module E, a bounded cocycle is a map b : G → E such that supg∈G kb(g)k < ∞ and b(gh) = gb(h) + b(g) for all g, h ∈ G. Such a cocycle b is called a boundary if there exists an element φ ∈ E such that b(g) = gφ − φ for every g ∈ G. Then the bounded cohomology in degree 1 of G with coefficients in E is the defined as H1 b(G, E) = Z1(G, E)/B1(G, E), where Z1(G, E) is the space of all bounded cocycles b : G → E and B1(G, E) ⊆ Z1(G, E) is the subspace of all boundaries b : G → E. Thus H1 b(G, E) = 0 if and only if every bounded bounded cocycle b : G → E is a boundary. See [5, 26] for details in the context of Banach algebras and [14, 15] in the context of locally compact groups. The bounded cohomology groups of G are canonically isomorphic to the Hochschild cohomology groups of the Banach algebra ℓ1(G), with the same coefficients. 4.1. Hopf G-modules. Since C is a Banach algebra, for any X the space L(X, C ) carries a natural structure of a C -module. For a ∈ C and T ∈ L(X, C ) define where the multiplication on the right is in C . (aT )(x) = aT (x), Definition 4.1. A Banach G-module E is called a Hopf G-module if for some left bounded Banach G-module X it is a subspace E ⊆ L(X, C ) which is both a G- module with respect to the action of G and a C -module with respect to the above action of C . The intuition behind the notion of Hopf G-modules is that they are "large" sub- modules of L(X, C ), as the two structures are, loosely speaking, transverse to each other. Indeed, consider a G-module X. The dual space X∗ has an induced G- module structure and we consider the two natural inclusions of X∗ into L(X, C ). The first one is obtained by mapping a functional φ to φ(x)1G. The second one is obtained by mapping φ to φ(x)1e, where 1e is the Dirac delta function at the identity element. Note that neither of these inclusions gives rise to a structure of a Hopf-G-module on X∗. Indeed, X∗ is not a C -submodule under the first one X∗ and it is not a G-submodule under the second one. Theorem 1.2. Let G be a finitely generated group. If G is exact then H1 for any weak-* closed Hopf G-module. b(G, E) = 0 INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 15 Proof. Let X be a left G-module, E ⊆ L(X, C ) be as in Theorem 1.2, b : G → E be a bounded cocycle and let C = supg∈G kb(g)k. We will show that b is a boundary. Define an operator Λ : X → ℓ∞(G, C ) by setting for x ∈ X. We have Λ(x)g =(cid:2)b(g)(cid:3) (x), kb(g)(x)kC ≤ kb(g)kL kxkX ≤ CkxkX, so the map is well-defined and continuous. The group G is exact so, by Theorem 1.1, there exists an invariant expectation M on G. Moreover, M can be chosen to be a weak-* limit of a net {ξβ}, where ξβ ∈ W00, as in the proof of Theorem 1.1. Since ℓ∞(G, C ) ⊆ L(ℓu(G), C ) we define φ : X → C by φ(x) = M(Λ(x)g). The process is illustrated by the following diagram X b(g) φ ✲ ✲ C ✲ Λ M ✲ L(ℓu(G), C ) Obviously, φ ∈ L(X, C ) and we need to show that φ ∈ E. By the definition of M, for every x ∈ X we have φ(x) = w∗ − lim β hξβ, Λ(x)gi = w∗ − lim β Xg∈G (ξβ)gb(g) (x). In other words, φ = C − lim β bβ, where bβ =Pg∈G(ξβ)gb(g). Since E is a Hopf G-module, each (ξβ)g ∈ C and only finitely many of them are non-zero, and b(g) ∈ E, we deduce that bβ belongs to E. Since E is weak-* closed in L(X, C ), φ is also an element of E. For any g ∈ G and x ∈ X we have (g · φ − φ) (x) = g ∗ φ(g−1x) − φ(x) = g ∗(cid:16)M(Λ(g−1x)h)(cid:17) − M(Λ(x)h). Note that for f ∈ ℓ∞(G, C ), the invariance of M can also be written in the following way g ∗ (M( fh)) = M((g ⊙ f )h) = M(cid:16)g ∗ fg−1h(cid:17) . 16 Thus and RONALD G. DOUGLAS AND PIOTR W. NOWAK (g · φ − φ) (x) = M(cid:16)g ∗ Λ(g−1x)g−1h(cid:17) − M(Λ(x)h) g ∗ Λ(g−1x)g−1h = g ∗(cid:16)(cid:16)b(g−1h)(cid:17) (g−1 x)(cid:17) = g · b(g−1h)(x) = g ·(cid:16)g−1 · b(h) + b(g−1) · h(cid:17) (x). For cocycles we have g · b(g−1) = −b(g). Thus for every h ∈ G we have g ∗ Λ(g−1x)g−1h = (b(h) − b(g)) (x) = Λ(x)h − b(g)(x). In the above expression, b(g)(x) is independent of h. Hence after applying M, by Lemma 3.10, we have M(b(g)(x)) = b(g)(x) and (g · φ − φ) (x) = M(Λ(x)h) − b(g)(x) − M(Λ(x)h) = −b(g)(x). Finally, setting Ξ = −φ we get b(g) = g · Ξ − Ξ, so that b is a boundary as required. (cid:3) 5. Concluding remarks 5.1. Dimension reduction and higher cohomology groups. In the case when G is amenable the fact that H1 b(G, E∗) = 0 for every Banach G-module E implies that Hn b(G, E∗) = 0 for all n ≥ 1. The method used to prove this is the dimension re- duction formula in bounded cohomology, see for example [11],[14, Section 10.3], [26, Theorem 2.4.6]. Nicolas Monod communicated to us the following argument showing that a similar fact is true in the case of exactness. Proposition 5.1 (N. Monod). If G is exact then for every n ≥ 1 we have Hn b(G, E) = 0 for every weak-* closed Hopf G-module of L(X, C ), where X is a left G-module. Sketch of proof. For every module E we have (G, E) = Hn Hn+1 b(G, ΣE), b where ΣE = ℓ∞(G, E)/E. Note that ΣE is a dual space, namely it is the dual of K, the kernel of the summation map ℓ1(G, E∗) → E∗, where E∗ denotes a given predual of E. There is a natural inclusion i : ℓ∞(G, E) → ℓ∞(G, ℓ∞(eG, X∗)), where eG = G and the notation allows one to keep track of the different copies of G. We have natural isometric isomorphisms This descends to a canonical inclusion ℓ∞(G, ℓ∞(eG, X∗)) = ℓ∞(G × eG, X∗) = ℓ∞(eG, ℓ∞(G, X∗)). ΣE ⊆ ℓ∞(eG, ΣX∗) and one can verify that, under the resulting identification, the module ΣE is a Hopf G-module, with respect to ℓ∞(eG), of ℓ∞(eG, ΣX∗) ≃ L(K, ℓ∞(eG)). (cid:3) INVARIANT EXPECTATIONS AND BOUNDED COHOMOLOGY 17 5.2. Relation between M and W. It is also interesting to investigate the relation between the set of positive expectations M+ (in the sense of remark 3.9) and the module W. One possibility is that W is generated by M+ in the sense that for every Ξ ∈ W there exist M, M′ ∈ M+ and constants c, c′ ∈ [0, +∞) such that Question 5.2. Is W generated by M+ in the above sense? Ξ = cM − c′M′. References [1] J. Brodzki, G.A. Niblo, P.W. Nowak, N. Wright, Amenable actions, invariant means and bounded cohomology, preprint, arXiv:1004.0295v1 [2] J. Brodzki, G.A. Niblo, P.W. Nowak, N. Wright, A homological characterization of topo- logical amenability, preprint arXiv:1008.4154v1 . [3] J. Brodzki, G.A. Niblo, N. Wright, On a cohomological characterization of Yu's property A, preprint, arXiv:1002.5040v2 [4] N. Brown, N. Ozawa, C∗-algebras and finite-dimensional approximations.Graduate Stud- ies in Mathematics, 88. American Mathematical Society, Providence, RI, 2008. [5] H.G. Dales, P. Aiena, J. Eschmeier, K. Laursen, G.A. Willis, Introduction to Banach alge- bras, operators, and harmonic analysis. London Mathematical Society Student Texts, 57. Cambridge University Press, Cambridge, 2003. [6] R.G. Douglas, P.W.Nowak, Every finitely generated group is weakly exact, preprint. [7] E. Guentner, N. Higson, S. Weinberger, The Novikov conjecture for linear groups. Publ. Math. Inst. Hautes ´Etudes Sci. No. 101 (2005), 243 -- 268. [8] E. Guentner, J. Kaminker, Exactness and the Novikov conjecture, Topology 41 (2002), no. 2, 411 -- 418. [9] M. Gromov, Random walk in random groups. Geom. Funct. Anal. 13 (2003), no. 1, 73 -- 146. [10] N. Higson, J. Roe, Amenable group actions and the Novikov conjecture. J. Reine Angew. Math. 519 (2000), 143 -- 153. [11] B.E. Johnson, Cohomology in Banach algebras. Memoirs of the American Mathematical Society, No. 127. American Mathematical Society, Providence, R.I., 1972. [12] E. Kirchberg, S. Wassermann, Operations on continuous bundles of C∗-algebras. Math. Ann. 303 (1995), no. 4, 677 -- 697. [13] E. Kirchberg, S. Wassermann, Exact groups and continuous bundles of C∗-algebras. Math. Ann. 315 (1999), no. 2, 169 -- 203. [14] N. Monod, Continuous bounded cohomology of locally compact groups. Lecture Notes in Mathematics, 1758. Springer-Verlag, Berlin, 2001. [15] N. Monod, An invitation to bounded cohomology. International Congress of Mathemati- cians. Vol. II, 1183 -- 1211, Eur. Math. Soc., Zurich, 2006. [16] N. Monod, A note on topological amenability, preprint arXiv:1004.0199v2 [17] P.W. Nowak, On exactness and isoperimetric profiles of discrete groups. J. Funct. Anal. 243 (2007), no. 1, 323 -- 344. [18] P.W. Nowak, Coarsely embeddable metric spaces without Property A. J. Funct. Anal. 252 (2007), no. 1, 126 -- 136. [19] P.W. Nowak, G.Yu, What is. . . Property A? Notices Amer. Math. Soc. 55 (2008), no. 4, 474-475. [20] P.W. Nowak, Isoperimetry of group actions. Adv. Math. 219 (2008), no. 1, 1 -- 26. [21] N. Ozawa, Amenable actions and exactness for discrete groups, C. R. Acad. Sci. Paris Ser. I Math., 330 (2000), 691 -- 695. [22] A.L.T. Paterson, Amenability. Mathematical Surveys and Monographs, 29. American Mathematical Society, Providence, RI, 1988. 18 RONALD G. DOUGLAS AND PIOTR W. NOWAK [23] J.-P. Pier, Amenable locally compact groups. Pure and Applied Mathematics (New York). A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1984. [24] J.-P. Pier, Amenable Banach algebras. Pitman Research Notes in Mathematics Series, 172. Longman Scientific & Technical, Harlow; John Wiley & Sons, Inc., New York, 1988. [25] J. Roe, Lectures on coarse geometry, University Lecture Series, 31. American Mathemati- cal Society, Providence, RI, 2003. [26] V. Runde, Lectures on amenability. Lecture Notes in Mathematics, 1774. Springer-Verlag, Berlin, 2002. [27] J. Spakula, Uniform K-homology theory. J. Funct. Anal. 257 (2009), no. 1, 88 -- 121. [28] R. Willett, Some notes on Property A. Limits of graphs in group theory and computer science, 191 -- 281, EPFL Press, Lausanne, 2009. [29] G. Yu, Baum-Connes conjecture and coarse geometry. K-Theory 9 (1995), no. 3, 223 -- 231. [30] G. Yu, The coarse Baum-Connes conjecture for spaces which admit a uniform embedding into Hilbert space, Invent. Math. (1) 139 (2000), 201-240. Department of Mathematics, Texas A&M University, College Station, TX 77843 E-mail address: [email protected], [email protected]
1606.06745
1
1606
2016-06-21T20:02:18
Embeddings between weighted complementary local Morrey-type spaces and weighted local Morrey-type spaces
[ "math.FA" ]
In this paper embeddings between weighted complementary local Morrey-type spaces ${\,^{^{\bf c}}\!}LM_{p\theta,\omega}({\mathbb R}^n,v)$ and weighted local Morrey-type spaces $LM_{p\theta,\omega}({\mathbb R}^n,v)$ are characterized. In particular, two-sided estimates of the optimal constant $c$ in the inequality \begin{equation*} \bigg( \int_0^{\infty} \bigg( \int_{B(0,t)} f(x)^{p_2}v_2(x)\,dx \bigg)^{\frac{q_2}{p_2}} u_2(t)\,dt\bigg)^{\frac{1}{q_2}} \le c \bigg( \int_0^{\infty} \bigg( \int_{{\,^{^{\bf c}}\!}B(0,t)} f(x)^{p_1} v_1(x)\,dx\bigg)^{\frac{q_1}{p_1}} u_1(t)\,dt\bigg)^{\frac{1}{q_1}} \end{equation*} are obtained, where $p_1,\,p_2,\,q_1,\,q_2 \in (0,\infty)$, $p_2 \le q_2$ and $u_1,\,u_2$ and $v_1,\,v_2$ are weights on $(0,\infty)$ and ${\mathbb R}^n$, respectively. The proof is based on the combination of duality techniques with estimates of optimal constants of the embeddings between weighted local Morrey-type and complementary local Morrey-type spaces and weighted Lebesgue spaces, which reduce the problem to the solutions of the iterated Hardy-type inequalities.
math.FA
math
EMBEDDINGS BETWEEN WEIGHTED COMPLEMENTARY LOCAL MORREY-TYPE SPACES AND WEIGHTED LOCAL MORREY-TYPE SPACES AMIRAN GOGATISHVILI, RZA MUSTAFAYEV, AND TU GC¸ E UNVER Abstract. In this paper embeddings between weighted complementary local Morrey-type spaces cLMpθ,ω(Rn, v) and weighted local Morrey-type spaces LMpθ,ω(Rn, v) are characterized. In particular, two-sided estimates of the optimal constant c in the inequality (cid:18)Z ∞ 0 (cid:18)ZB(0,t) f (x)p2 v2(x)dx(cid:19) q2 p2 u2(t)dt(cid:19) 1 q2 ≤ c(cid:18)Z ∞ 0 (cid:18)Z cB(0,t) f (x)p1 v1(x)dx(cid:19) q1 p1 u1(t)dt(cid:19) 1 q1 are obtained, where p1, p2, q1, q2 ∈ (0, ∞), p2 ≤ q2 and u1, u2 and v1, v2 are weights on (0, ∞) and Rn, respectively. The proof is based on the combination of duality techniques with estimates of optimal constants of the embeddings between weighted local Morrey-type and complementary local Morrey-type spaces and weighted Lebesgue spaces, which reduce the problem to the solutions of the iterated Hardy-type inequalities. 1. Introduction The classical Morrey spaces Mp,λ ≡ Mp,λ(Rn), were introduced by C. Morrey in [18] in order to study regularity questions which appear in the Calculus of Variations, and defined as follows: for 0 ≤ λ ≤ n and 1 ≤ p ≤ ∞, Mp,λ :=( f ∈ Lloc λ−n p k f kLp(B(x,r)) < ∞) , p (Rn) : k f kMp,λ := sup x∈Rn, r>0 r where B(x, r) is the open ball centered at x of radius r. Note that Mp,0(Rn) = L∞(Rn) and Mp,n(Rn) = Lp(Rn). These spaces describe local regularity more precisely than Lebesgue spaces and appeared to be quite useful in the study of the local behavior of solutions to partial differential equations, a priori estimates and other topics in PDE (cf. [13]). The classical Morrey spaces were widely investigated during the last decades, including the study of classical operators of Harmonic and Real Analysis - maximal, singular and potential operators - in generalizations of these spaces (so-called Morrey-type spaces). The local Morrey-type spaces and the complementary local Morrey-type spaces introduced by Guliyev in his doctoral thesis [16]. The local Morrey-type spaces LMpθ,ω and the complementary local Morrey-type spaces cLMpθ,ω were inten- sively studied during the last decades. The research mainly includes the study of the boundedness of classical operators in these spaces (see, for instance, [2 -- 10]), and the investigation of the functional-analytic properties of them and relation of these spaces with other known function spaces (see, for instance, [1, 11, 19]). We refer the reader to the surveys [2] and [3] for a comprehensive discussion of the history of LMpθ,ω and cLMpθ,ω. Let A be any measurable subset of Rn, n ≥ 1. By M(A) we denote the set of all measurable functions on A. The symbol M+(A) stands for the collection of all f ∈ M(A) which are non-negative on A. The family of all weight functions (also called just weights) on A, that is, measurable, positive and finite a.e. on A, is given by W(A). For p ∈ (0, ∞], we define the functional k · kp,A on M(A) by k f kp,A :=(cid:26) (cid:16)RA f (x)p dx(cid:17)1/p esssupA f (x) if if p < ∞ p = ∞ . If w ∈ W(A), then the weighted Lebesgue space Lp(w, A) is given by When A = Rn, we often write simply Lp,w and Lp(w) instead of Lp,w(A) and Lp(w, A), respectively. Lp(w, A) ≡ Lp,w(A) := { f ∈ M(A) : k f kp,w,A := k f wkp,A < ∞}. 2010 Mathematics Subject Classification. Primary 46E30; Secondary 26D10. Key words and phrases. local Morrey-type spaces, embeddings, iterated Hardy inequalities. 1 2 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Throughout the paper, we always denote by c and C a positive constant, which is independent of main parameters but it may vary from line to line. However a constant with subscript such as c1 does not change in different occurrences. By a . b, (b & a) we mean that a ≤ λb, where λ > 0 depends on inessential parameters. If a . b and b . a, we write a ≈ b and say that a and b are equivalent. We will denote by 1 the function 1(x) = 1, x ∈ R. Given two quasi-normed vector spaces X and Y, we write X = Y if X and Y are equal in the algebraic and the topological sense (their quasi-norms are equivalent). The symbol X ֒→ Y (Y ←֓ X) means that X ⊂ Y and the natural embedding I of X in Y is continuous, that is, there exist a constant c > 0 such that kzkY ≤ ckzkX for all z ∈ X. The best constant of the embedding X ֒→ Y is kI kX→Y. The weighted local Morrey-type spaces LMpθ,ω(Rn, v) and weighted complementary local Morrey-type spaces LMpθ,ω(Rn, v) are defined as follows: Let 0 < p, θ ≤ ∞. Assume that ω ∈ M+(0, ∞) and v ∈ W(Rn). where and where LMpθ,ω(Rn, v) :=(cid:26) f ∈ Lloc p,v(Rn) : k f kLMpθ,ω(Rn,v) < ∞(cid:27), cLMpθ,ω(Rn, v) :=(cid:26) f ∈ ∩t>0Lp,v( cB(0, t)) : k f k cLMpθ,ω(Rn,v) < ∞(cid:27), k f kLMpθ,ω(Rn,v) :=(cid:13)(cid:13)(cid:13)k f kp,v,B(0,r)(cid:13)(cid:13)(cid:13)θ,ω,(0,∞), cB(0,r)(cid:13)(cid:13)(cid:13)θ,ω,(0,∞). k f k cLMpθ,ω(Rn,v) :=(cid:13)(cid:13)(cid:13)k f kp,v, Remark 1.1. In [5] and [7] it were proved that the spaces LMpθ,ω(Rn) := LMpθ,ω(Rn,1) and cLMpθ,ω(Rn) := cLMpθ,ω(Rn,1) are non-trivial, i.e. consists not only of functions equivalent to 0 on Rn, if and only if (1.1) and (1.2) respectively. The same conclusion is true for LMpθ,ω(Rn, v) and cLMpθ,ω(Rn, v) for any v ∈ W(Rn). kωkθ,(t,∞) < ∞, kωkθ,(0,t) < ∞, for some for some t > 0, t > 0, The proof of the following statement is straightforward. Lemma 1.2. (i) If kωkθ,(t1,∞) = ∞ for some t1 > 0, then f ∈ LMpθ,ω(Rn, v) ⇒ f = 0 a.e. on B(0, t1). (ii) If kωkθ,(0,t2) = ∞ for some t2 > 0, then f ∈ cLMpθ,ω(Rn, v) ⇒ f = 0 a.e. on cB(0, t2). Let 0 < θ ≤ ∞. We denote by c Ωθ : =(cid:8)ω ∈ M+(0, ∞) : 0 < kωkθ,(t,∞) < ∞, t > 0(cid:9), Ωθ : =(cid:8)ω ∈ M+(0, ∞) : 0 < kωkθ,(0,t) < ∞, t > 0(cid:9). Let v ∈ W(Rn). It is easy to see that LMpθ,ω(Rn, v) and cLMpθ,ω(Rn, v) are quasi-normed vector spaces when ω ∈ Ωθ and ω ∈ c Ωθ, respectively. The following statements are immediate consequences of Fubini's Theorem and were observed in [5] and [7], for v = 1, respectively. Lemma 1.3. Let 0 < p ≤ ∞ and v ∈ W(Rn). Then (i) LMpp,ω(Rn, v) = Lp(w), where w(x) := v(x)kωkp,(x,∞), x ∈ Rn. (ii) cLMpp,ω(Rn, v) = Lp(w), where w(x) := v(x)kωkp,(0,x), x ∈ Rn. Recall that the embeddings between weighted local Morrey-type spaces and weighted Lebesgue spaces, that is, the embeddings (1.3) (1.4) Lp1(v1) ֒→ LMp2θ,ω(Rn, v2), cLM p2θ,ω(Rn, v2), Lp1(v1) ֒→ EMBEDDINGS BETWEEN cLMpθ,Ω(Rn, v) AND LMpθ,Ω(Rn, v) 3 (1.5) (1.6) Lp1(v1) ←֓ LMp2θ,ω(Rn, v2), cLM p2θ,ω(Rn, v2) Lp1(v1) ←֓ are completely characterized in [19]. Our principal goal in this paper is to investigate the embeddings between weighted complementary local Morrey- type spaces and weighted local Morrey type spaces and vice versa, that is, the embeddings (1.7) (1.8) cLMp1θ1,ω1(Rn, v1) ֒→ LMp2θ2,ω2(Rn, v2), cLMp2θ2,ω2(Rn, v2). LMp1θ1,ω1(Rn, v1) ֒→ An approach used in this paper consist of a duality argument combined with estimates of optimal constants of embeddings (1.3) - (1.6), which reduce the problem to the solutions of the iterated Hardy-type inequalities (1.9) with (cid:13)(cid:13)(cid:13)kH∗ f kp,u,(0,·)(cid:13)(cid:13)(cid:13)q,w,(0,∞) ≤ c k f kθ,v,(0,∞), f ∈ M+(0, ∞), (H∗ f )(t) :=Z ∞ t f (τ)dτ, t > 0, where u, v, w are weights on (0, ∞) and 0 < p, q ≤ ∞, 1 < θ < ∞. There exists different solutions of these inequalities. We will use characterizations from [14] and [15]. Note that in view of Lemma 1.3, embeddings (1.7) - (1.8) contain embeddings (1.3) - (1.6) as a special case. Moreover, by the change of variables x = y/y2 and t = 1/τ, it is easy to see that (1.8) is equivalent to the embedding cLMp1θ1, ω1(Rn, v1) ֒→ LMp2θ2, ω2(Rn, v2), where vi(y) = vi(y/y2)y−2n/pi and ωi(τ) = τ−2/θi ωi(cid:0)1/τ(cid:1), i = 1, 2. This note allows us to concentrate our attention on characterization of (1.7). On the negative side of things we have to admit that the duality approach works only in the case when, in (1.7) - (1.8), one has p2 ≤ θ2. Unfortunately, in the case when p2 > θ2 the characterization of these embeddings remains open. In particular, we obtain two-sided estimates of the optimal constant c in the inequality (cid:18)Z ∞ 0 (cid:18)ZB(0,t) f (x)p2v2(x)dx(cid:19) q2 p2 u2(t)dt(cid:19) 1 q2 ≤ c(cid:18)Z ∞ 0 (cid:18)Z cB(0,t) f (x)p1v1(x)dx(cid:19) q1 p1 u1(t)dt(cid:19) 1 q1 , where p1, p2, q1, q2 ∈ (0, ∞), p2 ≤ q2 and u1, u2 and v1, v2 are weights on (0, ∞) and Rn, respectively. The paper is organized as follows. We start with formulations of our main results in Section 2. The proofs of the main results are presented in Section 3. 2. Statement of the main results We adopt the following usual conventions. Convention 2.1. (i) Throughout the paper we put 0/0 = 0, 0 · (±∞) = 0 and 1/(±∞) = 0. (ii) We put p if 1−p ∞ if p if p−1 if 1 0 < p < 1, p = 1, 1 < p < ∞, p = ∞. p′ := (iii) To state our results we use the notation p → q for 0 < p, q ≤ ∞ defined by 1 p → q = 1 q − 1 p if q < p, and p → q = ∞ if q ≥ p. (iv) If I = (a, b) ⊆ R and g is a monotone function on I, then by g(a) and g(b) we mean the limits limt→a+ g(t) and limt→b− g(t), respectively. 4 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Our main results are the following theorems. Throughout the paper we will denote Theorem 2.2. Let 0 < θ2 = p2 ≤ p1 = θ1 < ∞. Assume that v1, v2 ∈ W(Rn), ω1 ∈ Ωθ1 and ω2 ∈ Ωθ2. Then 1 v2,Rn . Theorem 2.3. Let 0 < p1, p2, θ1, θ2 < ∞ and θ2 , p2 ≤ p1 = θ1. Assume that v1, v2 ∈ W(Rn), ω1 ∈ c Ωθ1 and ω2 ∈ Ωθ2. (t > 0, x > 0). kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ sup c eV(t) eV(t) +eV(x) and V(t, x) := 1 v2kp1→p2,B(0,x), p1,(0,·) kω2kp2,(·,∞)(cid:13)(cid:13)(cid:13)p1→p2,v−1 p1,(0,·)(cid:13)(cid:13)(cid:13)p1→p2,v−1 eV(x) := kv−1 kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈(cid:13)(cid:13)(cid:13)kω1k−1 t∈(0,∞)(cid:13)(cid:13)(cid:13)kω1k−1 0 (cid:13)(cid:13)(cid:13)kω1k−1 p1,(0,·)(cid:13)(cid:13)(cid:13) θ1,(0,t)(cid:13)(cid:13)(cid:13)kω2kp2,(·,∞)(cid:13)(cid:13)(cid:13)p1→p2,v−1 1 v2,B(0,t) d(cid:18) − kω2k p1→θ2 p1→p2,v−1 kω1k−1 kI k cLMp1 θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ sup t∈(0,∞) kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈(cid:18)Z ∞ 1 v2,B(0,t)kω2kθ2,(t,∞); p1→θ2 θ2,(t,∞)(cid:19)(cid:19) 1 p1→θ2 . 1 v2,B(0,t); (i) If p1 ≤ θ2, then (ii) If θ2 < p1, then (i) If θ1 ≤ p2, then (ii) If p2 < θ1, then Theorem 2.4. Let 0 < p1, p2, θ1, θ2 < ∞ and θ2 = p2 ≤ p1 , θ1. Assume that v1, v2 ∈ W(Rn), ω1 ∈ c Ωθ1 and ω2 ∈ Ωθ2. kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈(cid:18)Z ∞ p1→p2,v−1 0 (cid:13)(cid:13)(cid:13)kω2kp2,(·,∞)(cid:13)(cid:13)(cid:13)θ1→p2 θ1,(0,∞)(cid:13)(cid:13)(cid:13)kω2kp2,(·,∞)(cid:13)(cid:13)(cid:13)p1→p2,v−1 + kω1k−1 1 v2,Rn . θ1,(0,t) (cid:19)(cid:19) 1 v2,B(0,t)d(cid:18) − kω1k−θ1→p2 1 θ1→p2 In view of Lemma 1.3, Theorems 2.2 - 2.4 are straightforward consequences of [19, Theorem 3.1] and [19, Theorem 4.2]. To state further results we need the following definitions. Definition 2.5. Let U be a continuous, strictly increasing function on [0, ∞) such that U(0) = 0 and limt→∞ U(t) = ∞. Then we say that U is admissible. Let U be an admissible function. We say that a function ϕ is U-quasiconcave if ϕ is equivalent to an increasing function on (0, ∞) and ϕ/U is equivalent to a decreasing function on (0, ∞). We say that a U-quasiconcave function ϕ is non-degenerate if U(t) ϕ(t) The family of non-degenerate U-quasiconcave functions is denoted by QU. ϕ(t) = lim t→∞ = lim t→0+ ϕ(t) U(t) = lim t→∞ 1 ϕ(t) lim t→0+ = 0. Definition 2.6. Let U be an admissible function, and let w be a non-negative measurable function on (0, ∞). We say that the function ϕ, defined by ϕ(t) = U(t)Z ∞ 0 w(τ)dτ U(τ) + U(t) , t ∈ (0, ∞), is a fundamental function of w with respect to U. One will also say that w(τ)dτ is a representation measure of ϕ with respect to U. Remark 2.7. Let ϕ be the fundamental function of w with respect to U. Assume that Z ∞ 0 w(τ)dτ U(τ) + U(t) < ∞, t > 0, Z 1 0 w(τ)dτ U(τ) =Z ∞ 1 w(τ)dτ = ∞. Then ϕ ∈ QU. EMBEDDINGS BETWEEN cLMpθ,Ω(Rn, v) AND LMpθ,Ω(Rn, v) 5 Remark 2.8. Suppose that ϕ(x) < ∞ for all x ∈ (0, ∞), where ϕ is defined by If ϕ(x) = esssup t∈(0,x) U(t)esssup τ∈(t,∞) w(τ) U(τ) , t ∈ (0, ∞). lim sup t→0+ w(t) = lim sup t→+∞ 1 w(t) = lim sup t→0+ U(t) w(t) = lim sup t→+∞ w(t) U(t) = 0, then ϕ ∈ QU. Theorem 2.9. Let 0 < p1, p2, θ1, θ2 < ∞, p2 < p1, θ1 ≤ p2 < θ2. Assume that v1, v2 ∈ W(Rn), ω1 ∈ Suppose thateV is admissible and (i) If p1 ≤ θ2, then ϕ1(x) := sup t∈(0,∞)eV(t) V(x, t) kω1k−1 θ1,(0,t) ∈ Q 1 . p1→p2 eV V(t, x) kω2kθ2,(t,∞). kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) ϕ1(x) sup t∈(0,∞) (ii) If θ2 < p1, then kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) ϕ1(x)(cid:18)Z ∞ 0 V(t, x)p1→θ2 d(cid:18) − kω2k p1→θ2 θ2,(t,∞)(cid:19)(cid:19) c Ωθ1 and ω2 ∈ Ωθ2. 1 p1→θ2 . Theorem 2.10. Let 0 < p1, p2, θ1, θ2 < ∞, p2 < p1 and p2 < min{θ1, θ2}. Assume that v1, v2 ∈ W(Rn), ω1 ∈ c Ωθ1 and ω2 ∈ Ωθ2. Suppose thateV is admissible and ϕ2(x) :=(cid:18)Z ∞ 0 (i) If max{p1, θ1} ≤ θ2, then θ1,(0,t) (cid:19)(cid:19) [eV(t)V(x, t)]θ1→p2 d(cid:18) − kω1k−θ1→p2 1 θ1→p2 ∈ Q . 1 p1→p2 eV kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) + kω1k−1 ϕ2(x) sup t∈(0,∞) θ1,(0,∞) sup V(t, x)kω2kθ2,(t,∞) t∈(0,∞)eV(t)kω2kθ2,(t,∞); ≈(cid:18)Z ∞ 0 ϕ2(x) θ1 →θ2·θ1→p2 θ2→p2 eV(x)θ1→p2(cid:18)Z ∞ 0 eV(t)p1→θ2d(cid:18) − kω2k 0 θ1,(0,∞)(cid:18)Z ∞ V(t, x)p1→θ2d(cid:18) − kω2k θ2,(t,∞)(cid:19)(cid:19) p1→θ2 p1→θ2 . 1 + kω1k−1 p1→θ2 θ2,(t,∞)(cid:19)(cid:19) θ1→θ2 θ1,(0,x) (cid:19)(cid:19) p1→θ2 d(cid:18) − kω1k−θ1→p2 1 θ1→θ2 (ii) If p1 ≤ θ2 < θ1, then kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈(cid:18)Z ∞ 0 ϕ2(x) + kω1k−1 θ1,(0,∞) sup θ1→θ2·θ1 →p2 θ2→p2 eV(x)θ1→p2(cid:18) sup t∈(0,∞)eV(t)kω2kθ2,(t,∞); t∈(0,∞) (iii) If θ1 ≤ θ2 < p1, then kI k cLMp1 θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) (iv) If θ2 < min{p1, θ1}, then kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) V(t, x)kω2kθ2,(t,∞)(cid:19)θ1→θ2 θ1,(0,x) (cid:19)(cid:19) d(cid:18) − kω1k−θ1→p2 1 θ1→θ2 0 ϕ2(x)(cid:18)Z ∞ θ1,(0,∞)(cid:18)Z ∞ V(t, x)p1→θ2d(cid:18) − kω2k 0 eV(t)p1→θ2d(cid:18) − kω2k p1→θ2 θ2,(t,∞)(cid:19)(cid:19) θ2,(t,∞)(cid:19)(cid:19) p1→θ2 1 p1→θ2 1 p1→θ2 ; + kω1k−1 6 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Theorem 2.11. Let 0 < θ1 < p = p1 = p2 < θ2 < ∞. Assume that v1, v2 ∈ W(Rn) ∩ C(Rn), ω1 ∈ 1 v2,B(0,t)kω2kθ2,(t,∞). kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup c Ωθ1 and ω2 ∈ Ωθ2. t∈(0,∞)(cid:13)(cid:13)(cid:13)kω1k−1 θ1,(0,·)(cid:13)(cid:13)(cid:13)∞,v−1 Theorem 2.12. Let 0 < θ1, θ2 < ∞ and 0 < p = p1 = p2 < min{θ1, θ2}. Assume that v1, v2 ∈ W(Rn) such that v−1 1 v2 ∈ C(Rn). Suppose that ω1 ∈ c Ωθ1, ω2 ∈ Ωθ2 and 0 < kω−1 2 kθ2→p,(x,∞) < ∞ holds for all x > 0. (i) If θ1 ≤ θ2, then kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞)(cid:18)eV(x)θ1→pZ ∞ x θ1,(0,∞) sup + kω1k−1 d(cid:18) − kω1k−θ1→p t∈(0,∞)eV(t)kω2kθ2,(t,∞); θ1,(0,t)(cid:19) +Z x (ii) If θ2 < θ1, then θ1,(0,t)(cid:19)(cid:19) 1 0 eV(t)θ1→p d(cid:18) − kω1k−θ1→p θ1→p kω2kθ2,(x,∞) x 0 kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈(cid:18)Z ∞ (cid:18)Z ∞ 0<τ≤xeV(τ)kω2kθ2,(τ,∞)(cid:19)θ1→θ2 θ2→p(cid:18) sup θ1,(0,t)(cid:19)(cid:19) θ1→θ2 d(cid:18) − kω1k−θ1→p +(cid:18)Z ∞ (cid:18)Z x θ1,(0,t)(cid:19)(cid:19) θ1→θ2 0 eV(t)θ1→pd(cid:18) − kω1k−θ1→p θ2→peV(x)θ1→pkω2kθ1→θ2 t∈(0,∞)eV(t)kω2kθ2,(t,∞). θ1,(0,∞) sup + kω1k−1 0 3. Proofs of main results θ1→θ2 1 θ1,(0,x)(cid:19)(cid:19) d(cid:18) − kω1k−θ1→p θ1,(0,x)(cid:19)(cid:19) θ2,(t,∞)d(cid:18) − kω1k−θ1→p 1 θ1→θ2 Before proceeding to the proof of our main results we recall the following integration in polar coordinates formula. We denote the unit sphere {x ∈ Rn : x = 1} in Rn by S n−1. If x ∈ Rn\{0}, the polar coordinates of x are r = x ∈ (0, ∞), x′ = x x ∈ S n−1. There is a unique Borel measure σ = σn−1 on S n−1 such that if f is Borel measurable on Rn and f ≥ 0 or f ∈ L1(Rn), then ZRn f (x)dx =Z ∞ 0 ZS n−1 f (rx′)rn−1dσ(x′)dr (see, for instance, [12, p. 78]). Lemma 3.1. Let 0 < p1, p2, θ1, θ2 ≤ ∞ and p1 < p2. Assume that v1, v2 ∈ W(Rn), ω1 ∈ cLMp1θ1,ω1(Rn, v1) 6֒→ LMp2θ2,ω2(Rn, v2). Proof. Assume that cLMp1θ1,ω1(Rn, v1) ֒→ LMp2θ2,ω2(Rn, v2) holds. Then there exist c > 0 such that c Ωθ1 and ω2 ∈ Ωθ2. Then holds for all f ∈ M+(Rn). Let τ ∈ (0, ∞) and f ∈ M(Rn): supp f ⊂ B(0, τ). It is easy to see that (3.1) and k f kLMp2 θ2,ω2 (Rn,v2) ≤ c k f k cLMp1θ1,ω1 (Rn,v1) k f kLMp2 θ2,ω2 (Rn,v2) =(cid:13)(cid:13)(cid:13)k f kp2,v2,B(0,t)(cid:13)(cid:13)(cid:13)θ2,ω2,(0,∞) ≥(cid:13)(cid:13)(cid:13)k f kp2,v2,B(0,t)(cid:13)(cid:13)(cid:13)θ2,ω2,(τ,∞) cB(0,t)(cid:13)(cid:13)(cid:13)θ1,ω1,(0,∞) k f k cLMp1θ1,ω1 (Rn,v1) =(cid:13)(cid:13)(cid:13)k f kp1 ,v1, ≥ kω2kθ2,(τ,∞) k f kp2,v2,B(0,τ) EMBEDDINGS BETWEEN cLMpθ,Ω(Rn, v) AND LMpθ,Ω(Rn, v) (3.2) Combining (3.1) with (3.2), we can assert that =(cid:13)(cid:13)(cid:13)k f kp1 ,v1, cB(0,t)(cid:13)(cid:13)(cid:13)θ1,ω1,(0,τ) ≤ kω1kθ1,(0,τ) k f kp1,v1,B(0,τ). Since ω1 ∈ c Ωθ1 and ω2 ∈ Ωθ2, we conclude that Lp1(B(0, τ), v1) ֒→ Lp2(B(0, τ), v2), which is a contradiction. kω2kθ2,(τ,∞) k f kp2 ,v2,B(0,τ) ≤ c kω1kθ1,(0,τ) k f kp1 ,v1,B(0,τ). 7 (cid:3) The following lemma is true. Lemma 3.2. Let 0 < p1, p2, θ1, θ2 < ∞, p2 ≤ p1 and p2 < θ2. Assume that v1, v2 ∈ W(Rn), ω1 ∈ Then c Ωθ1 and ω2 ∈ Ωθ2. kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) = kI k sup g∈M+(0,∞) Proof. By duality, interchanging suprema, we have that p2 cLMp1 θ1,ω1 (Rn,v1)→Lp2(cid:0)v2(·)H∗g(·) ,(0,∞) kgk θ2 θ2−p2 ,ω −p2 2 1 p2(cid:1) 1 p2 .  kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) = sup f ∈M+(Rn) k f kLMp2 θ2,ω2 (Rn,v2) k f k cLMp1θ1,ω1 (Rn,v1) = sup f ∈M+(Rn) 1 k f k cLMp1θ1,ω1 (Rn,v1) sup g∈M+(0,∞) = sup g∈M+(0,∞) 1 kgk 1 p2 θ2 θ2−p2 ,ω −p2 2 ,(0,∞) sup f ∈M+(Rn) (cid:18)Z ∞ 0 (cid:18)Z ∞ 0 (cid:18)ZB(0,τ) f (x)p2v2(x)p2 dx(cid:19)g(τ)dτ(cid:19) 1 p2 kgk 1 p2 θ2 θ2−p2 ,ω −p2 2 ,(0,∞) (cid:18)ZB(0,τ) f (x)p2 v2(x)p2 dx(cid:19)g(τ)dτ(cid:19) 1 k f k cLMp1θ1,ω1 (Rn,v1) p2 . Applying Fubini's Theorem, we get that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) (3.3) = sup g∈M+(0,∞) = sup g∈M+(0,∞) kgk kgk 1 p2 θ2 θ2−p2 1 p2 θ2 θ2−p2 1 ,ω 1 (cid:18)ZRn f (x)p2v2(x)p2(cid:18)Z ∞ x g(τ)dτ(cid:19) dx(cid:19) 1 p2 k f k cLMp1 θ1,ω1 (Rn,v1) sup f ∈M+(Rn) −p2 2 ,(0,∞) ,ω −p2 2 ,(0,∞) kI k cLMp1θ1,ω1 (Rn,v1)→Lp2(cid:0)v2(·)H∗g(·) 1 p2(cid:1). (cid:3) Proof of Theorem 2.9. By Lemma 3.2, we have that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) = sup g∈M+(0,∞) 1 kgk 1 p2 θ2 θ2−p2 ,ω −p2 2 ,(0,∞) kI k cLMp1θ1,ω1 (Rn,v1)→Lp2(cid:0)v2(·)H∗g(·) 1 p2(cid:1). 8 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Since θ1 ≤ p2, applying [19, Theorem 4.2, (a)], we obtain that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ Using polar coordinates, we have that sup g∈M+(0,∞) sup t∈(0,∞) kω1k−p2 θ1,(0,t)kH∗g( · )k p1 p1−p2 ,(v−1 1 v2)p2 ,B(0,t) kgk θ2 θ2−p2 ,ω −p2 2 ,(0,∞) 1 p2 .  kH∗g( · )k p1 p1−p2 1 v2)p2 ,B(0,t) = kH∗gk p1 ,(v−1 p1−p2 p1−p2 p1 ,v ,(0,t) , t > 0, v(r) :=ZS n−1 (v−1 1 v2)(rx′) p1 p2 p1−p2 rn−1dσ(x′), r > 0. where Thus, we obtain that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ Taking into account that sup t∈(0,∞) sup g∈M+(0,∞) kω1k−p2 θ1,(0,t)(cid:13)(cid:13)(cid:13)H∗g(cid:13)(cid:13)(cid:13) p1 p1−p2 ,ω −p2 2 ,(0,∞) kgk θ2 θ2−p2 p1−p2 p1 ,v ,(0,t) 1 p2 .  (3.4) Z t 0 v(r)dr =Z t 0 ZS n−1 =ZB(0,t) (v−1 1 v2) (v−1 1 v2)(rx′) p1 p2 p1−p2 dσ(x′)rn−1dr p1 p2 p1−p2 (x)dx =eV(t) p1 p2 p1−p2 , (i) if p1 ≤ θ2, then applying [14, Theorem 3.2, (i)], we arrive at kI k cLMp1 θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) ϕ1(x) sup t∈(0,∞) V(t, x)kω2kθ2,(t,∞); (ii) if θ2 < p1, then applying [14, Theorem 3.2, (ii)], we arrive at kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup x∈(0,∞) ϕ1(x)(cid:18)Z ∞ 0 The proof is completed. Remark 3.3. In view of Remark 2.8, if V(t, x)p1→θ2d(cid:18) − kω2k p1→θ2 θ2,(t,∞)(cid:19)(cid:19) 1 p1→θ2 . (cid:3) lim sup t→0+ eV(t)kω1k−1 θ1,(0,t) = lim sup t→+∞ eV(t)kω1kθ1,(0,t) = lim sup t→0+ kω1kθ1,(0,t) = lim sup t→+∞ kω1k−1 θ1,(0,t) = 0, . 1 p1→p2 Proof of Theorem 2.10. By Lemma 3.2, applying [19, Theorem 4.2, (c)], we have that then ϕ1 ∈ Q eV kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ kω1k−1 g∈M+(0,∞) sup θ1,(0,∞) (cid:18)Z ∞ 0 sup + g∈M+(0,∞) kH∗g( · )k p1 p1−p2 ,(v−1 1 v2)p2 ,Rn kgk θ2 θ2−p2 ,ω −p2 2 ,(0,∞) kH∗g( · )k θ1−p2 θ1 p1 p1−p2 1 v2)p2 ,B(0,t) ,(v−1 kgk θ2 θ2−p2 ,ω −p2 2 ,(0,∞) 1 p2  d(cid:18) − kω1k  − θ1 p2 θ1−p2 θ1,(0,t)(cid:19)(cid:19) θ1−p2 θ1 1 p2 .  EMBEDDINGS BETWEEN cLMpθ,Ω(Rn, v) AND LMpθ,Ω(Rn, v) 9 Using polar coordinates, we have that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ kω1k−1 g∈M+(0,∞) sup θ1,(0,∞) (cid:18)Z ∞ 0 kH∗gk kH∗gk p1 p1−p2 p1−p2 p1 ,v ,(0,∞) kgk θ2 θ2−p2 ,ω −p2 2 ,(0,∞) 1 p2  d(cid:18) − kω1k − θ1 p2 θ1−p2 θ1,(0,t)(cid:19)(cid:19) θ1−p2 θ1 ,(0,t) θ1 θ1−p2 p1 p1−p2 p1−p2 p1 ,v kgk θ2 θ2−p2 ,ω −p2 2 ,(0,∞)  + sup g∈M+(0,∞) := C1 + C2. 1 p2  Assume first that p1 ≤ θ2. On using the characterization of the boundedness of the operator H∗ in weighted Lebesgue spaces (see, for instance, [17, 20]), we arrive at (i) Let θ1 ≤ θ2. Applying [14, Theorem 3.1, (i)], we obtain that C1 ≈ kω1k−1 θ1,(0,∞) sup t∈(0,∞)eV(t) kω2kθ2,(t,∞). C2 ≈ sup x∈(0,∞) ϕ2(x) sup t∈(0,∞) V(t, x) kω2kθ2,(t,∞). Consequently, the proof is completed in this case. (ii) Let θ2 < θ1. Using [14, Theorem 3.1, (ii)], we have that C2 ≈(cid:18)Z ∞ 0 ϕ2(x) θ1 →θ2·θ1→p2 θ2→p2 eV(x)θ1→p2(cid:18) sup t∈(0,∞) and the statement follows in this case. V(t, x)kω2kθ2,(t,∞)(cid:19)θ1→θ2 θ1,(0,x) (cid:19)(cid:19) d(cid:18) − kω1k−θ1→p2 1 θ1→θ2 , Let us now assume that θ2 < p1. Then, using the characterization of the boundedness of the operator H∗ in weighted Lebesgue spaces, we have that (iii) Let θ1 ≤ θ2, then [14, Theorem 3.1, (iii)] yields that C1 ≈ kω1k−1 θ1,(0,∞)(cid:18)Z ∞ C2 ≈ sup x∈(0,∞) ϕ2(x)(cid:18)Z ∞ 0 0 eV(t)p1→θ2 d(cid:18) − kω2k V(t, x)p1→θ2d(cid:18) − kω2k p1→θ2 θ2,(t,∞)(cid:19)(cid:19) 1 p1→θ2 . p1→θ2 θ2,(t,∞)(cid:19)(cid:19) 1 p1→θ2 , and these completes the proof in this case. (iv) If θ2 < θ1, then on using [14, Theorem 3.1, (iv)], we arrive at C2 ≈(cid:18)Z ∞ 0 ϕ2(x) θ1 →θ2·θ1→p2 θ2→p2 eV(x)θ1→p2(cid:18)Z ∞ 0 and in this case the proof is completed. V(t, x)p1→θ2d(cid:18) − kω2k p1→θ2 θ2,(t,∞)(cid:19)(cid:19) θ1→θ2 θ1,(0,x) (cid:19)(cid:19) p1→θ2 d(cid:18) − kω1k−θ1→p2 1 θ1→θ2 , (cid:3) Remark 3.4. Assume that ϕ2(x) < ∞, x > 0. In view of Remark 2.7, if Z 1 0 (cid:18)Z t 0 ωθ1 1 (cid:19)− θ1 θ1−p2 then ϕ2 ∈ Q . 1 p1→p2 eV 1 (t)dt =Z ∞ ωθ1 1 eV(t) θ1 p2 θ1−p2(cid:18)Z t 0 θ1 θ1−p2 ωθ1 1 (cid:19)− ωθ1 1 (t)dt = ∞, 10 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Proof of Theorem 2.11. By Lemma 3.2, applying [19, Theorem 4.2, (b)], we get that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) = sup t∈(0,∞) sup g∈M+(0,∞) kω1k−p θ1,(0,t)kH∗g( · )k∞,(v−1 1 v2)p,B(0,t) kgk θ2 θ2−p ,ω −p 2 ,(0,∞) 1 p .  Recall that, whenever F,G are non-negative measurable functions on (0, ∞) and F is non-increasing, then esssup t∈(0,∞) F(t)G(t) = esssup t∈(0,∞) F(t)esssup τ∈(0,t) G(τ). (3.5) Observe that (3.6) holds for all t > 0, where v(τ) :=(cid:0)supy=τ v−1 On using (3.5), we get that kH∗g( · )k∞,(v−1 1 v2)p,B(0,t) = sup τ∈(0,t) 1 (y)v2(y)(cid:1)pH∗g(y) = kH∗gk∞,v,(0,t) sup y=τ(cid:0)v−1 1 (y)v2(y)(cid:1)p, τ > 0. kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) = = sup g∈M+(0,∞) sup t∈(0,∞) kω1k−p θ1,(0,t)kH∗gk∞,v,(0,t) kgk θ2 θ2−p ,ω −p 2 ,(0,∞) sup g∈M+(0,∞) kH∗gk∞,kω1k−p θ1,(0,·) v(·),(0,∞) kgk θ2 θ2−p ,ω −p 2 ,(0,∞) 1 p  Using the characterization of the boundedness of H∗ in weighted Lebesgue spaces, we obtain that kI k cLMp1θ1,ω1 (Rn,v1)→LMp2θ2,ω2 (Rn,v2) ≈ sup t∈(0,∞) 1 . 1 p v(s)  p(cid:19) 1 (y)v2(y)(cid:19) θ1,(0,y)v−1 1 (x)v2(x)(cid:19) θ1,(0,x)v−1 kω1k−1 θ1,(0,s) kω1k−1 sup y=s kω1k−1 θ1,(0,·)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞,v−1 1 p kH∗gk∞,v,(0,∞) kgk θ2  ∞,v,(0,t)d(cid:18) − kω1k −p 2 ,(0,∞) θ2−p ,ω θ1 θ1−p kgk θ2 θ2−p ,ω −p 2 ,(0,∞) − θ1 p θ1−p θ1,(0,t)(cid:19)(cid:19) θ1−p θ1 1 p  = sup t∈(0,∞) = sup t∈(0,∞) = sup t∈(0,∞) s∈(0,t) s∈(0,t) x∈B(0,t) kω2kθ2,(t,∞)(cid:18) sup kω2kθ2,(t,∞)(cid:18) sup kω2kθ2,(t,∞)(cid:18) sup kω2kθ2,(t,∞)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)kω1k−1 θ1,(0,∞) (cid:18)Z ∞ kH∗gk g∈M+(0,∞) sup 0 sup  + g∈M+(0,∞) := C3 + C4. Proof of Theorem 2.12. By Lemma 3.2, applying [19, Theorem 4.2, (d)], and using (3.6), we get that 1 v2,B(0,t) . (cid:3) kI k cLMp1 θ1,ω1 (Rn,v1)→LMp2 θ2,ω2 (Rn,v2) ≈ kω1k−1 Again, using the characterization of the boundedness of H∗ in weighted Lebesgue spaces, we obtain that C3 ≈ kω1k−1 θ1,(0,∞) sup (i) Let θ1 ≤ θ2, then by [15, Theorem 4.1], we have that C4 ≈ sup x∈(0,∞)(cid:18)eV(x)θ1→pZ ∞ x d(cid:18) − kω1k−θ1→p θ1,(0,t)(cid:19) +Z x t∈(0,∞)eV(t)kω2kθ2,(t,∞). θ1,(0,t)(cid:19)(cid:19) 1 0 eV(t)θ1→p d(cid:18) − kω1k−θ1→p θ1→p kω2kθ2,(x,∞), EMBEDDINGS BETWEEN cLMpθ,Ω(Rn, v) AND LMpθ,Ω(Rn, v) and the statement follows in this case. (ii) Let θ2 < θ1, then [15, Theorem 4.4] yields that C4 ≈(cid:18)Z ∞ +(cid:18)Z ∞ 0 0 x (cid:18)Z ∞ 0<τ≤xeV(τ)kω2kθ2,(τ,∞)(cid:19)θ1→θ2 θ2→p(cid:18) sup θ1,(0,t)(cid:19)(cid:19) θ1→θ2 d(cid:18) − kω1k−θ1→p (cid:18)Z x θ1,(0,t)(cid:19)(cid:19) θ1→θ2 0 eV(t)θ1→pd(cid:18) − kω1k−θ1→p θ2→peV(x)θ1→pkω2kθ1→θ2 θ1,(0,x)(cid:19)(cid:19) d(cid:18) − kω1k−θ1→p θ1,(0,x)(cid:19)(cid:19) θ2,(t,∞)d(cid:18) − kω1k−θ1→p and the proof is completed in this case. 1 θ1→θ2 1 θ1→θ2 , 11 (cid:3) Acknowledgments. The research of A. Gogatishvili was partially supported by the grant P201-13-14743S of the Grant Agency of the Czech Republic and RVO: 67985840 and by Shota Rustaveli National Science Foun- dation grants no. DI/9/5-100/13 (Function spaces, weighted inequalities for integral operators and problems of summability of Fourier series). References [1] Ts. Batbold and Y. Sawano, Decompositions for local Morrey spaces, Eurasian Math. J. 5 (2014), no. 3, 9 -- 45. [2] V.I. Burenkov, Recent progress in studying the boundedness of classical operators of real analysis in general Morrey-type spaces. I, Eurasian Math. J. 3 (2012), no. 3, 11 -- 32. [3] V. I. Burenkov, Recent progress in studying the boundedness of classical operators of real analysis in general Morrey-type spaces. II, Eurasian Math. J. 4 (2013), no. 1, 21 -- 45. [4] V. I. Burenkov and M. L. Goldman, Necessary and sufficient conditions for the boundedness of the maximal operator from Lebesgue spaces to Morrey-type spaces, Math. Inequal. Appl. 17 (2014), no. 2, 401 -- 418, DOI 10.7153/mia-17-30. [5] V. I. Burenkov and H. V. Guliyev, Necessary and sufficient conditions for boundedness of the maximal operator in local Morrey-type spaces, Studia Math. 163 (2004), no. 2, 157 -- 176. [6] V. I. Burenkov, H. V. Guliyev, and V. S. Guliyev, Necessary and sufficient conditions for the boundedness of fractional maximal operators in local Morrey-type spaces, J. Comput. Appl. Math. 208 (2007), no. 1, 280 -- 301. [7] V.I. Burenkov, H.V. Guliyev, and V.S. Guliyev, On boundedness of the fractional maximal operator from complementary Morrey-type spaces to Morrey-type spaces, The interaction of analysis and geometry, Contemp. Math., vol. 424, Amer. Math. Soc., Providence, RI, 2007, pp. 17 -- 32. [8] V.I. Burenkov, V.S. Guliyev, A. Serbetci, and T.V. Tararykova, Necessary and sufficient conditions for the boundedness of genuine singular integral operators in local Morrey-type spaces, Eurasian Math. J. 1 (2010), no. 1, 32 -- 53. [9] V.I. Burenkov, A. Gogatishvili, V.S. Guliyev, and R.Ch. Mustafayev, Boundedness of the fractional maximal operator in local Morrey- type spaces, Complex Var. Elliptic Equ. 55 (2010), no. 8-10, 739 -- 758. [10] V.I. Burenkov, A. Gogatishvili, V.S. Guliyev, and R.Ch. Mustafayev, Boundedness of the Riesz potential in local Morrey-type spaces, Potential Anal. 35 (2011), no. 1, 67 -- 87. [11] V.I. Burenkov and E.D. Nursultanov, Description of interpolation spaces for local Morrey-type spaces, Tr. Mat. Inst. Steklova 269 (2010), no. Teoriya Funktsii i Differentsialnye Uravneniya, 52 -- 62 (Russian, with Russian summary); English transl., Proc. Steklov Inst. Math. 269 (2010), no. 1, 46 -- 56. [12] G. B. Folland, Real analysis, 2nd ed., Pure and Applied Mathematics (New York), John Wiley & Sons, Inc., New York, 1999. Modern techniques and their applications; A Wiley-Interscience Publication. [13] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, 2nd ed., Springer-Verlag, Berlin, 1983. [14] A. Gogatishvili, R. Ch. Mustafayev, and L.-E. Persson, Some new iterated Hardy-type inequalities, J. Funct. Spaces Appl. (2012), Art. ID 734194, 30. [15] A. Gogatishvili, B. Opic, and L. Pick, Weighted inequalities for Hardy-type operators involving suprema, Collect. Math. 57 (2006), no. 3, 227 -- 255. [16] V.S. Guliyev, Integral operators on function spaces on the homogeneous groups and on domains in Rn, Doctor's degree dissertation. Mat. Inst. Steklov., Moscow, 1994 (Russian). [17] A. Kufner and L.-E. Persson, Weighted inequalities of Hardy type, World Scientific Publishing Co. Inc., River Edge, NJ, 2003. MR1982932 (2004c:42034) [18] C. B. Morrey, On the solutions of quasi-linear elliptic partial differential equations, Trans. Amer. Math. Soc. 43 (1938), no. 1, 126 -- 166, DOI 10.2307/1989904. [19] R. Ch. Mustafayev and T. Unver, Embeddings between weighted local Morrey-type spaces and weighted Lebesgue spaces, J. Math. Inequal. 9 (2015), no. 1, 277 -- 296, DOI 10.7153/jmi-09-24. [20] B. Opic and A. Kufner, Hardy-type inequalities, Pitman Research Notes in Mathematics Series, vol. 219, Longman Scientific & Technical, Harlow, 1990. 12 A. GOGATISHVILI, R.CH.MUSTAFAYEV, AND T. UNVER Institute of Mathematics, Academy of Sciences of the Czech Republic, Zitn´a 25, 115 67 Praha 1, Czech Republic E-mail address: [email protected] Department of Mathematics, Faculty of Science and Arts, Kirikkale University, 71450 Yahsihan, Kirikkale, Turkey E-mail address: [email protected] Department of Mathematics, Faculty of Science and Arts, Kirikkale University, 71450 Yahsihan, Kirikkale, Turkey E-mail address: [email protected]
1411.2217
3
1411
2016-03-23T08:07:53
K(X,Y) as subspace complemented of L(X,Y)
[ "math.FA" ]
Let X,Y be two Banach spaces ; in the first part of this work, we show that K(X,Y) contains a complemented copy of c0 if Y contains a copy of c0 and each bounded sequence in Y has a subsequece which is w* convergente. Afterward we obtain some results of M.Feder and G.Emmanuele: Finally in this part we study the relation between the existence of projection from L(X,Y) on K(X,Y) and the existence of pro- jection from K(X,Y ) on K(X,Y) if Y has the approximation property. In the second part we study the Radon-Nikodym property in L(X,Y):
math.FA
math
K(X,Y) COMME SOUS-ESPACE COMPL´EMENT´E DE L(X, Y ) DAHER MOHAMMAD R´esum´e. Soient X,Y des espaces de Banach. Dans ce travail nous montrons que K(X, Y ) contient une copie compl´ement´ee de c0, si Y contient ℓ∞ isomorphiquement, ou Y contient c0 isomorphi- quement et toute suite born´ee dans Y ∗ admet une sous-suite qui converge pr´efaiblement. Dans la suite, nous prouvons qu'il existe un espace de Banach X tel que K(X) est compl´ement´e dans L(X), et K(X) n'est pas compl´ement´e dans son bidual. Abstract. Let X,Y be Banach spaces. In this work, we show that K(X, Y ) contains a complemented copy of c0, if Y contains a copy of ℓ∞, or Y contains a copy of c0 and every bounded sequence in Y ∗ has a subsequence which is w∗−convergente In the follwing, we show that there exists a Banach space X such that K(X) is complemented space in L(X) but K(X) is not complemented in its bidual. Classification : 46B07, Secondaire 47B10 Mots cl´es : projection dans L(X, Y ). Introduction. Soient X,Y des espaces de Banach . On d´esigne par L(X, Y ) l'espace des op´erateurs born´es de X dans Y et par K(X, Y ) le sous-espace de L(X, Y ) form´e des op´erateurs compacts. Dans ce travail, nous montrons que K(X, Y ) contient une copie compl´ement´ee de c0, si Y contient ℓ∞ isomorphiquement, ou Y contient c0 isomorphiquement et toute suite born´ee dans Y ∗ admet une sous- suite qui converge pr´efaiblement. Dans la suite, nous retrouvons des r´esultats de M.Feder [Fe], J.Johnson [Joh], et G.Emmanuele [Emm] concernant l'existence d'une projection continue de L(X, Y ) sur K(X, Y ). Finalement, nous racact´erisons l'existence d'une projection conti- nue :K(X, Y ) → K(X, Y )∗∗, si Y ou X ∗ a la propri´et´e de l'approxima- tion born´ee. Pour tout Banach X, on d´esigne par BX la boule unit´e ferm´e de X et par X ∗ le dual de X. Notons pour tout x ∈ X et tout x∗ ∈ X ∗ (x, x∗) = x∗(x). 1 2 DAHER MOHAMMAD D´esignons par (ek)k≥0 la base canonique de c0. Dans la suite on fixe deux espaces de Banach X, Y de dimensions infinies. Proposition 0.1. Supposons que Y contient c0 isomorphiquement. Alors il existe un op´erateur σ : ℓ∞ → L(X, Y ) tel que a) ℓ∞ ≈ σ(ℓ∞). b) σ(ℓ∞) ∩ K(X, Y ) = σ(c0). c) Il existe une projection continue P : L(X, Y ) → σ(ℓ∞) telle que PK(X,Y1) est une projection sur σ(c0), o`u Y1 est un sous-espace de Ba- nach de Y isomorphe `a c0. d) Supposons Y contient ℓ∞ isomorphiquement, ou toute suite born´ee dans Y ∗ admet une sous-suite qui converge pr´efaiblement. Alors il existe Λ ⊂ N et une projection continue Q : L(X, Y ) → σ(ℓ∞(Λ)) telle que QK(X,Y ) est une projection sur σ(c0(Λ)). D´emonstration. a). Soient Y1 un sous-espace de Banach de Y et U : c0 → Y1 un tel isomorphisme. D'apr`es [Di, Chap.XII]-[Jos]-[Ness], il existe une suite (x∗ 0 pr´efaiblement. k)k≥0 dans la sph`ere unit´e de X ∗ telle que x∗ k → On d´efinit l'op´erateur σ : ℓ∞ → L(X, Y ) par σ(α)(x) =Xk≥0 α = (αk)k≥0 ∈ ℓ∞, x ∈ X. k→+∞ Pour tout k ∈ N, il existe xk ∈ X tel que (xk, x∗ k) 6= 0. Cela implique αk(x, x∗ k)U(ek), que σ est injectif. Etape 1 : Montrons que σ est un op´erateur born´e. Pour tout α ∈ ℓ∞ et tout x ∈ X on a kσ(α)(x)kY = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xk≥0 k)ek(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)c0 αk(x, x∗ kUk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xk≥0 k)U(ek)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y U(Xk≥0 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k)ek)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y αk(x, x∗ αk(x, x∗ ≤ ≤ kUk kαkℓ∞ × sup k≥0 (x, x∗ k) ≤ kUk × kαkℓ∞ × kxk . Il en r´esulte que σ est un op´erateur born´e. Etape 2 : Montrons que σ est un isomorphisme sur son image. Fixons k ∈ N. D´esignons par z∗ k la forme lin´eaire d´efinie sur Y1 par (Uej, z∗ k) = δjk, j ∈ N, o`u δjk est le symbole de Kronecker. K(X,Y) COMME SOUS-ESPACE COMPL´EMENT´E DE L(X, Y ) 3 Soient α ∈ ℓ∞, ε > 0 et k ∈ {j ∈ N; αj 6= 0}. Il existe xk ∈ BX tel k) + ε/ αk . Pour tout k ∈ {j ∈ N ; αj 6= 0} on kk ≤ (xk, x∗ que 1 = kx∗ a αk ≤ αk × [(xk, x∗ = (σ(α)(xk), z∗ k) + ε/ αk] = αk × (xk, x∗ kkY ∗ k) + ε ≤ kσ(α)k × sup k≥0 kz∗ 1 k) + ε + ε. . Cela entraıne que σ est un isomorphisme sur son image. b). Montrons que σ(c0) ⊂ σ(ℓ∞) ∩ K(X, Y ). Soient α ∈ c0 et ε > 0. Il existe k0 ∈ N tel que αk ≤ ε pour tout k ≥ k0. Consid´erons Tn : X → Y l'op´erateur d´efinie par Tn(x) = αk(x, x∗ k)Uek, x ∈ X, n ∈ N. nXk=0 Il est ´evident que la suite (Tn)n≥0 est dans K(X, Y ) et que kTn − σ(α)kL(X,Y ) < ε pour tout n ≥ k0, donc σ(α) est un op´erateur compact. Montrons que σ(ℓ∞) ∩ K(X, Y ) ⊂ σ(c0). Remarquons d'abord que si α1, ..., αn ∈ C et y =Xk≤n 0, donc par densit´e pour tout y ∈ Y1 (y, z∗ 0. j ) → j→∞ αkU(ek), (y, z∗ j ) → j→∞ Soit α ∈ ℓ∞ tel que σ(α) ∈ K(X, Y ). Supposons que α /∈ c0. Il existe donc ε0 > 0, tel que pour tout k ∈ N, il existe mk ∈ N v´erifiant mk ≥ k et αmk ≥ ε0. (0.1) Posons T = σ(α) ∈ K(X, Y1). Pour tout k ∈ N, il existe xk dans k) ≥ 1 − 1/k + 2. Il est clair que la boule unit´e de X tel que (xk, x∗ αk = (T xk, z∗ k) pour tout k ∈ N. k)/(xk, x∗ Comme T est un op´erateur compact, il existe une sous-suite (xmkj )j≥0 y ∈ Y1. D'autre part, pour tout j ∈ N dans X telle que T (xmkj ) → j→+∞ αmkj =h(T xmkj − y, z∗ ) + (y, z∗ mkj mkj 0. Par cons´equent αmkj )i /(xmkj mkj ) → et (y, z∗ k) ≥ 1 − 1/k + 2 pour tout k ∈ N, ce qui est impossible d'apr`es (0.1). Donc α ∈ c0. 0, car (xk, x∗ j→+∞ j→+∞ → , x∗ mkj ) c). D'apr`es le th´eor`eme de Hahn-Banach, pour tout k ∈ N, il existe y∗ k ∈ Y ∗ qui prolonge z∗ 1 et kz∗ k ∈ Y ∗ = ky∗ kkY ∗ kkY ∗ 1 4 DAHER MOHAMMAD Soit maintenant T ∈ L(X, Y ). D´efinissons a(T ) ∈ ℓ∞ par αk(T ) = k), k ∈ N et P : L(X, Y ) → σ(ℓ∞) par P (T ) = k)/(xk, x∗ (T xk, y∗ σ [α(T )] , T ∈ L(X, Y ). Etape 1 : Montrons que P est une projection. Soient α ∈ ℓ∞. Posons T = σ(α). Remarquons que (T xk, z∗ k)/(xk, x∗ k) = αk(U(ek), z∗ k)/(xk, x∗ k) = αk. αk(T ) = (T xk, y∗ k)/(xk, x∗ k) = k)(xk, x∗ Donc P est une projection. Etape 2 : Montrons que P est continue. Pour tout T ∈ L(X, Y ) nous avons kσ(α(T ))k ≤ kσk × kα(T )kℓ∞ ≤ Etape 3 : Soit T ∈ K(X, Y1). Montrons que α(T ) ∈ c0. Comme P est une projection, d'apr`es b), T = σ(α(T )) ∈ K(X, Y ) ∩ k)] ≤ kσk×[kT k]×(cid:2)supk≥0(ky∗ kk /(1 − 1/k + 2))(cid:3) .(cid:4) kσk×supk≥0 [(T xk, y∗ k)/(xk, x∗ σ(ℓ∞) = σ(c0), c'est-`a-dire que α(T ) ∈ c0.(cid:4) d). Cas 1 : Y contient ℓ∞ isomorphiquement. Comme ℓ∞ est injectif, on peut supposer que Y ≈ ℓ∞. Consid´erons G : Y → ℓ∞ un isomorphisme. Notons (fn)n≥0 la base canonique de ℓ1, n = G∗fn ∈ Y ∗, n ∈ N et U : c0 → Y1, la restriction de G−1 `a c0. h∗ Soit maintenant T ∈ L(X, Y ). D´efinissons β(T ) ∈ ℓ∞ par βk(T ) = (T xk, h∗ k), k ∈ N et Q : L(X, Y ) → σ(ℓ∞) par Q(T ) = σ [β(T )] , T ∈ L(X, Y ). Par un argument analogue `a celui de c), on montre que Q est une projection continue et QK(X,Y ) est une projec- tion sur σ(c0) (ici Λ = N).(cid:4) k)/(xk, x∗ Cas 2 : Toute suite born´ee dans Y ∗ admet une sous-suite pr´efaiblement convergeante. Il existe une sous-suite (y∗ nk)k≥0 telle que y∗ nk → k→∞ y∗ pr´efaiblement dans Y ∗. D´esingons par Λ = {nk; k ∈ N} . Soit U : c0(Λ) → Y1 un isomorphisme. On d´efinit l'op´erateur σ : k)U(ek), x ∈ X, α ∈ ℓ∞(Λ) ℓ∞(Λ) → L(X, Y ) par σ(α)(x) =Xk∈Λ Pour tout T ∈ L(X, Y ) on d´efinit γ(T ) ∈ ℓ∞(Λ) par γk(T ) = k), k ∈ Λ et Q : L(X, Y ) → σ(ℓ∞(Λ)) par Q(T ) = αk(x, x∗ k − y∗)/(xk, x∗ (T xk, y∗ σ [γ(T )] , T ∈ L(X, Y ). Montrons que Q est une projection. Consid´erons T = σ(α), α ∈ ℓ∞(Λ). Pour tout k ∈ Λ on a γk(T ) = (T xk, y∗ k−y∗)/(xk, x∗ k) = (T xk, z∗ k−y∗)/(xk, x∗ αj(xk, x∗ j )(U(ej), z∗ αk, car pour tout j ∈ Λ (U(ej), y∗) = limk→∞(U(ej), z∗ nk) = 0. k) ="Xj∈Λ k − y∗)# /(xk, x∗ k) = K(X,Y) COMME SOUS-ESPACE COMPL´EMENT´E DE L(X, Y ) 5 Par un argument analogue `a celui de c), on montre que Q est conti- nue. Montrons finalement que γ(T ) ∈ c0(Λ), pour tout T ∈ K(X, Y ). Soit T ∈ K(X, Y ). Supposons que γ(T ) /∈ c0, il existe ε0 > 0 tel que pour tout k ∈ N, il existe mk ∈ Λ v´erifiant mk ≥ k et (cid:12)(cid:12)(cid:12)γmk (T )(cid:12)(cid:12)(cid:12) ≥ ε0. (0.2) → j→∞ D'autre part, il existe une sous-suite (xmkj )j≥0 telle que T xmkj y ∈ Y. Pour tout j ∈ N on a γmkj (T ) =h(T xmkj − y, y∗ mkj − y∗) + (y, y∗ mkj − y∗)i /(xmkj − y∗) → j→∞ , x∗ mkj ). 0, ceci est Il est clair que γmkj (T ) → j→∞ impossible, donc γ(T ) ∈ c0(Λ).(cid:4) 0, car (y, ymkj Les deux corollaires suivants sont une cons´equence de la proposition 0.1-d. Corollaire 0.1. Supposons Y contient ℓ∞ isomorphiquement. Alors K(X, Y ) contient une copie compl´ement´ee de c0. Corollaire 0.2. Supposons que Y contient c0 isomorphiquement et que toute suite born´ee dans Y ∗ admet une sous-suite qui converge pr´efaiblement. Alors K(X, Y ) contient une copie compl´ement´ee de c0. Corollaire 0.3. [Fe, Coroll.4]-[Ch, Coroll.4]-[Joh, Th. 4]Supposons que Y contient c0 isomorphiquement. Alors il n'existe aucune projection continue :L(X, Y ) → K(X, Y ). D´emonstration. Suppososns qu'il existe une projection continue P : L(X, Y ) → K(X, Y ). Cas 1 : Y ne contient pas ℓ∞ isomorphiquement et X ne contient pas une copie compl´ement´ee de ℓ1. D'apr`es [Kalt, Th.4], K(X, Y ) ne contient pas ℓ∞ isomorphiquement. D'autre part, la proposition 0.1, nous montre qu'il existe un isomor- phisme σ : ℓ∞ → σ(ℓ∞) ⊂ L(X, Y ). En appliquant [Kalt, Prop.2], on voit que P ◦ σ : ℓ∞ → K(X, Y ) est faiblement compact, ceci est impossible, car P ◦ σc0 un isomorphisme.(cid:4) Cas 2 : Y ne contient pas ℓ∞ isomorphiquement et X contient une = σc0 copie compl´ement´ee de ℓ1. 6 DAHER MOHAMMAD Ce cas implique qu'il existe une porojection continue P1 : L(ℓ1, Y ) → K(ℓ1, Y ). D'apr`es [Kalt, Th.6], K(ℓ1, Y ) ne contient pas c0 isomorphiquement, ce qui est impossible, car K(ℓ1, Y ) contient ℓ∞ isomorphiquement.(cid:4) Cas 3 : Y contient ℓ∞ isomorphiquement. D'apr`es la proposition 0.1-d), il existe une projection Q : K(X, Y ) → σ(c0). Comme ℓ∞ est un espace de Grothendieck, d'apr`es [Groth, Groth] σ ◦ Q ◦ P : ℓ∞ → un σ(c0) est faiblement compact, ce qui est impossible, car σ ◦ Q ◦ Pc0 isomorphisme.(cid:4) Corollaire 0.4. [Emm, Th.2]-[Ch, Coroll.7]-[Joh, Croll.7]Supposons qu'il existe une projection continue P : L(X, Y ) → K(X, Y ). Alors K(X, Y ) ne contient pas c0 isomorphiquement. D´emonstration. D'apr`es le corollaire 0.3, Y ne contient pas c0 isomorphiquement. Montrons d'abord que K(X, Y ) ne contient pas ℓ∞ isomorphique- ment. Supposons que K(X, Y ) contient ℓ∞ isomorphiquement. Le r´esultat de [Kalt, Th.4], nous indique que X contient une copie compl´ement´ee de ℓ1, il en r´esulte qu'il existe donc une projection continue :L(ℓ1, Y ) → K(ℓ1, Y ), ce qui est impossible d'apr`es [Kalt, Th.6]. On en d´eduit que K(X, Y ) ne contient pas ℓ∞ isomorphiquement. Supposons qu'il existe un isomorphisme σ1 : c0 → σ1(c0) ⊂ K(X, Y ). Pour tout α ∈ ℓ∞ et tout x ∈ X on a < ; J est un sous-ensemble de N fini) < ; J est un sous-ensemble de N fini ; J est un sous-ensemble de N fini) < +∞. αnσ1(en)x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y sup((cid:13)(cid:13)(cid:13)(cid:13)(cid:13)XJ kxk × sup αnσ1(en)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L(X,Y ) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)XJ αnen(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)c0 kxk × kσ1k sup((cid:13)(cid:13)(cid:13)(cid:13)(cid:13)XJ Comme Y ne contient pas c0 isomorphiquement, la s´erieXn≥0 ℓ∞ → L(X, Y ) par σ2(α)(x) = Xk≥0 converge dans Y pour tout x ∈ X [Kalt, prop.3]. On d´efinit σ2 : αkσ1(ek)(x), x ∈ X . Consid´erons αnσ1(en)x P ◦ σ2 : ℓ∞ → K(X, Y ). D'apr`es [Kalt, Prop2], P ◦ σ2 est faiblement compact, car K(X, Y ) ne contient pas ℓ∞ isomorphiquement, c'est im- possible, car P ◦ σ2c0 = σ1.(cid:4) K(X,Y) COMME SOUS-ESPACE COMPL´EMENT´E DE L(X, Y ) 7 Proposition 0.2. Supposons que L(X, Y ) ne contient pas ℓ∞ isomr- phiquement. Alors K(X, Y ) ne contient pas c0 isomorphiquement. D´emonstration. Supposons qu'il existe un isomorphisme σ1 : c0 → σ1(c0) ⊂ K(X, Y ). Remarquons d'apr`es la proposition 0.1 que Y ne contient pas c0 iso- morphiquement et que < +∞ pour tout α ∈ ℓ∞ et tout x ∈ X. L'espace Y ne contient pas c0 αnσ1(en)x αnσ1(en)(x)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y ; J est un sous-ensemble fini de N sup (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) XJ isomorphiquement, donc d'apr`es [Kalt, Prop.3], la s´erieXk≥0 σ2(α)x =Xn≥0 converge dans Y pour tout x ∈ X. On d´efinit σ2 : ℓ∞ → L(X, Y ) par αnσ1(en)x, α ∈ ℓ∞, x ∈ X. Comme L(X, Y ) ne contient pas ℓ∞ isomorphiquement, d'apr`es [Kalt, Prop.2] σ2 est faiblement com- pact. Il en r´esulte que σ1 est faiblement compact, ceci est impossible, par cons´equent K(X, Y ) ne contient pas c0 isomorphiquement.(cid:4) Soit X, Y deux espaces de Banach. D´esignons par X ∨ ⊗ Y (resp. Xb⊗Y ) le produit injectif de X,Y (resp. le produit projectif de X, Y ). D´esignons d'autre part, par B(X × Y ) l'espace des formes bilin´eaires u : X × Y → C telles que sup{u(x, y) ; kxk , kyk ≤ 1} < ∞. Thorme 0.1. Il existe un espace de Banach Z qui poss`edes des pro- pri´et´es suivantes : 1) K(Z) est un espace compl´ement´e de L(Z). 2) K(Z) n'est pas compl´ement´e dans son bidual. Proposition 0.3. Supposons que Y ou X ∗ a la propri´et´e de l'approxi- mation born´ee. Alors K(X, Y ) est complement´e dans K(X, Y )∗∗, si et seulement si K(X, Y ) et Y sont compl´ement´es dans L(X, Y ) et dans Y ∗∗ respectivement. D´emonstration de proposition 0.3. Supposons qu'il existe une projection continue Q : K(X, Y )∗∗ → K(X, Y ) . Montrons que Y est compl´ement´e dans son bidual. Consid´erons x0 dans la sph`ere unit´e de X et x∗ dans la sph`ere unit´e de X ∗ tels que (x0, x∗) = 1. On d´efinit l'op´erateur H : Y → K(X, Y ) 8 DAHER MOHAMMAD par H(y)(x) = (x, x∗)y, x ∈ X, y ∈ Y et l'op´erateur G : K(X, Y ) → Y, par G(T ) = T x0, T ∈ K(X, Y ). Notons Q1 = G ◦ Q ◦ H ∗∗ : Y ∗∗ → Y. Il est clair que Q1 est une projection continue sur Y. Montrons que K(X, Y ) est compl´ement´e dans L(X, Y ). Cas 1 : Y a la propri´et´e de l'approximation born´ee. D'apr`es [Joh, Lemma 2], il existe un isomorphisme U : L(X, Y ) → K(X, Y )∗∗ tel que la restriction de U `a K(X, Y ) est l'identit´e. Notons P = Q ◦ U : L(X, Y ) → K(X, Y ). Il est facile de voir que P est une projection continue sur K(X, Y ). Cas 2 : X ∗ a la propri´et´e de l'approximation born´ee. Il existe une suite g´en´eralis´ee (Vi)i∈I d'op´erateurs de rang finis :X ∗ → X ∗ telle que Vi → IX ∗ uniform´ement sur tout compact de X ∗ et supi∈I kVikL(X ∗) < ∞. Posons pour i ∈ I et T ∈ L(X, Y ), U T i = Vi ◦ T ∗ : Y ∗ → X ∗ et H T : X → Y. Il est ´evident i ∈ K(X, Y ). On d´efinit l'op´erateur U : L(X, Y ) → K(X, Y )∗∗, que H T par U(T ) = limU H T i , T ∈ L(X, Y ) (limite pr´efaible suivant U dans K(X, Y )∗∗). On se propose de montrer que P = Q ◦ U : L(X, Y ) → K(X, Y ) est une projection continue. i = Q1 ◦ (U T i )∗ X Remarquons d'abord que P est continue, car la suite (Vi)i∈I est born´ee dans L(X ∗). Consid´erons T ∈ K(X, Y ). Comme T ∗ est com- pact,(cid:13)(cid:13)U T 0. Il en r´esulte que(cid:13)(cid:13)H T i − T ∗(cid:13)(cid:13)L(Y ∗,X ∗) → 0, ceci implique que(cid:13)(cid:13)(U T i − Q1 ◦ T(cid:13)(cid:13)L(X,Y ) =(cid:13)(cid:13)H T cons´equent P (T ) = Q ◦ U(T ) = Q(T ) = T.(cid:4) i )∗ − T ∗∗(cid:13)(cid:13)L(X ∗∗,Y ∗∗) → i − T(cid:13)(cid:13)L(X,Y ) → 0, par Supposons qu'il existe une projection P : L(X, Y ) → K(X, Y ) et Q1 : Y ∗∗ → Y une projection continue. Montrons que K(X, Y ) est compl´ement´e dans K(X, Y )∗∗. Comme Y ou X ∗ `a la propri´et´e de l'approximation born´ee, K(X, Y ) = ∨ X ∗ ⊗ Y. On d´efinit l'op´erateur J : Xb⊗Y ∗ → K(X, Y )∗ par (JXk≤n j ⊗ yj) = Xk≤n,j≤m k,Xj≤m xk ⊗ y∗ (xk, x∗ x∗ les y∗ k ∈ Y ∗, les x∗ j ∈ X ∗ et les yj ∈ Y. j ) × (yj, y∗ k), les xk ∈ X, Montrons que J est un op´erateur born´e. Consid´erons (xk)k≤n ∈ X, (y∗ k)k≤n ∈ Y ∗, (x∗ j )j≤m ∈ X ∗, (yj)j≤m ∈ Y. On a alors K(X,Y) COMME SOUS-ESPACE COMPL´EMENT´E DE L(X, Y ) 9 x∗ k, k≤n xk ⊗ y∗ j ⊗ yj)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k,Xj≤m (JXk≤n j ⊗ yj) (xk))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj≤m X(y∗ kkY ∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) j ⊗ yj)# (xk)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Y "Xj≤m ≤ Xk≤n j ⊗ yj)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)K(X,Y ) kkY ∗ kxkkX(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xj≤m ≤ Xk≤n ky∗ ky∗ (x∗ x∗ x∗ . Donc J est continue. Soit maintenant R ∈ K(X, Y )∗∗. On d´efinit ξR ∈ (Xb⊗Y ∗)∗ par ξR(u) = (J(u), R), u ∈ Xb⊗Y ∗. D'apr`es [Du, coroll.2,chap.VIII-2], ξR ∈ B(X × Y ∗) = L(X, Y ∗∗). Consid´erons V : L(X, Y ∗∗) → L(X, Y ) l'opr´erateur d´efini par V (S) = Q1 ◦ S ∈ L(X, Y ), S ∈ L(X, Y ∗∗) et Q : K(X, Y )∗∗ → K(X, Y ) l'op´erateur d´efini par Q(R) = P [V (ξR)] , R ∈ K(X, Y )∗∗. On remarque que Q est une projection continue sur K(X, Y ).(cid:4) D´emonstration du th´eor`eme 0.1. D'apr`es [Tab, Th4.1.1], il existe un espace de Banach Z tel que Z ∗ = ℓ1 et K(Z) soit compl´ement´e dans L(Z). Remarquons que l'es- pace Z n'est pas r´eflexif, comme Z ∗∗ = ℓ∞ est un espace de Grothen- dieck [Groth] Z ne peut pas etre compl´ement´e dans son bidual. D'autre part, Z ∗ = ℓ1 a la propri´et´e de l'approximation born´ee, d'apr`es [Du, Chap.VIII-3] Z a la propri´et´e de l'approximation born´ee. En appliquant la proposition 0.3, on voit que K(Z) n'est pas compl´ement´e dans son bidual.(cid:4) Problme 0.1. Existe-il deux espaces de Banach X et Y tel que K(X, Y ) soit compl´ement´e dans son bidual et L(X, Y ) 6= K(X, Y )? Remarque 0.1. Si Y ou X ∗ a la propri´et´e de l'approximation born´ee, il existe une projection de L(X, Y )∗∗ sur K(X, Y )∗∗. Preuve. En effet, Supposons que Y a la propri´et´e de l'approximation born´ee. Soit P1 : K(X, Y )∗∗∗∗ → K(X, Y )∗∗ une projection continue telle que (P1V, ξ) = (V, ξ) pour ξ ∈ K(X, Y )∗ et V ∈ K(X, Y )∗∗. 10 DAHER MOHAMMAD L'application P1 ◦ U ∗∗ : L(X, Y )∗∗ → (K(X, Y )∗∗ est une projection continue (U est l'op´erateur d´efini auparavant).(cid:4) Remarque 0.2. Soit X un espace de Banach. Supposons qu'il existe une projection continue P : L∞(T, X) → L∞(T) dimension finie. ∨ ⊗ X. Alors X est de Preuve. En effet, Comme L∞(T, X) contient une copie (isom´etrique) compl´ement´ee de ∨ ∨ ⊗X contient une copie compl´ement´ee de ℓ∞ ℓ∞(X) et L∞(T) ⊗X, alors ∨ K(ℓ1, X) = ℓ∞ ⊗ X est compl´ement´e de L(ℓ1, X) = ℓ∞(X). D'apr`es [Kalt, th.6], X est de dimension finie.(cid:4) R´ef´erences [Ch] [Du] [Di] I. Chenciu, Complemented spaces of operators, Proc. Am. Math. Soc. Vol. 133, n ◦. 9, 2621-2623, (2005). J. Diestel, J. Uhl, Vector measures, Math. Surveys N ◦.15 , (1977), A. M. S. J. Diestel, Sequence and series in Banach spaces, Graduate Texts in Math. n ◦. 92, (Springer Verlag, 1984). [Emm] G. Emmanuele, A remark on the containment of c 0 in spaces of compact operators, Math. Proc. Camb. phil. Soc. 111, 331-335, (1992) [Fe] M. Feder, On the non existence of projection onto the space of compact operators, Cand. Math. Bull. 25, 78-81, (1992). [Fe1] M. Feder and P. Saphar, Spaces of compact operators and their dual spaces, Israel J. Math. 21, 38-49, (1974). [Groth] A. Grothendieck, Sur les applications lin´eaires faiblement compactes d'espaces de type C(K), Canad. J. Math. 5, 129-173, (1953). [Joh] [Jos] J. Johnson, Remarks on Banach spaces of compact operators, J. F. A. 32, 304-311, (1979). B. Josefson, Weak sequential convergence in the dual of a Banach spaces deos not imply norm convergence, Ark. Mat. 13, 79-89, (1975). [Kalt] N. J. Kalton, Spaces of compact operators, Math. Ann. 208, 267-278, (1974). [Ness] A. Nessenzweig, w ∗ sequential convergence, Israel J. Math. 22, 79-89, (1975). [Tab] M. Tarbard, Operators on Banach spaces of Bourgain-Delbaen-Type, arXiv :1309.7469V1 (2012). [Ros] H. P. Rosenthal, On relativity disjoint families of measures with some applications to space theory, Stud. Math. 37, 13-36, (1970). [email protected]
1504.02355
1
1504
2015-04-09T15:33:23
A $0-2$ law for cosine families with $\limsup$ to $\infty$
[ "math.FA", "math.OA" ]
For $\left(C(t)\right)_{t\in\mathbb R}$ being a cosine family on a unital normed algebra, we show that the estimate $\limsup_{t\to\infty^{+}}\|C(t) - I\| <2$ implies that $C(t)=I$ for all $t\in\mathbb R$. This generalizes the result that $\sup_{t\geq0}\|C(t)-I\|<2$ yields that $C(t)=I$ for all $t\geq0$. We also state the corresponding result for discrete cosine families and for semigroups.
math.FA
math
A 0 − 2 LAW FOR COSINE FAMILIES WITH lim sup TO ∞ FELIX L. SCHWENNINGER AND HANS ZWART Abstract. For (C(t))t∈R being a cosine family on a unital normed algebra, we show that the estimate lim supt→∞+ kC(t) − Ik < 2 implies that C(t) = I for all t ∈ R. This generalizes the result that supt≥0 kC(t) − Ik < 2 yields that C(t) = I for all t ≥ 0. We also state the corresponding result for discrete cosine families and for semigroups. In the recent past, laws of the form 1. Introduction (limsup-law ) (sup-law ) lim sup t→0 kC(t) − Ik < r =⇒ lim t∈R kC(t) − Ik < r =⇒ C(t) = I ∀t ∈ R, t→0kC(t) − Ik = 0, sup where r > 0 and (C(t))t∈R is a cosine family of elements in a unital Banach algebra A (with identity element I) were studied, see [1, 2, 3, 6] and [7, 9] for the special case where (C(t))t∈R is strongly continuous and A = B(X) is the Banach algebra of bounded operators on a Banach space X. For both, the limsup-law and the sup-law the largest possible constant r was shown to be 2. In this note we consider the condition (1.1) lim sup t→∞ kC(t) − Ik < 2, lim sup which is weaker than the premise in the sup-law, and show that (limsup-∞-law ) for r = 2 holds, see Theorem 2.5. A related question was raised in [8, Remark 2.6] for, more general, scaled versions of these laws. More precisely, it was asked whether for a ≥ 0 the following holds for some r, (1.2) t→∞ kC(t) − Ik < r =⇒ C(t) = I ∀t ∈ R, t→∞ kC(t) − cos(at)k < r =⇒ C(t) = cos(at) ∀t ∈ R. lim sup Let us mention that 'scaled version' (where the unity element I gets replaced by cos(at)I) of limsup-law and sup-law have a different optimal constant r = 8 , see 3√3 [2, 4, 5]. In the following, we show that (limsup-∞-law) holds, using techniques by J. Esterle [6]. Finally we state the corresponding result for semigroups, for which zero-one-laws have been studied much earlier than for cosine families. Date: March 20, 2019. 1991 Mathematics Subject Classification. Primary 47D09; Secondary 47D06. Key words and phrases. Cosine families, Semigroup of operators, Zero-two law, Zero-one law. The first named author has been supported by the Netherlands Organisation for Scientific Research (NWO), grant no. 613.001.004. 1 2 SCHWENNINGER AND ZWART 2. A lim supt→∞ - law In the following, for a unital normed algebra A, let I denote the identity element. Lemma 2.1. Let (C(t))t∈R be a cosine family in a unital Banach algebra. If lim sup t→∞ kC(t) − Ik = 0, then C(t) = I for all t ∈ R. Proof. From the assumption follows that limt→∞ kC(t) − Ik = 0. By d'Alembert's defining identity for cosine families, (2.1) C(t + s) + C(t − s) = 2C(t)C(s), for all s, t ∈ R. Thus, letting t → ∞, we derive 2I = 2C(s) for all s ∈ R. (cid:3) The following lemma is a slight extension of Esterle's Lemma 2.1 in [6], as we also allow for t0 = ∞. The proof is completely analogous the case case t0 = 0. Lemma 2.2. Let (c(t))t∈R be a complex-valued cosine family and t0 ∈ {0,∞}. Then, we have one of the following situations. (i) lim supt→t0 c(t) − 1 = ∞, (ii) lim supt→t0 c(t) − 1 = 2, (iii) lim supt→t0 c(t) − 1 = 0. Moreover, in case (iii), it follows that (2.2) c(t) =(cid:26) 1 cos(at) if if t0 = ∞, t0 = 0, for some a ≥ 0. Proof. As mentioned the proof is analogous to the one in [6, Lemma 2.1]. In case (iii) and t0 = ∞, it follows by Lemma 2.1 that c(t) = 1 for all t ∈ R. (cid:3) Proposition 2.3. Let (C(t))t∈R be a cosine family on a unital Banach algebra A. If lim supt→∞ Proof. Let A denote the space of characters on A. For all t ∈ R we have that (2.3) ρ (C(t) − I) < 2, then ρ (C(t) − I) = 0 for all t ∈ R. ρ (C(t) − I) = sup χ∈ Aχ (C(t) − I) = sup χ∈ Aχ(C(t)) − 1. Thus, by the assumption we get that lim supt→∞ χ(C(t)) − 1 < 2 for χ ∈ A. As (χ (C(t)))t∈R is a complex-valued cosine family, Lemma 2.2 then implies that χ(C(t)) = 1 for all t ∈ R and χ ∈ A. Using this in (2.3), we deduce that ρ(C(t) − I) = 0 for all t ∈ R. (cid:3) As pointed by Esterle [6], for a commutative unital Banach algebra A, for x ∈ A with kxk ≤ 1 we can define (2.4) √I − x := ∞ Xn=0 (−1)nαnxn, A 0 − 2 LAW FOR lim sup TO ∞ 3 1 2 ( 1 n! 2−1)...( 1 2 −n+1) = (−1)n−1 n2n−1(cid:0)2(n−1) where (−1)nαn, with α0 = 1, αn = 1 n−1 (cid:1), n > 0, are the Taylor coefficients of the function z → √1 − z at the origin (with convergence radius equal to 1). Since (−1)n−1αn > 0 for n ≥ 1, (2.5) (−1)n−1αnkxkn = 1 −p1 − kxk. Lemma 2.4 (Esterle 2015, [6]). Let (C(t))t∈R be a cosine family in a unital Banach algebra.If kC(2s) − Ik ≤ 2 and that ρ(C(s) − I) < 1 for some s ∈ R, Then, √I − x(cid:13)(cid:13)(cid:13) ≤ αnkxkn = Xn=1 ∞ Xn=1 I − (cid:13)(cid:13)(cid:13) ∞ 1 where the square root is defined as described above. C(s) =rI − I − C(2s) 2 , With the above preparatory results, the limsup-∞-law is now easy to show. The proof can be done analogously to the one in [6, Theorem 3.2], which in turn can be seen as an elegant refinement of the technique used in the three-lines-proof in [1]. Theorem 2.5. Let (C(t))t∈R be a cosine family in a unital Banach algebra A. Then, lim supt→∞ kC(t) − Ik < 2 implies that C(t) = I for all t ∈ R. Proof. By Proposition 2.3, we have that ρ(C(t) − I) = 0 for t ∈ R. Thus, we can apply Lemma 2.4 and Eq. (2.5) so that for all s ∈ R, 2 With S := lim sups→∞ kC(s) − Ik, this yields that S 2 ≤ 1, kI − C(s)k ≤ 1 −s1 −(cid:13)(cid:13)(cid:13)(cid:13) S ≤ 1 −r1 − I − C(2s) ≤ 1. (cid:13)(cid:13)(cid:13)(cid:13) which implies that S = 0. Hence, Lemma 2.1 concludes the assertion. (cid:3) Remark 2.6. After finishing this note, Esterle pointed out that, alternatively, [5, Theorem 2.3] implies that for a bounded cosine sequence with ρ(C(1) − I) = 0, it follows that C(t) = cos(at)I for all t ∈ R and some a ∈ R. Thus, lim supt→∞ kC(t)− Ik < 2 implies C(t) = I for all t ∈ R and therefore, the use of Lemma 2.4 can be omitted. Remark 2.7. It is clear that Theorem 2.5 generalizes the sup-law with r = 2. We remark that the known proofs of the sup-law, see [3, 9], which use a diagonalisation argument and the limsup-law, can not be generalized to the assertion of Theorem 2.5. 2.1. A discrete lim sup-law. For discrete cosine families, or cosine sequences (C(n))n∈Z, the following was proved in [9] (There, it was formulated for the special case of C : Z → B(X) for a Banach space X. However, the proof is the same for general Banach-algebra-valued cosine families) Theorem 2.8 ([9]). Let (C(n))n∈Z be a discrete cosine family in a unital Banach algebra. Then, sup n∈NkC(n) − Ik < 3 2 =⇒ C(n) = I ∀n ∈ Z. 4 SCHWENNINGER AND ZWART The proof is based on an elegant idea of Arendt, which can directly be applied to weaken the sup to lim sup in the theorem. Theorem 2.9. Let (C(n))n∈Z be a discrete cosine family in a unital Banach alge- bra. Then, lim sup n→∞ kC(n) − Ik < 3 2 =⇒ C(n) = I ∀n ∈ Z. This result is optimal as can be seen by C(n) = cos( 2nπ 3 ) which yields lim supn→∞ kC(n)− Ik = 3 2 , see [9, Theorem 3.2]. 3. The corresponding semigroup result Let us finally state the corresponding result for (discrete) semigroups in a unital normed algebra, which is a corollary of a well-known result by Wallen [10]. Theorem 3.1. Let {Tn}n∈N be a semigroup in a normed unital algebra. Then, (3.1) lim sup n→∞ kTn − Ik < 1 =⇒ Tn = I ∀n ∈ N. Proof. If lim supn→∞ kTn−Ik < 1, then lim inf n∈N [10], the assertion follows. Remark 3.2. Clearly, Theorem 3.1 implies that for a semigroup on [0,∞), (T (t))t≥0, we have that j=1 kTj−1k < 1. By Wallen (cid:3) 1 nPn (3.2) lim sup t→∞ kT (t) − Ik < 1 =⇒ T (t) = I ∀t ≥ 0. References [1] W. Arendt. A 0 − 3/2 - Law for Cosine Functions. Ulmer Seminare, Funktionalanalysis und Evolutionsgleichungen, pages 17:349 -- 350, 2012. [2] A. Bobrowski, W. Chojnacki, and A. Gregosiewicz. On close-to-scalar one-parameter cosine families. Submitted, 2015. [3] W. Chojnacki. Around Schwenninger and Zwart's zero-two law for cosine families. Submitted, 2015. [4] W. Chojnacki. On cosine families close to scalar cosine families. J. Aust. Math. Soc., 2015. preprint available at arXiv:1411.0854, DOI:10.1017/S1446788715000038. [5] J. Esterle. Bounded cosine functions close to continuous scalar bounded cosine functions. arXiv:1502.00150, 2015. [6] J. Esterle. A short proof of the zero-two law for cosine families. preprint, 2015. [7] S. Fackler. Regularity of semigroups via the asymptotic behaviour at zero. Semigroup Forum, 87(1):1 -- 17, 2013. [8] F. L. Schwenninger and H. Zwart. Less than one implies zero. Submitted, available at arXiv:1310.6202, 2015. [9] F. L. Schwenninger and H. Zwart. Zero-two law for cosine families. J. Evol. Equ., 2015. to appear, DOI:10.1007/s00028-015-0272-8. [10] L. J. Wallen. On the magnitude of xn − 1 in a normed algebra. Proc. Amer. Math. Soc., 18:956, 1967. Department of Applied Mathematics,, University of Twente, P.O. Box 217,, 7500 AE Enschede, The Netherlands E-mail address: [email protected] Department of Applied Mathematics,, University of Twente, P.O. Box 217,, 7500 AE Enschede, The Netherlands E-mail address: [email protected]
1706.07517
2
1706
2017-06-26T16:16:48
Strong hypercontractivity and strong logarithmic Sobolev inequalities for log-subharmonic functions on stratified Lie groups
[ "math.FA", "math.AP" ]
On a stratified Lie group $G$ equipped with hypoelliptic heat kernel measure, we study the behavior of the dilation semigroup on $L^p$ spaces of log-subharmonic functions. We consider a notion of strong hypercontractivity and a strong logarithmic Sobolev inequality, and show that these properties are equivalent for any group $G$. Moreover, if $G$ satisfies a classical logarithmic Sobolev inequality, then both properties hold. This extends similar results obtained by Graczyk, Kemp and Loeb in the Euclidean setting.
math.FA
math
Strong hypercontractivity and strong logarithmic Sobolev inequalities for log-subharmonic functions on stratified Lie groups Nathaniel Eldredge∗ September 5, 2018 . A F h t a m [ 2 v 7 1 5 7 0 . 6 0 7 1 : v i X r a Abstract On a stratified Lie group G equipped with hypoelliptic heat kernel measure, we study the behavior of the dilation semigroup on Lp spaces of log-subharmonic functions. We consider a notion of strong hyper- contractivity and a strong logarithmic Sobolev inequality, and show that these properties are equivalent for any group G. Moreover, if G satisfies a classical logarithmic Sobolev inequality, then both proper- ties hold. This extends similar results obtained by Graczyk, Kemp and Loeb in the Euclidean setting. Contents 1 Introduction 1.1 Background and motivation . . . . . . . . . . . . . . . . . . . 1.2 Statement of results . . . . . . . . . . . . . . . . . . . . . . . 2 Stratified Lie groups and hypoelliptic heat kernels 2.1 Stratified Lie groups . . . . . . . . . . . . . . . . . . . . . . . 2.2 The dilation semigroup . . . . . . . . . . . . . . . . . . . . . . 2.3 Sub-Riemannian geometry on G . . . . . . . . . . . . . . . . 2.4 Properties of the heat kernel . . . . . . . . . . . . . . . . . . . 3 Log-subharmonic functions 2 2 6 9 9 10 13 15 18 ∗School of Mathematical Sciences, University of Northern Colorado, 501 20th St. Box 122, Greeley, CO 80639 USA. Email [email protected]. 1 4 Examples and special cases 4.1 Euclidean space . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 The Heisenberg group . . . . . . . . . . . . . . . . . . . . . . 4.3 H-type groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Complex stratified Lie groups . . . . . . . . . . . . . . . . . . 5 Convolution and approximation 6 Differentiation under the integral sign 7 Proofs of the main results 7.1 LSI implies sLSI 7.2 7.3 . . . . . . . . . . . . . . . . . . . . . . . . . sHC implies sLSI . . . . . . . . . . . . . . . . . . . . . . . . . sLSI implies sHC . . . . . . . . . . . . . . . . . . . . . . . . . 8 Weaker notions of subharmonicity 9 Acknowledgments 1 Introduction 20 20 21 22 22 23 25 28 28 29 30 33 34 1.1 Background and motivation The topic of this paper is inspired by two papers of P. Graczyk, T. Kemp, and J.-J. Loeb [14, 15], in which they introduced notions of strong hypercon- tractivity and a strong logarithmic Sobolev inequality for log-subharmonic functions on real Euclidean space equipped with an appropriate probability measure, and showed the intrinsic equivalence of these two notions. In the present paper, we extend their results to the setting of a stratified real Lie group equipped with hypoelliptic heat kernel measure, which we view in this context as a natural generalization of Euclidean space with Gaussian measure. As motivation, we begin by recalling the classical notion of hypercon- tractivity and its relationship to the logarithmic Sobolev inequality. Let µ be standard Gaussian measure on Rn, and let A be the self-adjoint Ornstein -- Uhlenbeck operator on L2(µ) given by Af (x) = −∆f (x) + x · ∇f (x). (For this introduction, we will work formally and ignore domain considerations.) 2 Hypercontractivity is the statement that ke−tAfkLq(µ) ≤ kfkLp(µ), t ≥ tN (p, q) := 1 2 log(cid:18) q − 1 p − 1(cid:19) , f ∈ Lp(µ), 1 < p ≤ q < ∞. (1.1) This result was proved by E. Nelson [31, 32] with improvements by J. Glimm [13]; see [19] for a broad historical survey of results in this area. Intuitively, (1.1) says that after a certain characteristic time tN , known as "Nelson's time,", the Ornstein -- Uhlenbeck semigroup e−tA improves integrability from Lp(µ) to Lq(µ). The value of tN given in (1.1) is sharp. In the same context, the logarithmic Sobolev inequality, in its "L1 form," is the statement that f log f dµ ≤ 1 2 ∇f2 f dµ + kfkL1(µ) log kfkL1(µ), f ∈ C 1(Rn), f > 0 (1.2) or equivalently, in the perhaps more familiar "L2 form", L2(µ) log kfkL2(µ), f2 log f dµ ≤ ∇f2 dµ + kfk2 f ∈ C 1(Rn) (1.3) where the equivalence follows by replacing f by f2 or vice versa. The ear- liest known version of this inequality is due to A.J. Stam [35], with another version discovered independently by P. Federbush [12]. The form given here was obtained by L. Gross [16], who coined the name. Gross also showed, at the level of Markovian semigroups, that (1.2) and (1.1) are equivalent. For instance, (1.3) can be formally obtained from (1.1) by setting p = 2, q = 1 + e2t, so that t(p, q) = t, and differentiating at t = 0. In 1983, S. Janson [22] discovered a fascinating phenomenon of hyper- contractivity in a complex setting. Consider (1.1) with n = 2 and identify R2 with C. Janson showed that if we restrict the inequality (1.1) to the space H of holomorphic functions, then we obtain the following improvement: ke−tAfkLq(µ) ≤ kfkLp(µ), t ≥ tJ (p, q) := 1 2 log(cid:18) q p(cid:19) , In this result, the critical time tJ := 1 smaller than Nelson's time, so integrability improves faster if the initial f ∈ H ∩ Lp(µ), 0 < p ≤ q < ∞. 2 log(cid:16) p q(cid:17) ("Janson's time") is strictly (1.4) 3 function f is holomorphic. Moreover, Janson's result has content even if p = 1 or 0 < p < 1. Inequalities of this form have come to be called (complex) strong hypercontractivity. For alternate proofs, extensions (including to Cn), and related results, see [5, 23, 40]. Part of the reason for this strengthening in the holomorphic case is that, since holomorphic functions are harmonic, the action of A on holomorphic functions reduces to that of the first-order operator Ef (x) = x·∇f (x), which is simply the generator of dilations on Cn = R2n. This idea was pursued by Graczyk, Kemp, and Loeb in [14, 15], in which they chose to explicitly consider the behavior of the dilation semigroup e−tE on real Euclidean space Rn. In this setting, the holomorphic functions are replaced by the log- subharmonic (LSH) functions; i.e. those nonnegative functions f for which log f is subharmonic. (This is effectively a generalization: when n is even and f is holomorphic, then f is log-subharmonic.) In the case of Gaussian measure µ, they proved the following version of strong hypercontractivity: ke−tE fkLq(µ) ≤ kfkLp(µ), t ≥ tJ (p, q), f ∈ LSH ∩ Lp(µ), 0 < p ≤ q < ∞. (1.5) They also obtained a corresponding strong logarithmic Sobolev inequality: f log f dµ ≤ 1 2 Ef dµ + kfkL1(µ) log kfkL1(µ), f ∈ LSH. (1.6) The classical logarithmic Sobolev inequality (1.2) is a key ingredient in their proof; indeed, (1.2) implies (1.6) rather directly. More generally, Graczyk, Kemp and Loeb proved, for a wider class of measures µ, that the statements (1.5) and (1.6)1 are equivalent. For instance, as with (1.1) and (1.2), one may formally obtain (1.6) from (1.5) by taking p = 1, q = e2t and differentiating at t = 0. In some cases, the hypothesis f ∈ LSH∩Lp(µ) in (1.5) must be strength- ened to f ∈ LSH ∩ Lq(µ); they call this statement partial strong hypercon- tractivity. We discuss this subtle issue in Remark 1.4, later in this paper. Another line of research inspired by Janson's strong hypercontractivity (1.4) was to study the phenomenon in non-Euclidean settings. In the papers [17, 18], L. Gross considered the case of a complex Riemannian manifold M equipped with an arbitrary smooth probability measure µ, where the Ornstein -- Uhlenbeck operator A is taken to be the generator of the Dirichlet 1Here, and for the rest of this section, the inequalities stated above should be read as including appropriate constants in the obvious places. 4 form E(f ) = ´M ∇f2 dµ. Gross showed, under certain assumptions, that if (M, µ) satisfies the logarithmic Sobolev inequality (1.2) then it satisfies the strong hypercontractivity property (1.4). In this context, it still happens that A reduces, on holomorphic functions, to a first-order vector field, whose geometric and complex-analytic properties become crucially important. In the paper [11], L. Gross, L. Saloff-Coste and the present author were interested in extending the complex Riemannian results of [17, 18] into a complex sub-Riemannian setting. We replaced the complex Riemannian manifold M with a stratified complex Lie group G equipped with a left- invariant sub-Riemannian geometry, taking the measure µ to be the hy- poelliptic heat kernel associated to this geometry. A relevant feature of stratified Lie groups is that, like Euclidean space, they admit a canonical group of dilations. In this setting, the Ornstein -- Uhlenbeck operator A fails to be holomorphic, so we studied instead its L2(µ)-orthogonal projection B onto the holomorphic functions, which, we showed, coincides with the vector field E generating the dilations. We were able to show that, if the logarith- mic Sobolev inequality (1.2) holds, then complex strong hypercontractivity (1.4) holds with the operator B in place of A. Of course, in retrospect, this statement is really (1.5) for holomorphic functions. We remark in passing that the logarithmic Sobolev inequality (1.2) is known to hold for a few stratified complex Lie groups (specifically, the com- plex Heisenberg -- Weyl groups), but it is not currently known whether it holds for all of them. The aim of the present paper is, in a sense, to unify [11] with [14, 15] by considering statements akin to (1.5) (in its "partial" form) and (1.6), in the setting of a real stratified Lie group G, again equipped with a left- invariant sub-Riemannian geometry and the associated hypoelliptic heat kernel measure. Our main Theorem 1.1 is, roughly, that (1.5) and (1.6) are equivalent in any stratified Lie group G, and if G satisfies the logarithmic Sobolev inequality (1.2) then (1.5) and (1.6) are both true. Again, we stress that (1.2) is known to hold for some stratified Lie groups (specifically, the H-type groups), but it is not currently known whether it holds for all of them; see Remark 1.2 below. In our view, stratified Lie groups are a natural setting in which to gen- eralize (1.5), (1.6), since the dilation structure of a stratified Lie group is perhaps the most direct generalization of the dilation structure of Euclidean space. Rather than considering more general measures µ as in [14, 15], we have chosen to restrict our attention to the canonical hypoelliptic heat kernel measure: partly because it is the natural generalization of Gaussian measure in this setting, and partly because we need to make use of strong 5 heat kernel estimates from the literature (Theorem 2.14 below). 1.2 Statement of results We briefly summarize the notation required to state our results. Complete definitions are given in Sections 2 and 3 below. Let G be a stratified Lie group equipped with a left-invariant sub- Riemannian metric h·,·i, for which the horizontal space is given by the first layer of the stratification of the Lie algebra of G. Let m be some normaliza- tion of Haar (Lebesgue) measure on G. We denote by ∇ and ∆ the canonical sub-gradient and sub-Laplacian induced by the metric, and by ρs the hy- poelliptic heat kernel for ∆ at time s > 0. A function f ∈ C 2(G) is said to be log-subharmonic (LSH) if f > 0 and ∆ log f ≥ 0 (we discuss alternative formulations in Section 8). We denote by E the vector field which generates the canonical dilations δr of the group G, and by e−tE f = f ◦ δe−t the corresponding operator semigroup. The Lp norms and spaces in the following statements are taken with respect to the heat kernel probability measure ρs dm at some fixed time s. We let Lp+ =Sq>p Lq, and write f ∈ W 1,p+ if f,∇f ∈ Lp+, where · is the norm induced by the metric h·,·i. The aim of this paper is to study the relationship between the following three statements, for fixed constants c, β ≥ 0. The time parameter s > 0 may be taken as arbitrary; each of the following statements holds for one s > 0 iff it holds for all s > 0, with the same constants c, β (see Remark 2.20 below). • The classical logarithmic Sobolev inequality: cs 2 G ∇f2 f G f log f ρs dm ≤ ρs dm + kfkL1 log kfkL1 + βkfkL1, f ∈ C 1(G), f ≥ 0 (LSI) We have stated this in its "L1 form". By replacing f by f 2, one can see that (LSI) is equivalent to the "L2 form": G f 2 log f ρs dm ≤ csG ∇f2 ρs dm +kfk2 L2 log kfkL2 + f ∈ C 1(G) β 2kfk2 L2, (L2-LSI) 6 The original formulation (1.2) of the logarithmic Sobolev inequality corresponds to taking β = 0. When β > 0, (LSI) is sometimes referred to as a "defective logarithmic Sobolev inequality". As remarked in Section 1.1, (LSI) is well known to be equivalent to the hypercontractivity of the Ornstein -- Uhlenbeck semigroup of G. • The strong logarithmic Sobolev inequality: G f log f ρs dm ≤ cG Ef ρs dm + kfkL1 log kfkL1 + βkfkL1, f ∈ LSH ∩ W 1,1+ (sLSI) The name "strong logarithmic Sobolev inequality" comes from [14]. However, in our present context, we show in Theorem 7.1 that (LSI) implies (sLSI), so (sLSI) is in fact logically weaker. Of course, (sLSI) applies to a much smaller class of functions. • (Partial) strong hypercontractivity: ke−tEfkLq ≤ M (p, q)kfkLp, t ≥ tJ (p, q), f ∈ LSH ∩ Lq, 0 < p ≤ q < ∞ (sHC) where M (p, q) := exp(β · (p−1 − q−1)), tJ (p, q) := c log(cid:18) q p(cid:19) . (1.7) Here tJ (p, q) is Janson's time. Note that the word "contractivity" is more apt when β = 0, since in that case, M (p, q) = 1 and (sHC) says that e−tE is a contraction from a subset of Lp into Lq. The word "partial" comes from [15] and refers to the hypothesis f ∈ LSH ∩ Lq (rather than Lp); see Remark 1.4 below. In comparing these statements to [14, 15], note that our c is their c 2 . The main results of this paper can be summarized as follows: Theorem 1.1. In any stratified Lie group G, the statements (sLSI) and (sHC) are equivalent. If G satisfies (LSI), then (sLSI) and (sHC) are both satisfied. That is, (LSI) =⇒ (sLSI) ⇐⇒ (sHC). The implication (LSI) =⇒ (sLSI) is Theorem 7.1; (sHC) =⇒ (sLSI) is Theorem 7.2; and (sLSI) =⇒ (sHC) is Theorem 7.6. 7 Remark 1.2. It is an open problem to determine which stratified Lie groups satisfy the logarithmic Sobolev inequality (LSI), and we hope this paper may provide additional motivation for further work on this difficult question. The current state of the art, as far as we are aware, is that (LSI) is true for H-type groups ([9, 21]; see Example 4.3 below for definitions and references), and of course in the "step 1" Euclidean case (Example 4.1). In all other stratified Lie groups, including all those of step ≥ 3, it is apparently unknown whether (LSI) holds or not. Corollary 1.3. If G is an H-type group, then (sLSI) and (sHC) are both true. Remark 1.4. In the statement (sHC), the hypothesis f ∈ Lq may seem somewhat unnatural, given that the result is to bound the Lq norm by the Lp norm. It is reasonable to conjecture that if (sHC) holds for all f ∈ LSH∩Lq then in fact it holds for all f ∈ LSH ∩ Lp. However, the obvious density argument is not available, because we do not know in this setting whether LSH ∩ Lq(ρs) is dense in LSH ∩ Lp(ρs). (Density arguments in stratified Lie groups can be subtle; for example, it is shown in [28, Proposition 8] that polynomials are dense in L2(ρs) only for groups of step m ≤ 4.) This issue arose, in the Euclidean setting, in the work of Graczyk, Kemp and Loeb [15]; our (sHC) is the analogue of the statement "partial strong hypercontractivity" appearing in their Theorem 1.17.1(b). The stronger statement, requiring only f ∈ LSH ∩ Lp (in our notation), is their The- orem 1.17.1(a); but they are able to show this only under significantly stronger assumptions on the measure, one of which is that the measure be α-subhomogeneous for some α ≤ 1/c, where c is our constant in the strong logarithmic Sobolev inequality (sLSI) (recall that our c is their c 2 ). This approach does not appear to succeed in our stratified Lie group setting, given current technology. For instance, if we consider the Heisenberg group H3, we can use the heat kernel estimates of [8, 26] to see that the heat kernel is α-subhomogeneous only for α ≥ 2. However, current methods for proving (LSI) in this setting produce a constant c > 1 2 (see [3, Sections 1.2 and 6.1] in conjunction with [7, Theorem 1.6]), although we do not know whether this is sharp. The situation for H-type groups is similar [8, 9, 21, 27]. 8 2 Stratified Lie groups and hypoelliptic heat ker- nels In this section, we review the standard definitions and properties of strati- fied Lie groups (also known as Carnot groups), and of the sub-Riemannian geometry and hypoelliptic heat kernels on these groups. The material in this section is adapted from [11]. Some motivating examples are discussed in Section 4. 2.1 Stratified Lie groups A comprehensive reference on stratified Lie groups is [4]. Definition 2.1. Let g be a finite-dimensional real Lie algebra. We say g is stratified of step m if it admits a direct sum decomposition g = Vj mMj=1 (2.1) and [V1, Vj] = Vj+1, [V1, Vm] = 0. A finite-dimensional real Lie group G is stratified if it is connected and simply connected and its Lie algebra g is stratified. Stratified Lie groups are also known as Carnot groups. Various equiva- lent definitions can be found in [4, Chapters 1 and 2]. As a trivial example, Euclidean space Rn with its usual addition is a (commutative) stratified Lie group of step 1. (Here the Lie bracket is sim- ply 0.) The simplest nontrivial example of a stratified Lie group is the Heisenberg group H3, which has step 2. See Section 4 below for further discussion of these and other examples. It is easy to check that a stratified Lie group G is necessarily nilpotent, and thus diffeomorphic to its Lie algebra g via the exponential map. In particular, a stratified Lie group is diffeomorphic to Euclidean space Rn as a smooth manifold (though certainly not isomorphic to Rn as a Lie group). Notation 2.2. Let Lx, Rx : G → G denote the left and right translation maps Lx(y) = xy and Rx(y) = yx. Notation 2.3. Let e denote the identity element of G. We identify the Lie algebra g with the tangent space TeG. For ξ ∈ g, leteξ,bξ denote, respectively, the unique left-invariant and right-invariant vector fields on G with eξ(e) = bξ(e) = ξ. 9 Notation 2.4. Being a connected nilpotent Lie group, G is unimodular, so it has a bi-invariant Haar measure which is unique up to scaling. For our purposes, there is no particular natural choice of scaling, so from now on m will denote some fixed Haar measure on G. Integrals like ´G f (x) dx will denote Lebesgue integrals with respect to m. It is easy to verify that the Haar measure on G is the push-forward under the exponential map of Lebesgue measure on g. Notation 2.5. Convolution on G is defined by (f ∗ g)(x) = G f (xy−1)g(y)dy = G f (z)g(z−1x)dz (2.2) when the integral exists. the integral sign, we obtain the identities Suppose ϕ ∈ C 1 eξg(x) = d dtt=0 g(xetξ) and bξg(x) = d eξ[ϕ ∗ f ] = ϕ ∗ (eξf ), c (G), f ∈ C 1(G) and ξ ∈ g. By using the formulas dtt=0 g(etξx) and differentiating under bξ[ϕ ∗ f ] = (bξϕ) ∗ f. 2.2 The dilation semigroup (2.3) Definition 2.6. For λ ≥ 0, the dilation map δλ on g is defined by δλ(v1 + ··· + vm) = λjvj vj ∈ Vj j = 1, . . . , m. (2.4) mXj=1 By an abuse of notation, we will also use δλ to denote the corresponding map on G defined by δλ(exp(v)) = exp(δλ(v)). It is straightforward to verify that for each λ > 0, the dilation δλ on g is an automorphism of the Lie algebra, and the dilation δλ on G is an automorphism of the Lie group. Also, δλµ = δλ ◦ δµ λ, µ ≥ 0. (2.5) Moreover, the derivative at the identity of δλ : G → G is (δλ)∗ = δλ : g → g. Since δλ is a group automorphism, we have the identity δλ ◦ Lx = Lδλ(x) ◦ δλ. we have Hence if ξ ∈ g = TeG and eξ is the corresponding left-invariant vector field, (δλ)∗eξ(x) = (δλ)∗(Lx)∗ξ = (δλLx)∗ξ = (Lδλ(x)δλ)∗ξ = (Lδλ (x))∗(δλ)∗ξ. 10 In particular, if ξ ∈ Vj, then (δλ)∗ξ = λjξ and so or in other words (δλ)∗eξ(x) = λj(Lδλ(x))∗ξ = λjeξ(δλ(x)) eξ(f ◦ δλ) = λj(eξf ) ◦ δλ. (2.6) (2.7) The dilation structure is a fundamental property of stratified Lie groups, and since the aim of this paper is to generalize results on the dilation in Eu- clidean space, stratified Lie groups are a natural setting to consider. Indeed, in a certain sense, we are studying what happens if we are allowed to dilate at different rates in different directions (linearly in V1 directions, quadratically in V2 directions, and so on). Definition 2.7. We define the dilation vector field or Euler vector field E on G by d dr(cid:12)(cid:12)(cid:12)r=0 (Ef )(x) = f (δer (x)) f ∈ C 1(G). (2.8) The main object of study in this paper is the one-parameter semigroup of dilation operators e−tEf = f ◦ δe−t, t ≥ 0. (2.9) Notation 2.8. Let ξj,k be a basis for g = G adapted to the stratification {Vj}, so that {ξj,k : 1 ≤ k ≤ dim Vj} is a basis for Vj. Then each x ∈ G can be written uniquely as x = exp(cid:16)Pm a smooth system of coordinates on G. j=1Pdim Vj k=1 xj,kξj,k(cid:17), so that xj,k is In this system of coordinates, we have E = mXj=1 dim VjXk=1 jxj,k ∂ ∂xj,k . (2.10) In the Euclidean case m = 1, we have E =P xk pointing radially away from the origin with magnitude x. Notation 2.9. The homogeneous dimension of g or G is ∂xk ∂ = x · ∇, a vector field D = mXj=1 j dim Vj. 11 We note that δλ scales the Lebesgue measure m by m(δλ(A)) = λDm(A). Thus for an integrable function f , we have G f ◦ δλ dm = λ−D G f dm. (2.11) (2.12) To conclude this subsection, we observe that the vector fields discussed above can be expressed in terms of each other in a well-behaved manner. Define the adapted basis {ξj,k} for g and the coordinates {xj,k} on G as in Notation 2.8. For each j, k, the vector fields 2.3) coincide at the identity but in general nowhere else. ∂xj,k ∂ , fξj,k, cξj,k (see Notation Lemma 2.10. We can write ∂ ∂xj,k + mXα=j+1 dim VαXβ=1 aα,β j,k ∂ ∂xα,β fξj,k = and (2.13) (2.14) ∂ ∂xj,k = fξj,k + mXα=j+1 dim VαXβ=1 bα,β j,k fξj,k where the coefficient functions aα,β xα,β. j,k , bα,β j,k are polynomials in the coordinates We can likewise express {cξj,k} and { ∂ polynomial coefficients, as well as {fξj,k} and {cξj,k}. Proof. By the Baker -- Campbell -- Hausdorff formula, in the coordinates {xj,k}, the group operation on G has the form ∂xj,k} in terms of each other, with (xy)j,k = xj,k + yj,k + Rj,k(x, y) where Rj,k is a polynomial which only depends on the coordinates xα,β, yα,β with α < j. See [4, Proposition 2.2.22 (4)] for details. Then (2.13) follows dt(cid:12)(cid:12)t=0x exp(tξj,k). We then obtain (2.14) by . (Note, for instance, that from (2.13) we ∂xj,k ∂ , so that (2.14) is trivially satisfied when j = m. One can solving the system (2.13) for immediately, since fξj,k(x) = d have gξm,k = ∂ ∂xm,k then proceed by downward induction on j.) 12 substitute this into (2.13). An identical argument applies to {cξj,k} and { ∂ terms of {cξj,k}, first write {cξj,k} in terms of { ∂ cj,kfξj,k =Xj,k E =Xj,k cj,kcξj,k Corollary 2.11 (See also [28, Lemma 4]). We can write ∂xj,k}. To write fξj,k in ∂xj,k} as just noted, and then (2.15) where the coefficient functions cj,k, cj,k are polynomials in the xα,β coordi- nates. Proof. Substitute (2.14) into (2.10). 2.3 Sub-Riemannian geometry on G In this subsection, we review some facts about the sub-Riemannian geom- etry of a stratified Lie group G and its hypoelliptic sub-Laplacian. For background on the general notions of sub-Riemannian geometry, see [29, 33, 36, 37]. Fix an inner product h·,·i on V1 ⊂ g. From now on, when we speak of a stratified Lie group G, we really mean a triple (G,h·,·i, m), including a choice of inner product on V1 and a choice of normalization for the Haar measure. Objects such as the sub-Laplacian, heat kernel, etc, which we discuss below, are not really intrinsic to the Lie group G, but depend on the choice of h·,·i and m. The inner product gives rise to a sub-Riemannian geometry on the smooth manifold G in the following way. For x ∈ G, let Hx = (Lx)∗V1 ⊂ TxG, so that H is a left-invariant subbundle of the tangent bundle T G. This H is called the horizontal bundle or horizontal distribution. Then the inner product h·,·i on V1 induces, by left translation, an inner product h·,·ix on Hx, which is a left-invariant sub-Riemannian metric on G. We may drop the subscript x when no confusion will arise. Since V1 generates g, the horizontal bundle H satisfies Hormander's bracket generating condition. We will use · to denote the norm on Hx induced by h·,·i. This metric gives rise to a canonical left-invariant sub-Laplacian ∆ on G, which is easiest to define in terms of a basis. Let ξ1, . . . , ξn be an orthonormal basis for V1, and let 2 2 ∆ = eξ1 + ··· + eξn 13 (2.16) where, as in Notation 2.3, eξi is the extension of ξi to a left-invariant vector field on G. It is easy to check this definition is independent of the basis chosen. Since H satisfies the bracket generating condition, the operator ∆ is hypoelliptic [20]. c (G), is essentially self-adjoint on L2(G, m). It is shown in [39] that ∆, with domain C ∞ As a consequence of (2.7), we have ∆[f ◦ δλ] = λ2(∆f ) ◦ δλ. Likewise, if es∆/4 is the heat semigroup for ∆, we have es∆/4[f ◦ δλ] = (esλ2∆/4f ) ◦ δλ. (2.17) (2.18) Much more information about the sub-Laplacian can be found in [4]. We may also define the sub-gradient ∇ by ∇f (x) = nXi=1 (eξif )(x)eξi(x) ∈ Hx, f ∈ C 1(G). (2.19) This too is well-defined independent of the chosen basis. Finally, let d be the Carnot -- Carath´eodory distance induced by the sub- Riemannian metric (see [4, Section 5.2]). Intuitively, d(x, y) is the length of the shortest horizontal path joining x and y. The Chow -- Rashevskii and ball-box theorems [29, 30] imply that d(x, y) is finite and that d is a metric which induces the manifold topology on G (which in turn is just the Eu- clidean topology on the finite-dimensional vector space G = g). A straight- forward computation shows that d is left-invariant with respect to the group structure on G: and invariant with respect to the inverse: d(x, y) = d(zx, zy), x, y, z ∈ G d(e, x−1) = d(e, x) and also homogeneous with respect to the dilation δλ: d(δλ(x), δλ(y)) = λd(x, y). See [4, Propositions 5.2.4 and 5.2.6] for details. (2.20) (2.21) (2.22) 14 2.4 Properties of the heat kernel It is shown in [39] that the Markovian heat semigroup es∆/4 admits a right convolution kernel ρs, i.e. es∆/4f = f ∗ ρs; it is also shown that ρs is C ∞ and strictly positive. This function ρs is the (hypoelliptic) heat kernel associated to (G,h·,·i, m). Since es∆/4 is Markovian, the heat kernel measure ρs dm is a probability measure. In this subsection, we collect several properties of the heat kernel from the literature. Notation 2.12. For s > 0 and 0 < p ≤ ∞, we write Lp(ρs) as short for Lp(G, ρs dm). Let Lp+(ρs) := [q>p Lp−(ρs) := \q<p Lq(ρs) Lq(ρs). We will say fn → f in Lp+(ρs) (respectively, in Lp−(ρs)) if fn → f in Lq(ρs) for some q > p (respectively, for all q < p). Also, W 1,p+(ρs) will denote the space of functions f with f,∇f ∈ Lp+(ρs). We shall only have occasion to deal with C 1 functions in W 1,p+, so we do not discuss weak derivatives here. Notice that Lp+(ρs) and Lp−(ρs) are vector spaces, and L∞−(ρs) is an algebra. By Holder's inequality, if f ∈ Lp+(ρs) and g ∈ L∞−(ρs) then f g ∈ Lp+(ρs). We also note that if f ∈ L1(ρs) is positive and bounded away from zero then log f ∈ L∞−(ρs). Lemma 2.13. The heat kernel ρs is invariant under the group inverse op- eration: we have ρs(x) = ρs(x−1). Proof. See [34, Theorem III.2.1 (4)] or the discussion in [6, Proposition 3.1 (3)]. We shall need to make use of sharp upper and lower estimates for the heat kernel. Theorem 2.14. For each 0 < ǫ < 1 there are constants C(ǫ), C ′(ǫ) such that for every x ∈ G and s > 0, C(ǫ) m(B(e,√s)) e−d(e,x)2/(1−ǫ)s ≤ ρs(x) ≤ (2.23) where m(B(e,√s)) is the Lebesgue (Haar) measure of the d-ball centered at the identity (or any other point of G) with radius √s. m(B(e,√s)) C ′(ǫ) e−d(e,x)2/(1+ǫ)s 15 Proof. The upper bound is Theorem IV.4.2 of [39]. The lower bound is Theorem 1 of [38]. Note that our choice to consider the semigroup es∆/4 rather than es∆ accounts for a missing factor of 4 in the exponents compared to the results stated in [38, 39]. Corollary 2.15. Any polynomial p in the coordinates xj,k (see Notation 2.8) is in L∞−(ρs). Proof. It suffices to consider p(x) = xr j,k for some fixed j, k, r. Let S = {x : d(e, x) = 1} be the unit sphere of d, and let C = supS p which is finite by the compactness of S. Then the inequality p(x) ≤ Cd(e, x)rj holds trivially on S. The scaling relations p(δλ(x)) = λrjp(x) and (2.22) now imply that p(x) ≤ Cd(e, x)rj for all x ∈ G. The result now follows, via massive overkill, from the upper bound in Theorem 2.14. Corollary 2.16. If ∇f ∈ Lp+(ρs) then Ef ∈ Lp+(ρs). Proof. Combine Corollary 2.11 with Corollary 2.15. Lemma 2.17. (Special case of [39, Theorem IV.3.1]) Let r ≥ 0 and 0 < s < t < ∞. There is a constant C, depending on r, s, t, such that for all y ∈ G, sup d(e,x)≤r ρs(xy) ≤ Cρt(y). Proof. Replacing x, y by x−1, y−1 and using Lemma 2.13 and (2.21), it is enough to show the result for ρs(yx) instead of ρs(xy). Let x ∈ B(e, r) be arbitrary. By the bounds in Theorem 2.14, we have ρs(yx) ρt(y) ≤ C ′′(s, t, ǫ) exp(cid:18)−(cid:18) d(e, yx)2 (1 + ǫ)s − d(e, y)2 (1 − ǫ)t(cid:19)(cid:19) . By the left invariance of the distance d and the triangle inequality, we have d(e, yx) ≥ d(e, y) − r, which yields ρs(yx) 1 ρt(y) ≤ C ′′(s, t, ǫ) exp(cid:18)−d(e, y)2(cid:18) ≤ C ′′(s, t, ǫ) exp(cid:18)−d(e, y)2(cid:18) (1 + ǫ)s − 1 (1 + ǫ)s − 1 (1 − ǫ)t(cid:19) + (1 − ǫ)t(cid:19) + 1 2rd(e, y) (1 + ǫ)s − 2rd(e, y) (1 + ǫ)s(cid:19) . r2 (1 + ǫ)s(cid:19) Now since s < t, we can take ǫ sufficiently small that we now choose some r0 with 1 (1+ǫ)s − 1 (1−ǫ)t > 0. If r0 >(cid:18) 1 (1 + ǫ)s − 1 (1 − ǫ)t(cid:19)−1 2r (1 + ǫ)s 16 then for all y with d(e, y) ≥ r0, the exponent is negative and we have ρs(yx) ρt(y) ≤ C ′′(s, t, ǫ). This suffices, since by continuity the supremum over all y ∈ B(e, r0) is finite. Lemma 2.18. Let s > 0, p > 1 and 0 < t0 ≤ t1 < p p−1 s. Then sup t∈[t0,t1] ρt ρs ∈ Lp(ρs). (cid:12)(cid:12)(cid:12)(cid:12) ρs(x)(cid:12)(cid:12)(cid:12)(cid:12) Proof. For any t ∈ [t0, t1] we have by Theorem 2.14 that ρt(x) p ρs(x) = ≤ ≤ ρt(x)p ρs(x)p−1 C ′(ǫ)pm(B(e,√s))p−1 C(ǫ)p−1m(B(e,√t))p C ′(ǫ)pm(B(e,√s))p−1 exp(cid:18)−(cid:18) C(ǫ)p−1m(B(e,√t0))p exp(cid:18)−(cid:18) (1+ǫ)t1 − p−1 (1 + ǫ)t1 − p t1 − p − 1 (1 − ǫ)s = p lim ǫ→0 p − 1 s > 0 p (1 + ǫ)t − p (1 + ǫ)t1 − p − 1 (1 − ǫ)s(cid:19) d(e, x)2(cid:19) (1 − ǫ)s(cid:19) d(e, x)2(cid:19) . p − 1 The right side is independent of t, and will be integrable on G (with respect to m) provided that (1−ǫ)s > 0. But since p we can choose ǫ sufficiently small that this coefficient is indeed positive. Lemma 2.19. The heat kernel ρs obeys the scaling relation ρs(δλ(y)) = λ−Dρsλ−2(y). (2.24) Proof. This follows from the corresponding scaling properties of the semi- group es∆/4 (2.18) and of the Haar measure m (2.11). Remark 2.20. Using (2.7) and (2.24), one may verify, by replacing f by an appropriate dilation f ◦ δr, that each of the statements (LSI), (sLSI), (sHC) in our main theorem holds for one s > 0 iff it holds for all s > 0, with the same constants c, β. Lemma 2.21. Suppose f ∈ C 2(G) ∩ L1(ρs) and Ef,∇f, ∆f ∈ L1(ρs). Then Ef ρs dm = ∆f ρs dm. (2.25) G s 2 G Moreover, the same result holds if we assume ∆f ≥ 0 instead of ∆f ∈ L1(ρs). 17 Proof. Suppose first that f ∈ C ∞ c (G). By (2.18) we have G f ◦ δer ρs dm = G f ρse2r dm. Differentiating under the integral sign at r = 0, we obtain G s ρs dm = f d ds Ef ρs dm = 2sG 2 G Now to show the general case, let φ ∈ C ∞ c (G) be a cutoff function with φ = 1 on a neighborhood of the identity e, and set φn = φ ◦ δ1/n. Then it is easy to check that f ∆ρs dm = 2 G ∆f ρs dm. s 1 n ∇φn = Hence we have, pointwise and boundedly, (∇φ) ◦ δ1/n, ∆φn = 1 n2 (∆φ) ◦ δ1/n, Eφn = (Eφ) ◦ δ1/n φn → 1, ∇φn → 0, ∆φn → 0, Eφn → 0, the last following from the fact that Eφ = 0 on a neighborhood of e. Ap- plying our result to φnf , we have Eφn · f ρs dm +G G = G ∆φn · f ρs dm +G φn · Ef ρs dm g(∇φn,∇f ) ρs dm +G φn · ∆f ρs dm. By dominated convergence, using the integrability assumptions on f and its derivatives, letting n → ∞ gives the desired identity. If we only assume ∆f ≥ 0, then if we choose φn with a little more care, we can get φn → 1 monotonically. Then we can repeat the argument above, in which we have ´g φn · ∆f ρs dm → ´G ∆f ρs dm by monotone convergence instead of dominated convergence. 3 Log-subharmonic functions In general, a function f : G → [0,∞) on G is said to be log-subharmonic (LSH) if log f is subharmonic with respect to the sub-Laplacian ∆. There are many possible notions of subharmonicity in this setting. In this paper, we shall work primarily with a strong "classical" notion of subharmonicity, in order to avoid obscuring the main ideas with technicalities; but see Section 8 below, where we discuss how the results of this paper can be applied to functions which are log-subharmonic in a weaker sense. 18 Definition 3.1. Suppose f ∈ C 2(G). We will say f is subharmonic if ∆f ≥ 0. We will say f is log-subharmonic (LSH) if f > 0 and ∆ log f ≥ 0. Lemma 3.2. If f ∈ C 2(G) and f > 0, then f is LSH if and only if ∆f ≥ ∇f2 f . (3.1) In particular, LSH functions are subharmonic. Proof. By the chain and product rules, ∆ log(f ) = −∇f2 f 2 + so that ∆ log(f ) ≥ 0 iff ∆f ≥ ∇f 2 f . ∆f f = 1 f (cid:18)−∇f2 f + ∆f(cid:19) Proposition 3.3. Suppose f, g are LSH. The following functions are LSH: 1. Positive constants 2. f g 3. f p for any p > 0 4. f + g 5. f ◦ δλ for any λ > 0 Proof. Items 1 -- 3 are immediate. For item 4, we use a trick suggested in [14, Proposition 2.2]. We have that u = log f and v = log g are subharmonic. Fix x ∈ G and assume without loss of generality that ∆u(x) ≥ ∆v(x). Now ∆ log(f + g) = ∆ log(eu + ev) = ∆[v + log(eu−v + 1)] ≥ ∆ log(eu−v + 1). Let ψ(t) = log(et + 1) and note that ψ′(t) = et (1+et)2 > 0. By the chain and product rules, we have ∆ log(eu−v + 1) = ∆ψ(u − v) = ψ′′(u − v)∇[u − v]2 + ψ′(u − v)∆[u − v]. This is nonnegative at x since by assumption ∆u(x) ≥ ∆v(x). 1+et > 0 and ψ′′(t) = et Item 5 is an immediate consequence of (2.17) which implies that ∆ log(f ◦ δλ) = λ2(∆ log(f )) ◦ δλ. 19 Lemma 3.4. If f is LSH and ϕ ∈ C ∞ c (G) is nonnegative then ϕ∗ f is LSH. Proof. We will show that ϕ ∗ f satisfies (3.1). By rescaling, let us suppose without loss of generality that ´G ϕ dm = 1. Fix an orthonormal basis ξ1, . . . , ξn for V1. From (2.16) and (2.3), we have ∆[ϕ ∗ f ] = ϕ ∗ ∆f , and so Lemma 3.2 gives ∆[ϕ ∗ f ] = ϕ ∗ ∆f ≥ ϕ ∗ ∇f2 f = nXi=1 ϕ ∗ (eξif )2 f . Applying the multivariate Jensen inequality with the convex function ψ(u, v) = u2/v and the probability measure ϕ dm, we have ϕ ∗ f ! (x) = G (eξif )2 dy ϕ(y) f (y−1x) (eξif (y−1x))2 ≥ (cid:16)´G ϕ(y)eξif (y−1x) dy(cid:17)2 ´G ϕ(y)f (y−1x) dy (ϕ ∗ (eξif ))(x)2 (ϕ ∗ f )(x) (eξi[ϕ ∗ f ](x))2 (ϕ ∗ f )(x) = = using (2.3) again, since eξi is left-invariant. Thus we have = ∇(ϕ ∗ f )2 ∆[ϕ ∗ f ] ≥ nXi=1 (eξi[ϕ ∗ f ])2 ϕ ∗ f ϕ ∗ f and so by Lemma 3.2, ϕ ∗ f is LSH. 4 Examples and special cases 4.1 Euclidean space Example 4.1. As a trivial example, G = Rn with Euclidean addition is an (abelian) stratified Lie group of step 1. (Indeed, these are all the step 1 stratified Lie groups.) Here the Lie bracket is zero and the dilation is δλ(x) = λx. If we equip V1 = g = Rn with the Euclidean inner product, then the sub-Laplacian and sub-gradient are the usual Euclidean Laplacian and gradient, and the Carnot -- Carath´eodory distance d is Euclidean distance. 20 The heat kernel ρs is the Gaussian density, appropriately scaled. (Note that, in our normalization, standard Gaussian density corresponds to s = 2.) As such, the results of this paper include statements about Gaussian measure on Euclidean space, similar to those obtained in [14, 15]. It is well known that (LSI) is true for Gaussian measures [16], with constant c = 1 2 . 4.2 The Heisenberg group Example 4.2. The simplest nontrivial example of a stratified Lie group is the 3-dimensional real Heisenberg group G = H3, which we may realize as R3 equipped with the group operation (x1, x2, x3)(x′ 1, x′ 2, x′ 1, x2 + x′ 2, x3 + x′ 3 + (x1x′ 3) =(cid:18)x1 + x′ 1 2 1)(cid:19) . 2 − x2x′ (4.1) If we let ξi = ∂ ∂xi ∈ g = TeG for i = 1, 2, 3, the Lie bracket is given by [ξ1, ξ2] = ξ3, [ξ1, ξ3] = [ξ2, ξ3] = 0 so we have the decomposition g = V1 ⊕ V2 where V1 = span{ξ1, ξ2}, V2 = span{ξ3}. Thus the Heisenberg group is stratified of step 2. A natural inner product h·,·i on V1 is given by taking ξ1, ξ2 to be orthonormal. The corresponding left-invariant vector fields are given by eξ1 = ∂ ∂x1 − 1 2 x2 ∂ ∂x3 , ∂ ∂x2 + 1 2 x1 ∂ ∂x3 , eξ2 = ∂ ∂x3 . eξ3 = The Heisenberg group H3 was the first nontrivial stratified Lie group that was shown to satisfy the logarithmic Sobolev inequality (LSI). This statement can be found in [3] and follows from heat semigroup gradient bounds previously established in [25], via a variant of a standard Γ2-calculus argument from [2] or [1, pp. 69 -- 70]. A key ingredient is sharp upper and lower heat kernel estimates, obtained in [26]. As such, our Theorem 1.1 implies that (sLSI) and (sHC) are satisfied by H3 as well. The Heisenberg group construction immediately generalizes to the family of Heisenberg -- Weyl groups H2n+1, which is realized as R2n+1 with a group operation defined again by (4.1), where now we take x1, x2 ∈ Rn. groups, which we discuss next. The Heisenberg and Heisenberg -- Weyl groups are examples of H-type 21 4.3 H-type groups Example 4.3. Suppose that G is a (real) stratified Lie group of step 2. Let the inner product h·,·i on V1 be extended to an inner product on all of g = V1 ⊕ V2, still denoted by h·,·i, for which the decomposition g = V1 ⊕ V2 is orthogonal. For each z ∈ V2, define the linear map Jz : V1 → V1 by hJzv, wi = hz, [v, w]i. We say that (G,h·,·i) is H-type if, for each z ∈ V2 with hz, zi = 1, the map Jz is a partial isometry. H-type groups were introduced in [24]; see [4, Chapter 18] for more background on these groups. The Heisenberg and Heisenberg -- Weyl groups discussed in Example 4.2 are H-type (indeed, the H stands for Heisenberg). H-type groups satisfy the same type of heat semigroup gradient bounds as the Heisenberg group H3. This was shown independently in [9, 21]; for the required heat kernel estimates, see [8, 27]. Thus, such groups satisfy (LSI) as well, by the same general argument given in [3]. As we noted in Corollary 1.3, our Theorem 1.1 then implies that (sLSI) and (sHC) are also true in H-type groups. We do not know of any further examples of stratified Lie groups where (LSI) has been proved. 4.4 Complex stratified Lie groups Example 4.4. Suppose that G is a stratified Lie group which is also a complex Lie group, so that the Lie algebra g admits a complex structure J : g → g satisfying [Jv, w] = J[v, w]. Then g is a complex vector space and the subspaces Vi in the decomposition (2.1) are complex vector spaces as well. The complex structure on g induces a complex manifold structure on G for which the exponential map is holomorphic. In this setting, it is natural to ask that the inner product h·,·i on V1 be compatible with the complex structure, by being Hermitian: hJv, wi = −hv, Jwi. We call such G a complex stratified Lie group. A simple example is the complex Heisenberg group H3 C, or the complex Heisenberg -- Weyl groups H2n+1 . C Lemma 4.5. Suppose G is a complex stratified Lie group. Let f : G → C be holomorphic. Then for any ǫ > 0, the function g =pf2 + ǫ is LSH. We cannot say that f itself is LSH by our definition, because f need not be either C 2 nor strictly positive, but it is weakly LSH in the sense of Section 8; see Proposition 8.4. Proof. Let x ∈ G and suppose for the moment that f (x) 6= 0. By conti- nuity, there is a disk D ⊂ C \ {0} and an open neighborhood U of x such 22 that f (U ) ⊂ D. Let L(z) be a branch of the complex logarithm which is holomorphic on D. Then L ◦ f is holomorphic on U . We are assuming that the inner product on V1 is Hermitian, so the real and imaginary parts of any holomorphic function are harmonic with respect to the sub-Laplacian ∆ (see [11] for further details). Thus ∆ log f = ∆ Re L ◦ f = 0 on U . It follows that ∆ log g ≥ 0 on U (see Proposition 3.3 items 3 and 4). So we have shown ∆ log g ≥ 0 on {f 6= 0}. But since f is holomorphic, {f 6= 0} is dense in G (unless f ≡ 0 in which case the statement is trivial). Since g is strictly positive and C ∞, ∆ log g is continuous. Thus we have ∆ log g ≥ 0 everywhere. Corollary 4.6. Let G be a complex stratified Lie group satisfying (sHC). Then (sHC) also holds for all holomorphic f ∈ Lq(ρs). Proof. Apply (sHC) topf2 + ǫ, and let ǫ ↓ 0 using Lemma 4.5 and dom- inated convergence. In particular, by Theorem 1.1, this holds whenever G satisfies (LSI). This implication was one of the main results of [11], which also gave a density argument for holomorphic Lp that can be used to show that (sHC) also holds for holomorphic f ∈ Lp(ρs). See the related discussion in Remark 1.4. Unfortunately, we do not know of any similar density results for LSH functions in the real case. For a complex stratified Lie group G, the vector field E has an addi- tional significance: as shown in [11], it is the holomorphic projection of the Ornstein -- Uhlenbeck operator A. In the special case of Cn with the Gaussian heat kernel, if f is holomorphic then we actually have Af = Ef because the Laplacian term vanishes. For more general complex stratified groups, this is no longer true because Af may fail to be holomorphic, but its L2(ρs) projection onto the holomorphic functions equals Ef . There is not much overlap between the complex and H-type Lie groups; we showed in [10] that the only complex Lie groups which are also H-type are the complex Heisenberg -- Weyl groups H2n+1 . As such, these are the only complex stratified Lie groups for which (LSI) is currently known to hold. C 5 Convolution and approximation Lemma 5.1. If f ∈ Lp+(ρs), p ≥ 1, and ϕ ∈ Cc(G) then ϕ ∗ f ∈ Lp+(ρs). Proof. By considering positive and negative parts, we can assume without loss of generality that ϕ ≥ 0. Also, by rescaling we can assume ´G ϕ dm = 1. 23 Let q > p and r > 1 be so small that f ∈ Lqr(ρs). Let 1 r∗ = 1 r∗−1 s. Then by Lemma 2.18 we have (ρs). Next, let K be the support of ϕ, which is compact, and use and choose any t with s < t < rs = r∗ ρt ρs ∈ Lr∗ Lemma 2.17 to choose C such that supz∈K ρs(zy) ≤ Cρt(y) for all y ∈ G. see that To start, use Jensen's inequality with the probability measure ϕ dm to r + 1 (ϕ ∗ f )(x)q =(cid:12)(cid:12)(cid:12)(cid:12)G ϕ(z)f (z−1x) dz(cid:12)(cid:12)(cid:12)(cid:12) so that, by Fubini's theorem, q ≤ G ϕ(z)f (z−1x)q dz kϕ ∗ fkq Lq(ρs) ≤ K G = K G ϕ(z)f (z−1x)qρs(x) dx dz ϕ(z)f (y)qρs(zy) dy dz making the change of variables y = z−1x and using the translation invariance of m. Now for all z in the support K of ϕ, we have ρs(zy) ≤ Cρt(y). Since ´G ϕ(z) dz = 1 by assumption, we now have kϕ ∗ fkq Lq(ρs) ≤ C G f (y)qρt(y) dy = C G f (y)q ρt(y) ≤ Ckfkq ρs(y) Lqr(ρs)(cid:13)(cid:13)(cid:13)(cid:13) ρt ρs(cid:13)(cid:13)(cid:13)(cid:13)Lr∗ (ρs) ρs(y) dy by Holder's inequality. By our choices of t, q, r, both norms are finite. Lemma 5.2. Suppose f ∈ Lp+(ρs), p ≥ 1, and ϕ ∈ C ∞ ξ ∈ g, we have eξ[ϕ ∗ f ] ∈ Lp+(ρs). As a consequence, we also have using Lemma 2.10, we can write eξ =Pj,k cj,kcξj,k for some polynomials cj,k. Proof. By writing ξ as a linear combination of an adapted basis {ξj,k} and In particular, by Corollary 2.15 we have cj,k ∈ L∞−(ρs). Now using (2.3), we have ∇[ϕ ∗ f ] , ∆[ϕ ∗ f ], E[ϕ ∗ f ] ∈ Lp+(ρs). c (G). Then for any eξ[ϕ ∗ f ] =Xj,k cj,kcξj,k[ϕ ∗ f ] =Xj,k cj,k · [(cξj,kϕ) ∗ f ] which is in Lp+(ρs) by Lemma 5.1. 24 The desired statement for ∆[ϕ ∗ f ] follows by applying this twice to get eξ2[ϕ ∗ f ] ∈ Lp+(ρs), then summing over an orthonormal basis {ξi} for V1. For E[ϕ ∗ f ], use Corollary 2.16. Lemma 5.3. Suppose f ∈ L1+(ρs). There is a sequence of nonnegative ϕn ∈ C ∞ c (G) such that ϕn ∗ f → f almost everywhere and in L1+(ρs). If moreover f ∈ C(G) the convergence is uniform on compact sets. Proof. Let ϕ ∈ C ∞ c (G) be nonnegative with ´G ϕ dm = 1, and let ϕn = nDϕ◦ δn, so that ϕn is a sequence of standard mollifiers. It is standard that ϕn ∗ f → f almost everywhere, after passing to a subsequence if necessary, and that the convergence is uniform on compact sets if f is continuous. Now we note that the ϕn are all supported in some compact neighbor- hood K of the identity. As in the proof of Lemma 5.1, if we choose q, r > 1 such that f ∈ Lqr(ρs), then we can choose C, t, independent of n, such that Lqr(ρs)(cid:13)(cid:13)(cid:13)(cid:13) ρt ρs(cid:13)(cid:13)(cid:13)(cid:13)Lr∗ (ρs) kϕn ∗ fkq Lq(ρs) ≤ Ckfkq In particular, if 1 < q′ < q then {ϕn ∗ fq′ with respect to ρs dm, hence ϕn ∗ f converges in Lq′ : n ≥ 1} is uniformly integrable (ρs). 6 Differentiation under the integral sign As mentioned in Section 1, the strong logarithmic Sobolev inequality (sLSI) is essentially an infinitesimal version of the strong hypercontractivity in- equality (sHC). Thus, at a purely formal level, the equivalence between them is completely natural, and consists mainly of differentiating under the integral sign. The difficulty is to verify that this is justified. The following abstract lemma is a general principle for differentiating un- der the integral sign. We have not seen this particular form in the literature, so we give the proof. Lemma 6.1. Let (X, µ) be a probability space, and let F : [0, T ] × X → R be jointly measurable. Suppose that for each x ∈ X, we have F (·, x) ∈ C 1([0, T ]), so that ∂tF : [0, T ] → X is also jointly measurable. Furthermore, suppose that the family of functions {∂tF (t,·) : 0 ≤ t ≤ T} is uniformly in- tegrable on (X, µ). Set v(t) = X w(t) = X F (t, x) µ(dx) ∂tF (t, x) µ(dx). 25 (6.1) Then v ∈ C 1([0, T ]) and v′(t) = w(t) on [0, T ]. We use the term "uniformly integrable" here in the probabilist's sense: a family of functions G is uniformly integrable with respect to µ iff lim M →∞ sup g∈G g≥M g dµ = 0. This is the necessary and sufficient hypothesis for the Vitali convergence theorem. In particular, a uniformly integrable family is bounded in L1(µ). We also recall, for later use, the fact that if supg∈G kgkLp(µ) < ∞ for some p > 1 then G is uniformly integrable. Proof of Lemma 6.1. First, by the Vitali convergence theorem, w is contin- uous on [0, T ]. The uniform integrability also implies that ∂tF (t,·) is uniformly L1 bounded, so we have k∂tF (t,·)kL1(µ) ≤ K for some finite K. Thus for any 0 ≤ τ ≤ T we have τ 0 X ∂tF (t, x) µ(dx) dt ≤ KT < ∞. So by Fubini's theorem and the first fundamental theorem of calculus, we have τ 0 w(t) dt = τ 0 X = X τ = X = v(τ ) − v(0). 0 ∂tF (t, x) µ(dx) dt ∂tF (t, x) dt µ(dx) (F (τ, x) − F (0, x)) µ(dx) Hence by the second fundamental theorem of calculus, v is differentiable and v′ = w. The following lemma, which follows from the heat kernel bounds in Sec- tion 2.4, will be convenient in verifying the uniform integrability hypothesis for our applications. Lemma 6.2. Suppose f ∈ Lp+(ρs). Then for some q > p and any T < ∞ we have sup 0≤t≤T ke−tE fkLq(ρs) < ∞. 26 In particular, if f ∈ L1+(ρs), then {e−tE f : 0 ≤ t ≤ T} is uniformly integrable with respect to ρs dm. Proof. Choose q > p and r > 1 so small that f ∈ Lqr(ρs). Let r∗ = r r−1 . By Lemma 2.18, taking t0 = se−2T and t1 = s, we have sup0≤t≤T ρse−2t/ρs ∈ Lr∗ by (2.24) and (2.12) (ρs). Then for any t ∈ [0, T ] we have Lq(ρs) = G f ◦ δe−tqρs dm ke−tEfkq = G fqρse−2t dm ≤ G fq sup Lqr(ρs)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ kfkq sup 0≤t≤T ρse−2t 0≤t≤T ρs ! ρs dm ρs (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lr∗ (ρs) ρse−2t which is independent of t and finite. Lemma 6.3. Suppose r ∈ C 1([0, T ]) with 1 ≤ r(t) ≤ q, and suppose f ∈ W 1,q+(ρs) is positive. Set ft = e−tEf r(t). Then the functions ft, ft log ft, ∇ft, Eft, 0 ≤ t ≤ T are all uniformly bounded in Lp(ρs) norm for some p > 1. In particular, they are uniformly integrable. We note that the conclusion of this lemma implies ft ∈ W 1,1+(ρs) for each 0 ≤ t ≤ T . Proof. For ft, note that 1 + f q is in L1+, so by Lemma 6.2 we have that the family {e−tE [1 + f q] : 0 ≤ t ≤ T} is uniformly bounded in Lp for some p > 1. But f r(t) ≤ 1 + f q for each t, so ft ≤ e−tE[1 + f q], and ft is uniformly bounded in Lp. For ft log ft, note that since f q ∈ L1+, we also have (1+f q) log f q ∈ L1+. Then, since log f r(t) = r(t) log f ≤ q log f = log f q we have f r(t) log f r(t) ≤ (1 + f q) log f q. By the same argument as in the previous case, ft log ft is uniformly bounded in some Lp. 27 For ∇ft, note that ∇ft = r(t)e−tE[f r(t)−1](cid:12)(cid:12)∇[e−tEf ](cid:12)(cid:12) = r(t)e−te−tEhf r(t)−1∇fi using (2.7). Now f r(t)−1 ≤ 1 + f q−1, and we have (1 + f q−1)∇f ∈ L1+ by Holder's inequality. So by Lemma 6.2,(cid:8)e−tE(cid:2)(1 + f q−1)∇f(cid:3)(cid:9) is uniformly bounded in some Lp, p > 1, and the same thus holds for ∇ft. Since ∇f ∈ Lq+, we have Ef ∈ Lq+ as well, by Lemma 2.16. So a similar argument applies for Eft as for ∇ft, noting that Eft = r(t)e−tE[f r(t)−1]Ee−tE f = r(t)e−tEhf r(t)−1Efi . Lemma 6.4. Again suppose r ∈ C 1([0, T ]) with 1 ≤ r(t) ≤ q, and suppose f ∈ C 1(G) ∩ W 1,q+(ρs) is positive. Set ft = e−tE f r(t). Let v(t) = G w(t) = G ft ρs dm = ke−tEfkLr(t)(ρs) ∂tft ρs dm = G(cid:20)−Eft + r′(t) r(t) Then v ∈ C 1([0, T ]) and v′(t) = w(t) on [0, T ]. Proof. Differentiate under the integral sign using Lemma 6.1, with F (t, x) = ft(x). The continuity of ∂tft follows from the assumption that f ∈ C 1(G), and the uniform integrability hypothesis is verified by Lemma 6.3. ft log ft(cid:21) ρs dm. (6.2) 7 Proofs of the main results 7.1 LSI implies sLSI Theorem 7.1. In any stratified Lie group G, if (LSI) holds, then (sLSI) holds, with the same constants c, β. Proof. Suppose f ∈ LSH ∩ W 1,1+(ρs); by Corollary 2.16 we have Ef ∈ L1+(ρs). Since LSH functions are subharmonic (∆f ≥ 0), we can apply Lemmas 3.2 and 2.21 to obtain G ∇f2 f ρs dm ≤ G ∆f ρs dm = 2 s G Ef ρs dm. Inserting this inequality into (LSI) yields (sLSI). 28 7.2 sHC implies sLSI Theorem 7.2. In any stratified Lie group G, if (sHC) holds, then (sLSI) holds, with the same constants c, β. Proof. Suppose (sHC) holds with constants c, β. Fix f ∈ LSH ∩ W 1,1+(ρs). Set r(t) = et/c and choose T > 0 so small that f ∈ W 1,r(T )+(ρs). For t ∈ [0, T ], applying (sHC) with p = 1 and q = r(T ) yields ke−tE fkLr(t)(ρs) ≤ M (t)kfkL1(ρs) where M (t) := M (1, r(t)) = exp(β · (1 − e−t/c)). Define v(t), w(t) as in (6.2), and set α(t) = 1 M (t)ke−tE fkLr(t)(ρs) = 1 M (t) v(t)1/r(t). (7.1) (7.2) (7.3) Note that α(0) = kfkL1(ρs), so (7.1) says that α(t) ≤ α(0) for all t ∈ [0, T ]. Now applying Lemma 6.4 with q = r(T ), we have that v is continuously differentiable on [0, T ]; hence so is α(t). As such, we must have 0 ≥ α′(0) = −M ′(0) M (0)2 v(0)1/r(0) + 1 1 M (0) + 1 M (0) v(0)(1/r(0))−1v′(0) r(0) v(0)1/r(0) log v(0)−r′(0) r(0)2 . (7.4) Observing that and r(0) = 1 M (0) = 1 r′(0) = M ′(0) = 1 c β c v(0) = kfkL1(ρs) v′(0) = w(0) = −G Ef ρs dm + 1 c G f log f ρs dm 29 we see that (7.4) reads β 0 ≥ − + c kfkL1(ρs) −G c G f log f ρs dm − 1 Ef ρs dm 1 ckfkL1(ρs) log kfkL1(ρs) which after rearranging is precisely (sLSI). Remark 7.3. Theorem 7.2 does not rely on any properties of LSH func- tions, except the assumption that they satisfy (sHC). So more broadly, any appropriate class of functions satisfying (sHC) will also satisfy (sLSI). 7.3 sLSI implies sHC In this section, we show that if the strong logarithmic Sobolev inequality is satisfied for LSH functions, then so is strong hypercontractivity. We begin by noting that the semigroup e−tE is contractive on log- subharmonic functions. We assume some integrability on ∇f but this assumption will be removed later. Lemma 7.4. Suppose f ∈ LSH ∩ W 1,1+(ρs). Then for any t ≥ 0 we have ke−tE fkL1(ρs) ≤ kfkL1(ρs). Proof. Let T > 0 be arbitrary. Applying Lemma 6.4 with r(t) ≡ 1 = q, we have d dtke−tE fkL1(ρs) = −G Ee−tE f ρs dm. Now from Lemma 6.3, again with r(t) ≡ 1 = q, we have in particular that e−tE f,∇e−tEf, Ee−tEf ∈ L1(ρs). Also, f is subharmonic and hence so is e−tE f by (2.17). So by Lemma 2.21, we have G Ee−tE f ρs dm = s 2 G ∆e−tEf ρs dm ≥ 0. Hence ke−tEfkL1(ρs) is a decreasing function of t. The next step is to show that (sLSI) implies that (sHC) holds at time t = tJ . We take p = 1 and again assume, for now, sufficient integrability for ∇f. 30 Lemma 7.5. Suppose that (sLSI) holds. Let 1 ≤ q < ∞, and set tJ = tJ (1, q) = c log q M = M (1, q) = exp(β · (1 − q−1)). Suppose f ∈ LSH ∩ W 1,q+(ρs). Then ke−tJ EfkLq(ρs) ≤ MkfkL1(ρs). (7.5) Proof. Set r(t) = et/c, so that r(tJ ) = q, and let ft, v(t), w(t) be as in Lemma 6.4. The hypotheses of Lemma 6.4 are satisfied, with T = tJ , so we have v ∈ C 1([0, tJ ]) and v′(t) = w(t). On the other hand, for each 0 ≤ t ≤ tJ , we have ft ∈ LSH, by Propo- sition 3.3 items 3 and 5. Moreover, Lemma 6.3 implies ft ∈ W 1,1+(ρs). So (sLSI) applies to ft. In terms of v(t), w(t), this reads cw(t) ≤ v(t) log v(t) + βv(t) (7.6) where we note that r′(t) r(t) = 1 c . Since w(t) = v′(t), we may rewrite (7.6) as d dt log v(t) ≤ 1 c log v(t) + β c . (7.7) Define M (t) = M (1, r(t)) = exp(β · (1 − e−t/c)) α(t) = 1 M (t)ke−tEfkLr(t)(ρs) = M (t) 1 v(t)1/r(t) as in the proof of Theorem 7.2. Note that α(0) = kfkL1(ρs) and α(tJ ) = M −1ke−tJ EfkLq(ρs). Then we have log α(t) = e−t/c(cid:18)− 1 c c(cid:19) ≤ 0 log v(t) − log v(t) + d dt d dt β using (7.7). Hence α(t) is a decreasing function on [0, tJ ], so in particular α(tJ ) ≤ α(0), which is the desired statement. Theorem 7.6. In any stratified Lie group G, if (sLSI) holds, then (sHC) holds, with the same constants c, β. 31 Proof. Fix f ∈ LSH ∩ Lq(ρs) and t ≥ tJ (p, q). gives Suppose first that p = 1 and f ∈ LSH ∩ W 1,q+(ρs). Then Lemma 7.5 (7.8) Set τ = t − tJ (1, q), and let g = e−tJ Ef q ∈ LSH. We apply Lemma 6.3 with T = tJ and r(t) ≡ q, so that g = ftJ , to see that g ∈ W 1,1+(ρs). Applying Lemma 7.4 to g, we have ke−tJ EfkLq(ρs) ≤ M (1, q)kfkL1(ρs). ke−τ EgkL1(ρs) ≤ kgkL1(ρs) or in other words ke−tE fkq Lq(ρs) ≤ ke−tJ EfkLq(ρs). (7.9) Combining (7.8) and (7.9) gives (sHC) in this case. Next, suppose only that p = 1, f ∈ LSH ∩ Lq+(ρs), but make no as- sumptions about ∇f . Let ϕn be a sequence of standard mollifiers as in Lemma 5.3, and set fn = ϕn ∗ f , so that fn → f pointwise and in L1+(ρs); then e−tEfn → e−tE f pointwise as well. We have fn ∈ LSH by Lemma 3.4; fn ∈ Lq+(ρs) by Lemma 5.1; and ∇fn ∈ Lq+(ρs) by Lemma 5.2. So by the previous case, we have ke−tE fnkLq(ρs) ≤ M (1, q)kfnkL1(ρs) (7.10) and by Fatou's lemma, the same holds for f . Next, suppose p = 1 and f ∈ LSH ∩ Lq(ρs). Then for any 0 < α < 1, we have f α ∈ LSH ∩ Lq+(ρs), so that by the previous case, ke−tE f αkLq(ρs) ≤ M (1, q)kf αkL1(ρs). Letting α → 1, we have f α → f pointwise and in L1(ρs) (by dominated convergence, using for instance 1 + f as the dominating function). So by Fatou's lemma, the result holds for f . Finally, let 0 < p ≤ q be arbitrary and f ∈ LSH ∩ Lq(ρs). Set g = f p and r = q/p. Then we have g ∈ LSH ∩ Lr(ρs), and so by the previous case we have ke−tEgkLr (ρs) ≤ M (1, r)kgkL1(ρs), t ≥ tJ (1, r). Noting that M (1, r) = M (p, q)p and tJ (1, r) = tJ (p, q), this reads ke−tE fkp which is equivalent to (sLSI). Lq(ρs) ≤ M (p, q)pkfkp Lp(ρs) 32 8 Weaker notions of subharmonicity We have chosen to focus our attention in this paper on log-subharmonic functions which are C 2. In this section, we note that our results for strong hypercontractivity can be extended to functions which are LSH in a weaker sense. A comprehensive discussion of the various possible definitions of sub- harmonicity on stratified Lie groups is beyond the scope of this paper. We refer the reader to [4], in which the basic definition of subharmonic functions (Definition 7.2.2) is in terms of harmonic measure. Many other equivalent characterizations are given; perhaps the simplest is the following definition in terms of distributional derivatives. Definition 8.1. We say a function f : G → [−∞,∞) is weakly subhar- monic if f ∈ L1 loc(G, m) and ∆f ≥ 0 in the sense of distributions. We say a function f : G → [0,∞) is weakly log-subharmonic if either f ≡ 0 or log f is weakly subharmonic, and we write f ∈ wLSH. Strictly speaking, a function f is weakly subharmonic in this sense iff it has an m-version which is subharmonic in the sense of [4, Definition 7.2.2]; see [4, Theorem 8.2.15 and Corollary 8.2.4]. The distinction is irrelevant for our current purposes, since null sets will not concern us. Let us also call attention to [4, Corollary 8.2.3], where it is shown that a function is (weakly) subharmonic iff it satisfies a sub-averaging property, which is analogous to the definition of subharmonic used in [14, 15]. Lemma 8.2. Suppose f ∈ wLSH ∩ Lq+(ρs), where q ≥ 1. Then there is a sequence fn ∈ LSH ∩ Lq+(ρs) with fn → f almost everywhere and in L1+(ρs). Proof. If f ≡ 0 this is trivial by taking fn = 1/n. Otherwise, log f is weakly subharmonic. Let ϕn ∈ C ∞ c be a sequence of nonnegative standard mollifiers with ´G ϕn dm = 1, as in Lemma 5.3, and set gn = ϕn ∗ log f . Then gn → log f almost everywhere. By [4, Theorem 8.1.5 and Corollary 8.2.3], gn is also weakly subharmonic; moreover, since gn ∈ C ∞(G), we have by [4, Proposition 7.2.5] that ∆gn ≥ 0. Set fn = exp(gn), so that fn ∈ LSH and fn → f almost everywhere. Now fn ≤ ϕn ∗ f by Jensen's inequality. By Lemma 5.1, we have ϕn ∗ f ∈ Lq+(ρs), so the same is true for fn. And by Lemma 5.3, we have ϕn ∗ f → f in L1+(ρs), so fn → f in L1+(ρs) as well. Theorem 8.3. If (sHC) holds for all f ∈ LSH ∩ Lq(ρs), then it holds for all f ∈ wLSH ∩ Lq(ρs). 33 Proof. As in the proof of Theorem 7.6, it suffices to prove (sHC) with p = 1 and for all f ∈ wLSH ∩ Lq+(ρs). Using Lemma 8.2, choose fn ∈ LSH ∩ Lq+(ρs) with fn → f almost everywhere and in L1+(ρs). Then (sHC) holds for each fn. We have kfnkL1(ρs) → kfkL1(ρs), so by Fatou's lemma, (sHC) holds for f . In the setting of complex stratified Lie groups (Example 4.4), the mod- ulus of a holomorphic function is weakly LSH. Proposition 8.4. Let G be a complex stratified Lie group, and suppose f : G → C is holomorphic. Then f ∈ wLSH. Proof. Let fn =qf2 + 1 n . We showed in Lemma 4.5 that fn ∈ LSH. Now fn ↓ f, and so log f is a decreasing limit of subharmonic functions. By [4, Theorem 8.2.7], log f is therefore weakly subharmonic. 9 Acknowledgments I would like to thank Bruce K. Driver for suggesting this problem and draw- ing attention to the papers [14, 15] whose results are extended here. I would also like to thank Leonard Gross and Laurent Saloff-Coste for their collabo- ration on the paper [11], of which this paper is an outgrowth. This research was supported by a grant from the Simons Foundation (#355659, Nathaniel Eldredge). References [1] D. Bakry. On Sobolev and logarithmic Sobolev inequalities for Markov semigroups. In New trends in stochastic analysis (Charingworth, 1994), pages 43 -- 75. World Sci. Publ., River Edge, NJ, 1997. [2] Dominique Bakry and Michel ´Emery. Hypercontractivit´e de semi- groupes de diffusion. C. R. Acad. Sci. Paris S´er. I Math., 299(15): 775 -- 778, 1984. ISSN 0249-6291. [3] Dominique Bakry, Fabrice Baudoin, Michel Bonnefont, Djalil Chafaı. the Heisenberg 2008. http://dx.doi.org/10.1016/j.jfa.2008.09.002. On gradient bounds group. ISSN 0022-1236. J. Funct. Anal., doi: 10.1016/j.jfa.2008.09.002. for and the heat kernel on 255(8):1905 -- 1938, URL 34 [4] A. Bonfiglioli, E. Lanconelli, and F. Uguzzoni. Stratified Lie groups and potential theory for their sub-Laplacians. Springer Monographs in Mathematics. Springer, Berlin, 2007. ISBN 978-3-540-71896-3; 3-540- 71896-6. [5] Eric A. Carlen. Some integral identities and inequali- coher- 1991. URL entire for state ties ent ISSN 0022-1236. http://dx.doi.org/10.1016/0022-1236(91)90022-W. functions and their application to the 97(1):231 -- 249, 10.1016/0022-1236(91)90022-W. J. Funct. Anal., transform. doi: [6] Bruce K. Driver and Leonard Gross. Hilbert spaces of holomorphic functions on complex Lie groups. In New trends in stochastic analysis (Charingworth, 1994), pages 76 -- 106. World Sci. Publ., River Edge, NJ, 1997. [7] Bruce K. Driver and Tai Melcher. Hypoelliptic heat kernel in- J. Funct. Anal., 221(2):340 -- ISSN 0022-1236. doi: 10.1016/j.jfa.2004.06.012. URL equalities on the Heisenberg group. 365, 2005. http://dx.doi.org/10.1016/j.jfa.2004.06.012. [8] Nathaniel Eldredge. Precise estimates for the subelliptic heat J. Math. Pures Appl. (9), 92(1): 10.1016/j.matpur.2009.04. URL http://dx.doi.org/10.1016/j.matpur.2009.04.011. ISSN 0021-7824. kernel on H-type groups. 52 -- 85, 2009. 011. arXiv:0810.3218. doi: [9] Nathaniel Eldredge. Gradient estimates for the subelliptic heat 258(2):504 -- 533, kernel on H-type groups. 2010. URL http://dx.doi.org/10.1016/j.jfa.2009.08.012. arXiv:0904.1781. 10.1016/j.jfa.2009.08.012. J. Funct. Anal., ISSN 0022-1236. doi: [10] Nathaniel Eldredge. On complex H-type Lie algebras. Preprint. arXiv:1406.2396, 2014. URL http://arxiv.org/abs/1406.2396. [11] Nathaniel Eldredge, Leonard Gross, and Laurent Saloff-Coste. Strong hypercontractivity and logarithmic Sobolev inequalities on strat- the ified complex Lie groups. American Mathematical Society. arXiv:1510.05151, 2015. URL http://arxiv.org/abs/1510.05151. in Transactions of To appear 35 [12] Paul Federbush. Partially alternate derivation of a result of Nelson. Journal of Mathematical Physics, 10(1):50 -- 52, 1969. doi: 10.1063/1. 1664760. URL http://dx.doi.org/10.1063/1.1664760. [13] James Glimm. Boson fields with nonlinear self-interaction in two di- mensions. Comm. Math. Phys., 8:12 -- 25, 1968. ISSN 0010-3616. [14] Piotr Graczyk, Todd Kemp, and Jean-Jacques Loeb. Hypercon- tractivity for log-subharmonic functions. J. Funct. Anal., 258(6):1785 -- 1805, 2010. ISSN 0022-1236. doi: 10.1016/j.jfa.2009.08.014. URL http://dx.doi.org/10.1016/j.jfa.2009.08.014. [15] Piotr Graczyk, Todd Kemp, and Jean-Jacques Loeb. Strong logarithmic Sobolev inequalities for log-subharmonic functions. Canad. J. Math., 67 (6):1384 -- 1410, 2015. ISSN 0008-414X. doi: 10.4153/CJM-2015-015-8. URL http://dx.doi.org/10.4153/CJM-2015-015-8. [16] Leonard Gross. Logarithmic Sobolev inequalities. Amer. J. Math., 97 ISSN 0002-9327. doi: 10.2307/2373688. URL (4):1061 -- 1083, 1975. http://dx.doi.org/10.2307/2373688. [17] Leonard Gross. Hypercontractivity over complex manifolds. Acta 10.1007/ Math., 182(2):159 -- 206, 1999. BF02392573. URL http://dx.doi.org/10.1007/BF02392573. ISSN 0001-5962. doi: [18] Leonard Gross. Strong hypercontractivity and relative subharmonicity. J. Funct. Anal., 190(1):38 -- 92, 2002. ISSN 0022-1236. doi: 10.1006/ jfan.2001.3883. URL http://dx.doi.org/10.1006/jfan.2001.3883. Special issue dedicated to the memory of I. E. Segal. [19] Leonard Gross. Hypercontractivity, logarithmic Sobolev inequalities, and applications: a survey of surveys. In Diffusion, quantum theory, and radically elementary mathematics, volume 47 of Math. Notes, pages 45 -- 73. Princeton Univ. Press, Princeton, NJ, 2006. [20] Lars Hormander. Hypoelliptic second order differential equations. Acta Math., 119:147 -- 171, 1967. ISSN 0001-5962. doi: 10.1007/BF02392081. URL http://dx.doi.org/10.1007/BF02392081. [21] Jun-Qi Hu and Hong-Quan Li. Gradient estimates for the heat Potential Anal., 33(4):355 -- 386, 10.1007/s11118-010-9173-1. URL semigroup on H-type groups. 2010. doi: http://dx.doi.org/10.1007/s11118-010-9173-1. ISSN 0926-2601. 36 [22] Svante Janson. On hypercontractivity for multipliers on orthogonal polynomials. Ark. Mat., 21(1):97 -- 110, 1983. ISSN 0004-2080. doi: 10. 1007/BF02384302. URL http://dx.doi.org/10.1007/BF02384302. [23] Svante Janson. On complex hypercontractivity. J. Funct. Anal., 151 (1):270 -- 280, 1997. ISSN 0022-1236. doi: 10.1006/jfan.1997.3144. URL http://dx.doi.org/10.1006/jfan.1997.3144. [24] Aroldo Kaplan. Fundamental solutions for a class of hypoelliptic PDE generated by composition of quadratic forms. Trans. Amer. Math. Soc., 258(1):147 -- 153, 1980. ISSN 0002-9947. doi: 10.2307/1998286. URL http://dx.doi.org/10.2307/1998286. [25] Hong-Quan Li. Estimation optimale du gradient du semi-groupe de la chaleur sur le groupe de Heisenberg. J. Funct. Anal., 236(2):369 -- 394, 2006. ISSN 0022-1236. doi: 10.1016/j.jfa.2006.02.016. URL http://dx.doi.org/10.1016/j.jfa.2006.02.016. [26] Hong-Quan Li. Estimations asymptotiques du noyau de la chaleur sur les groupes de Heisenberg. C. R. Math. Acad. Sci. Paris, 344(8):497 -- 502, 2007. ISSN 1631-073X. doi: 10.1016/j.crma.2007.02.015. URL http://dx.doi.org/10.1016/j.crma.2007.02.015. [27] Hong-Quan Li. Estimations optimales du noyau de la chaleur sur les groupes de type Heisenberg. J. Reine Angew. Math., 646:195 -- 233, 2010. ISSN 0075-4102. doi: 10.1515/CRELLE.2010.070. URL http://dx.doi.org/10.1515/CRELLE.2010.070. [28] Fran¸coise Lust-Piquard. Ornstein-Uhlenbeck stratified groups. on 2010. http://dx.doi.org/10.1016/j.jfa.2009.11.012. J. Funct. Anal., doi: ISSN 0022-1236. semi-groups 258(6):1883 -- 1908, URL 10.1016/j.jfa.2009.11.012. [29] Richard Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002. ISBN 0-8218-1391-9. [30] Alexander Nagel, Elias M. Stein, and Stephen Wainger. Balls and metrics defined by vector fields. I. Basic properties. Acta Math., 155 (1-2):103 -- 147, 1985. ISSN 0001-5962. doi: 10.1007/BF02392539. URL http://dx.doi.org/10.1007/BF02392539. 37 [31] Edward Nelson. A quartic interaction in two dimensions. In Mathe- matical Theory of Elementary Particles (Proc. Conf., Dedham, Mass., 1965), pages 69 -- 73. M.I.T. Press, Cambridge, Mass., 1966. [32] Edward Nelson. The free Markoff field. J. Functional Analysis, 12: 211 -- 227, 1973. [33] Ludovic Rifford. Sub-Riemannian geometry and optimal transport. SpringerBriefs in Mathematics. Springer, Cham, 2014. ISBN 978-3- 319-04803-1; 978-3-319-04804-8. doi: 10.1007/978-3-319-04804-8. URL http://dx.doi.org/10.1007/978-3-319-04804-8. [34] Derek W. Robinson. Elliptic operators and Lie groups. Oxford Math- ematical Monographs. The Clarendon Press Oxford University Press, New York, 1991. ISBN 0-19-853591-0. Oxford Science Publications. [35] A. J. Stam. Some inequalities satisfied by the quantities of information Information and Control, 2:101 -- 112, 1959. of Fisher and Shannon. ISSN 0890-5401. [36] Robert S. Strichartz. Sub-Riemannian geometry. ential Geom., 24(2):221 -- 263, http://projecteuclid.org/euclid.jdg/1214440436. ISSN 0022-040X. 1986. J. Differ- URL [37] Robert S. Strichartz. Corrections to: "Sub-Riemannian geometry" [J. Differential Geom. 24 (1986), no. 2, 221 -- 263; MR0862049 (88b:53055)]. J. Differential Geom., 30(2):595 -- 596, 1989. ISSN 0022-040X. URL http://projecteuclid.org/euclid.jdg/1214443604. [38] N. Th. Varopoulos. Small time Gaussian estimates of heat diffusion kernels. II. The theory of large deviations. J. Funct. Anal., 93(1):1 -- 33, 1990. ISSN 0022-1236. doi: 10.1016/0022-1236(90)90136-9. URL http://dx.doi.org/10.1016/0022-1236(90)90136-9. [39] N. Th. Varopoulos, L. Saloff-Coste, and T. Coulhon. Analysis and geometry on groups, volume 100 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1992. ISBN 0-521-35382-3. [40] Zheng-Fang Zhou. The contractivity of the free Hamiltonian semi- group in the Lp space of entire functions. J. Funct. Anal., 96(2):407 -- 425, 1991. ISSN 0022-1236. doi: 10.1016/0022-1236(91)90067-F. URL http://dx.doi.org/10.1016/0022-1236(91)90067-F. 38
1601.03183
3
1601
2016-02-12T13:16:06
The essential spectrum of the Neumann--Poincare operator on a domain with corners
[ "math.FA", "math.SP" ]
Exploiting the homogeneous structure of a wedge in the complex plane, we compute the spectrum of the anti-linear Ahlfors-Beurling transform acting on the associated Bergman space. Consequently, the similarity equivalence between the Ahlfors-Beurling transform and the Neumann-Poincare operator provides the spectrum of the latter integral operator on a wedge. A localization technique and conformal mapping lead to the first complete description of the essential spectrum of the Neumann-Poincare operator on a planar domain with corners, with respect to the energy norm of the associated harmonic field.
math.FA
math
THE ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR ON A DOMAIN WITH CORNERS KARL-MIKAEL PERFEKT AND MIHAI PUTINAR Abstract. Exploiting the homogeneous structure of a wedge in the complex plane, we compute the spectrum of the anti-linear Ahlfors -- Beurling transform acting on the associated Bergman space. Conse- quently, the similarity equivalence between the Ahlfors -- Beurling trans- form and the Neumann -- Poincar´e operator provides the spectrum of the latter integral operator on a wedge. A localization technique and con- formal mapping lead to the first complete description of the essential spectrum of the Neumann -- Poincar´e operator on a planar domain with corners, with respect to the energy norm of the associated harmonic field. 1. Introduction Exactly a hundred years ago Torsten Carleman defended his doctoral dissertation titled " Uber das Neumann -- Poincar´esche Problem fur ein Ge- biet mit Ecken" [8]. The double-layer potential singular integral operator associated with a domain Ω ⊂ R2, known also as the Neumann -- Poincar´e (NP) operator, was at that time a central object of study, first for its role in solving boundary value problems of mathematical physics, but also as the main example in the emerging abstract spectral theories proposed by Hilbert, Fredholm and F. Riesz. While the NP operator is compact on smooth boundaries, the presence of corners produces continua in its essen- tial spectrum. For the modern reader these concepts make no sense without a well defined, complete functional space where the operator NP acts, not to mention also the current definitions of essential spectrum, spectral reso- lution, approximate or generalized eigenvalues, etc. In a tour de force Car- leman did solve the singular integral equation governed by the NP operator and analyzed the (asymptotic) structure of its solutions in a domain with corners. He made use of elementary and very ingenious geometric transfor- mations together with the, at his time new, theory of Fredholm determinants combined with the canonical factorization of entire functions of Hadamard. Carleman's work did not attract the visibility it deserves, nor did the prior results of his predecessors, among which we mention Zaremba [25]. Date: August 27, 2018. Key words and phrases. Neumann -- Poincar´e operator, energy norm, Bergman space, essential spectrum. 1 2 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR Only a few years after Carleman's defense, Radon [21] developed the theory of measures of bounded variation, and applied it to study the NP operator on the space of continuous functions C(∂Ω). He computed the essential spectral radius for boundaries ∂Ω of bounded rotation, extending Carleman's work. Note that we now understand that for non-smooth bound- aries, the spectrum of the NP operator depends drastically on the underlying space. For instance, when the NP operator is considered on Lp(∂Ω), p ≥ 2, Ω a curvilinear polygon, the complete spectral picture is available [16] -- and it is entirely different from what appears in the work of Carleman and Radon. The Hilbert space on which we will perform a spectral analysis is the energy space of potential fields with sources carried by ∂Ω. The energy space was advocated by Poincar´e in his foundational and novel approach to the Dirichlet problem. It stands out as a natural setting for the NP operator for at least two reasons. First, the invertibility properties of the NP operator acting on the energy space lead to finite energy solutions of boundary value problems for the Laplacian, and such solutions often carry a physical interpretation. Second, due to a symmetrization property, the NP operator has real spectrum on the energy space even when the boundary ∂Ω is non-smooth (this is not true for example on L2(∂Ω)). The recent survey [23] treats among other things qualitative aspects of the essential spectrum of the NP operator on various spaces of interest, for domains with corners. However, the case of the physically motivated energy space is noted for the lack of information concerning the structure of the essential spectrum. Ahlfors [2] observed a connection between the spectral radius of the NP operator (the largest Fredholm eigenvalue of Ω) and the quasiconformal reflection coefficient of ∂Ω. The reflection coefficient is notoriously difficult to compute for general domains which do not have any special geometric structure [14]. Yet, Ahlfors' inequality provides to date nearly all known spectral bounds of the NP operator in the energy norm. Very recently, the NP operator has received a resurgence of interest aris- ing from the mathematical theory of new materials and its need to solve various inverse problems. In particular, spectral analysis questions of the NP operator on non-smooth domains are central in the penetrating works of Ammari, Kang, Milton and their enthusiastic collaborators [3, 4, 5, 6]. And again, a century after Carleman's work, estimates of the location of the essential spectrum of the NP operator and the asymptotic behavior of its generalized eigenfunctions turn out to be highly sought results. The preprint [12] contains a detailed description of the spectral resolution of the NP operator acting on a lens domain, with respect to the energy space. In a previous work [19] we have obtained bounds for the spectrum of the NP operator, in the same energy space, on domains with corners, via distortion theorems of conformal mappings. In the same setting, a detailed numerical study of the spectrum has been done in the preprint [11]. Some interesting ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 3 geometric analysis questions pertaining to the NP operator also appear in [17]. The present note contains a new approach to the spectral analysis of the NP operator in a wedge in two variables. We exploit the similarity between the NP operator acting on the energy space (identifiable with a fractional Sobolev space on the boundary) and the Ahlfors -- Beurling singular integral operator [7] acting on the Bergman space of the underlying domain. The homothetic action of the commutative group of positive real numbers on the wedge turns out to simplify, at least conceptually, the computation. We then generalize, via a standard localization procedure and a conformal mapping technique, the wedge computation to domains with finitely many corners. The outcome is an exact picture of the essential spectrum of the NP operator, in the energy norm. A few comments on the nature of the singular integral transformations we deal with are in order: the NP operator is not symmetric, but only symmetrizable in a norm which is equivalent to a fractional Sobolev space norm on the boundary, see [19] for details. Therefore the NP operator is only scalar in the sense of Dunford, and any spectral resolution has to be understood in this generalized sense [9]. The spectral analysis of the anti-linear Ahlfors -- Beurling operator can today be naturally understood within the abstract theory of complex symmetric operators [10]. While our localization and conformal mapping argument shows that the essential spectrum is a continuum, it does not control the possible singular continuous part of the spectrum. Recall that the absolute continuity of the spectrum of an operator can be altered by a Hilbert-Schmidt perturbation (according to the classical Weyl-von Neumann theorem). Conversely, for a self-adjoint operator it is preserved by a trace class perturbation (according to the Kato- Rosenblum theorem). Let us therefore clarify that our localizations, while Dunford scalar, can not be jointly put on a normal form. On the other hand, it is known [24] that as a rectangle is elongated, the spectral radius of the associated Ahlfors -- Beurling operator changes. When combined with the results on the essential spectrum of the present article, it follows that isolated eigenvalues of the NP operator depend on the non-local geometry of the domain. The eigenvalues are thus very unlikely to be given a simple description. 2. Preliminaries Let Ω ⊂ C be a bounded Lipschitz domain with connected boundary. The Sobolev space of order 1/2 along the boundary, H 1/2(∂Ω), is defined in the usual way, using a bi-Lipschitz atlas to view ∂Ω as a manifold. A Hilbert space norm on H 1/2(∂Ω) is given by the Besov norm (1) kf k2 H 1/2(∂Ω) ∼ kf k2 L2(∂Ω) +Z∂Ω×∂Ω f (x) − f (y)2 x − y2 dσ(x) dσ(y), 4 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR where σ is the natural Hausdorff measure on ∂Ω. See for instance [22, Appendix II]. H −1/2(∂Ω) is then defined by duality with respect to the pairing of L2(∂Ω), and H −1/2 (∂Ω) is its subspace of elements f ∈ H −1/2(∂Ω) 0 such that hf, 1iL2(∂Ω) = 0. The Neumann -- Poincare operator K : H 1/2 (∂Ω) → H 1/2 (∂Ω) is defined by (2) Kf (x) = 2 π p. v.Z∂Ω ∂ny log x − yf (y) dσ(y), x ∈ ∂Ω, where ny is the outward normal derivative of ∂Ω at y. K is always a bounded operator. When evaluating the integral for x /∈ ∂Ω, we obtain the harmonic double-layer potential Df , Df (x) = ∂ny log x − yf (y) dσ(y), x ∈ C \ ∂Ω = Ω ∪ Ω c . 1 πZ∂Ω Yet another characterization of H 1/2(∂Ω) is that it consists precisely of the functions f such that Df ∈ H 1(C \ ∂Ω), i.e. such that kDf kH 1(C\∂Ω) =ZΩ∪Ω ∇Df 2 dx < ∞. c In other words, H 1/2(∂Ω) is the space of charges which yield potentials of finite energy. As an element of H 1(Ω), Df has an (interior) trace Tr Df ∈ H 1/2(∂Ω). Df and Kf are related by the jump formula (3) Tr Df = 1 2 (f + Kf ). In the case that ∂Ω is a C 2-curve, [13, Ch. 8] offers a very readable and self- contained introduction to the Neumann -- Poincar´e operator and its use in constructing finite energy solutions to the Dirichlet and Neumann problems for the Laplacian. By K ∗ : H −1/2(∂Ω) → H −1/2(∂Ω) we mean the adjoint of K with respect to the L2 (∂Ω)-pairing. In [19] the authors showed, for a general Lipschitz domain, that K ∗ : H −1/2 (∂Ω) is similar to a self adjoint operator. The only effect of considering the action of K ∗ on H −1/2 (∂Ω) rather than on H −1/2(∂Ω) is that it loses its isolated eigenvalue λ = 1 of multiplicity 1. (∂Ω) → H −1/2 0 0 0 One realization of such a self adjoint operator is the anti-linear Ahlfors -- Beurling operator TΩ acting on the Bergman space L2 a(Ω), (4) TΩf (z) = 1 π f (ζ) (ζ − z)2 dA(ζ), f ∈ L2 a(Ω), z ∈ Ω. p. v.ZΩ Here the Bergman space L2 functions in Ω. To be precise, K ∗ : H −1/2 L2 a(Ω) are similar as R-linear operators. a(Ω) consists of the holomorphic square-integrable (∂Ω) and TΩ : (∂Ω) → H −1/2 a(Ω) → L2 0 0 ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 5 From the similarity we have the following equality of spectra: σ(K) = σR(TΩ) ∪ {1}. Note that if λ is in the spectrum of TΩ, then by anti-linearity so is eiθλ for any θ ∈ R. However, we are interested only in the real spectrum, K : H 1/2(∂Ω)/C → H 1/2(∂Ω)/C being similar to a self-adjoint operator over the complex field. Therefore, to determine the spectrum of K using TΩ we consider only λ ≥ 0 in the spectrum of TΩ, and note that the spectrum of K consists of ±λ, for all such points λ, in addition to the simple eigenvalue 1. The Neumann -- Poincar´e operator (2) may be written more explicitly as Kf (x) = 2 π p. v.Z∂Ω hy − x, nyi y − x2 f (y) dσ(y), x ∈ ∂Ω, In this article we shall consider the case where Ω ⊂ C is a C 2-smooth curvilinear polygon. By this we mean that Ω is a bounded and simply connected domain whose boundary is curvilinear polygonal: there are a finite number of counter-clockwise consecutive vertices (aj)N j=1 ⊂ C, 1 ≤ N < ∞, connected by C 2-smooth arcs γj : [0, 1] → C with starting point aj and end point aj+1 (indices modulo N ), such that ∂Ω = ∪jγj and γj and γj+1 meet at the interior angle αj+1 at aj+1, 0 < αj+1 < 2π. Note that if γ ⊂ ∂Ω is a C 2 subarc, then the kernel (5) k(x, y) = hy − x, nyi y − x2 is actually bounded and continuous for y ∈ γ′, x ∈ ∂Ω, where γ′ is any strict subarc of γ. Similarly, k(x, y) is bounded and continuous for x ∈ γ1 and y ∈ γ2 if γ1 and γ2 are compact and disjoint subsets of ∂Ω (regardless of smoothness). From these observations we obtain the compactness of certain cut-offs which we shall later use to localize the operator K. We roughly follow the proof from [13] that K is compact if ∂Ω is C 2, in addition to a trivial observation about multiplication operators. Lemma 1. Let ρ be a smooth function on ∂Ω. Then Mρ, the operator of multiplication by ρ, is a bounded operator on H 1/2(∂Ω). Proof. This is a straightforward consequence of the Besov norm expression (1) for H 1/2(∂Ω). (cid:3) Lemma 2. Let Ω be a C 2-smooth curvilinear polygon with vertices (aj)N j=1. For each j, 1 ≤ j ≤ N , let ρj be a smooth function on ∂Ω such that ρj(x) = 1 for all x in a neighborhood of aj. Furthermore, suppose that the supports of ρj are pairwise disjoint (at a positive distance apart). Let ρN +1 = 1 − j=1 ρj. If j 6= k or j = k = N + 1, then Mρj KMρk : H 1/2(∂Ω) → H 1/2(∂Ω) is a PN compact operator. 6 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR Proof. Let d dσ denote (tangential) differentiation along ∂Ω, extended in the distributional sense to an operator d dσ : H 1/2(∂Ω) → H −1/2(∂Ω). We will actually show that Mρj KMρk is bounded as a map from H 1/2(∂Ω) into H 1(∂Ω), where H 1(∂Ω) is the space of functions f such that d dσ f ∈ L2(∂Ω). Since the embedding H 1(∂Ω) ֒→ H 1/2(∂Ω) is compact this is sufficient. For f ∈ H 1/2(∂Ω), let Df ∈ H 1(Ω) denote the double-layer potential of f in the interior of Ω. The tangential derivative is associated with the classical jump formula [15] d dσ Tr Df = 1 2(cid:18) d dσ f − K ∗ d dσ f(cid:19) . The validity of this formula for all f ∈ H 1/2(∂Ω) follows from the classical considerations by approximation and the continuity of all operators involved. Combined with the jump formula (3) we conclude that d dσ Kf (x) = −K ∗ d dσ f (x) = − = 2 π d dσ(y) 2 π p. v.Z∂Ω ∂nx log x − y d dσ f (y) dσ(y) ∂nx log x − y(f (y) − f (x)) dσ(y), where the last step equality follows from integration by parts. Note that if x, y ∈ supp ρN +1, then If instead x ∈ supp ρj and y ∈ supp ρk, k 6= j, then bounded. Hence, p. v.Z∂Ω (cid:12)(cid:12)(cid:12)(cid:12) d dσ(y) (6) d dσ Mρj KMρk f (x) −(cid:20) d (cid:12)(cid:12)(cid:12)(cid:12) It is clear that the term(cid:2) d Z∂Ω(cid:12)(cid:12)(cid:12)(cid:12)ρj(x)Z∂Ω .Z∂ΩZ∂Ω . d 1 x − y dσ(y) ∂nx log x − y is ∂nx log x − y(cid:12)(cid:12)(cid:12)(cid:12) . ρj(x)(cid:21) K(ρkf )(x)(cid:12)(cid:12)(cid:12)(cid:12) . ρj(x)Z∂Ω dσ ρj(x)(cid:3) K(ρkf )(x) is in L2(∂Ω), by Lemma 1 and ρk(x)f (x) − ρk(y)f (y) x − y dσ(y) dσ the boundedness of K on H 1/2(∂Ω). It remains to show that the right-hand side of (6) is in L2(∂Ω). However, this follows from the Cauchy-Schwarz inequality and Lemma 1: ρk(x)f (x) − ρk(y)f (y) 2 dσ(y)(cid:12)(cid:12)(cid:12)(cid:12) dσ(x) dσ(y) dσ(x) x − y ρk(x)f (x) − ρk(y)f (y)2 x − y2 . kMρk f k2 H 1/2(∂Ω) . kf k2 H 1/2(∂Ω). (cid:3) ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 7 3. The Wedge Let Wα = {z ∈ C : arg z < α/2} be a wedge of aperture α, 0 < α < 2π. Consider any linear fractional transformation L which maps Wα ∪ {∞} onto a bounded domain. Since TΩ is unitarily equivalent to TL(Ω) for any domain Ω [19], it follows that the spectrum of the Neumann -- Poincar´e operator K associated with L(Wα) may be determined by considering TWα.1 This section is devoted to proving the following. Theorem 3. The spectrum of K : H 1/2 (∂L(Wα)) → H 1/2 (∂L(Wα)) is given by σ(K) =nx ∈ R : x ≤(cid:12)(cid:12)(cid:12)1 − α π(cid:12)(cid:12)(cid:12)o ∪ {1}. 1 is a a simple eigenvalue. The remainder of the spectrum is essential, of uniform multiplicity 2. Remark. This result, stated somewhat differently, also appears in the preprint [12]. Our proof is rather different and we include it with full details below. It's sufficient to consider α < π, because the spectrum of TWα is the same . We begin with the as that of TW2π−α, since W2π−α is a rotation of Wα following simple proposition about the kernels of L2(Wα). Proposition 4 ([20]). L2(Wα) is a reproducing kernel Hilbert space. The reproducing kernel at the point z ∈ Wα is given by wπ/α−1zπ/α−1 (wπ/α + zπ/α)2 kz(w) = 1 α2 (7) . c Now let R± = {z ∈ C : arg z = ±α/2} be the two rays of the boundary ∂Wα. We consider the holomorphic Schwarz functions S± on Wα such that S±(ζ) = ¯ζ on R±, S±(ζ) = e∓iαζ. By first applying Stokes' theorem and then Cauchy-Goursat's theorem to each of the rays R+ and R−, we find for functions f with sufficient decay that for z ∈ Wα TWαf (z) = lim ε→0 f (ζ) ¯ζ − ¯z dζ − 1 ε2Zζ−z=ε f (ζ) f (ζ)(ζ − z) dζ# dζ(cid:21) i i 2π"Z∂Wα 2π(cid:20)ZR− 2πZ ∞ π Z ∞ f (y)(cid:20) 0 sin α i 0 f (y) = = = dζ −ZR+ f (ζ) S−(ζ) − ¯z 1 eiαy − ¯z S+(ζ) − ¯z 1 e−iαy − ¯z(cid:21) dy − y (eiαy − ¯z)(e−iαy − ¯z) dy. 1We avoid considering K directly on ∂Wα due to the technical difficulties associated with unbounded domains. 8 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR For instance, this formula is valid for linear combinations of kernels (7). Hence, it follows for x > 0 that (8) TWαf (x) = = sin α 0 π Z ∞ π Z ∞ −∞ sin α f (y) y eiαy − x2 dy etf (etx) et eiαet − 12 dt, motivating the following lemma. Lemma 5. For t ∈ R, let Ut : L2 a(Wα) → L2 a(Wα) be the operator Utf (z) = etf (etz). Then (Ut)t∈R is a strongly continuous group of unitary operators with gen- erator iA, Ut = eitA, where Af (z) = −i(f (z) + zf ′(z)). A is a (densely defined) self-adjoint operator with full spectrum, σ(A) = R. Furthermore, its spectrum has uniform multiplicity 1. Proof. The verification that Ut is a strongly continuous group of unitaries is straightforward. The formula for A follows immediately from the fact that iA is the strong limit of t−1(Ut − I) as t → 0. For t real, zit is bounded from below and above in Wα. Hence the operator Mt of multiplication by zit is bounded and invertible on L2 a(Wα). A computation shows that M −1 t AMt = A + tI. Hence A is similar to A + t for every t ∈ R. Since the spectrum of A is not empty, it must therefore be full. To show that the spectrum of A has multiplicity 1, we prove that for every z ∈ Wα, the reproducing kernel kz of L2(Wα) at z, see (7), is a cyclic vector of A. Suppose that f ∈ L2(Wα) is orthogonal to span{Ankz : n ≥ 0}. Clearly every function gn(w) = wnk(n) (w) is in this linear span. Note that z dρ(cid:19)n gn(w) =(cid:18) d , kz(ρw)(cid:12)(cid:12)(cid:12)ρ=1 1 ρ2 kz/ρ(w). and that for ρ > 0 kz(ρw) = Hence we have that 0 = hf, gniL2(Wα) =(cid:18) d dρ(cid:19)n ρ−2f (z/ρ)(cid:12)(cid:12)(cid:12)ρ=1 . Evaluating for n = 0, 1, 2, . . . we find that f (n)(z) = 0 for every n ≥ 0, proving that f = 0. (cid:3) ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 9 Let J : L2 a(Wα) → L2 a(Wα) be the (anti-linear) conjugation given by Jf (z) = f (¯z). Then TWαJ is a self-adjoint operator on L2 a(Wα) which additionally is J-symmetric [10] in the sense that TWαJ = J(TWαJ)J = JTWα. Lemma 6. For 0 < α < π, TWαJ is a positive operator, with spectrum σ(TWαJ) =nx ∈ R : 0 ≤ x ≤ 1 − α πo of uniform multiplicity 2. Proof. For z with ℑz < 1, let sin α F (z) = eitz et eiαet − 12 dt. Then F (A) is the operator on L2 a(Wα) such that F (A)f (z) = = eitAf (z) etf (etz) et eiαet − 12 dt et eiαet − 12 dt. −∞ π Z ∞ π Z ∞ π Z ∞ −∞ −∞ sin α sin α Comparing with (8) it is now clear that F (A) = TWαJ. We already know that F (A) = TWαJ is a self-adjoint operator. The change of variable s = eit gives us that F (x) =Z ∞ 0 six s s2 − 2s cos α + 1 ds s , x ∈ R. This Mellin transform can be computed by "partial fractions" (see also [16], pp. 453), which yields F (x) = sin(i(π − α)x) sin(iπx) , x ∈ R. F is thus a smooth even positive function on R, such that F (0) = 1 − α/π, F is decreasing for x ≥ 0 and F (x) → 0 as x → ∞. The statement of the lemma now follows in view of Lemma 5 and the spectral theorem for unbounded self-adjoint operators with its associated multiplicity theory. See [18] for a remarkably clear presentation of the multiplicity theory, and [1] for its application to the pushforward measure µ ◦ F −1, µ a scalar spectral measure for A. (cid:3) We can now prove the theorem by a comparison of TWα and TWαJ. Proof of Theorem 3. Since TWαJ is self-adjoint and J-symmetric, it holds that JTWαJ = TWα. Therefore (TWαJ)2 = (TWα)2. Hence the theorem follows from Lemma 6, the spectral theorem, and the symmetry of the spec- trum of TWα. (cid:3) 10 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR 4. General curvilinear polygons In this section we shall completely determine the essential spectrum of the Neumann -- Poincar´e operator associated with a curvilinear polygon Ω ⊂ C. Theorem 7. Let K : H 1/2(∂Ω) → H 1/2(∂Ω) be the Neumann -- Poincar´e operator of a C 2-smooth curvilinear polygon Ω ⊂ C with angles α1, . . . , αN . Then σess(K) =(cid:26)x ∈ R : x ≤ max 1≤j≤N(cid:12)(cid:12)(cid:12)1 − αj π(cid:12)(cid:12)(cid:12)(cid:27) . For an operator T : H 1/2(∂Ω) → H 1/2(∂Ω), denote by σea(T ) the essential spectrum of T in the sense of approximate eigenvalues. That is, λ ∈ σea(T ) n=1 ⊂ H 1/2(∂Ω) having no if and only if there is a bounded sequence (fn)∞ convergent subsequence, such that (T − λ)fn → 0. We call (fn) a singular sequence. Note that if S : H 1/2(∂Ω) → H 1/2(∂Ω) is another operator such that S − T is compact, then σea(S) = σea(T ). Lemma 8. For the Neumann -- Poincar´e operator K : H 1/2(∂Ω) → H 1/2(∂Ω) it holds that σess(K) = σea(K). Proof. Consider first K acting on H 1/2(∂Ω)/C, which only eliminates the simple isolated eigenvalue λ = 1 from the spectrum of K (K1 = 1). Since we also know that K is similar to a self-adjoint operator on H 1/2(∂Ω)/C, we have that σess(K) = σess(cid:16)KH 1/2(∂Ω)/C(cid:17) = σea(cid:16)KH 1/2(∂Ω)/C(cid:17) ⊂ σea(K). The reverse inclusion is obvious. (cid:3) We now begin the proof of Theorem 7 with a localization lemma. Recall that for a smooth function ρ on ∂Ω, we denote by Mρ the operator of multiplication by ρ on H 1/2(∂Ω). Lemma 9. Let Ω be a C 2-smooth curvilinear polygon with vertices (aj)N j=1. For each j, 1 ≤ j ≤ N , let ρj be a smooth function on ∂Ω such that ρ(x) = 1 for all x in a neighborhood of aj. Furthermore, suppose that the supports of ρj are pairwise disjoint (at a positive distance apart). Then, letting K : H 1/2(∂Ω) → H 1/2(∂Ω) be the Neumann -- Poincar´e operator, (9) Furthermore, K − NXj=1 Mρj KMρj is compact. (10) σess(K) = σea(Mρj KMρj ). N[j=1 ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 11 Proof. We construct a smooth partition of unity of ∂Ω by letting ρN +1 = j=1 ρj. We then have that 1 −PN If j 6= k or j = k = N + 1, then, by Lemma 2, Mρj KMρk is compact. This gives us the validity of (9), and hence that K = Mρj KMρk . N +1Xj=1 N +1Xk=1 σess(K) = σea(K) = σea NXj=1 Mρj KMρj . λ = 0 clearly belongs to both sides of (10). Suppose now that 0 6= λ ∈ σea(Mρk KMρk ) for some k. Let (fn)n be a corresponding singular sequence, so that (Mρk KMρk − λ)fn → 0 as n → ∞. Multiplying on the left by Mρj it follows that Mρj fn → 0 for every j 6= k, so that (fn) is a singular sequence also forPj Mρj KMρj , proving that NXj=1 Mρj KMρj . N[j=1 σea(Mρj KMρj ) ⊂ σea NXj=1 (11) Mρj KMρj fn − λfn → 0 Conversely, suppose that as n → ∞, for a sequence (fn) with no convergent subsequence. Let χ be a smooth function which is 1 on the support of ρ1, 0 on the support of ρj for every j 6= 1. Multiplying (11) with χ and noting that MχMρ1 = Mρ1Mχ = Mρ1 and that MχMρj = 0 for every other j, it follows that Mρ1KMρ1Mχfn− λMχfn → 0. Hence Mχfn is a singular sequence for Mρ1KMρ1, unless (Mχfn) has a convergent subsequence (Mχfk). In the latter case M1−χfk is a j=2 Mρj KMρj . Now the argument of this paragraph may be repeated until one finds a singular sequence for Mρj KMρj , for some j. We have hence proved that singular sequence forPN σea NXj=1 Mρj KMρj ⊂ σea(Mρj KMρj ). (cid:3) N[j=1 Let L(z) = (z + 1)/(z − 1), and let Vα = L(Wα), where Wα is the wedge of the preceding section, 0 < α < 2π. Then Vα is a lens domain, symmetric around the real and imaginary axis, with corners of angle α at −1 and 1. The next lemma says that the two corners have equal contribution to the essential spectrum of K : H 1/2(∂Vα) → H 1/2(∂Vα). 12 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR Lemma 10. Let ρ be a smooth function on ∂Vα, compactly supported in the left half-plane, such that ρ(x) = 1 for all x in a neighborhood of −1. Then σea(MρKMρ) = σess(K) =nx ∈ R : x ≤(cid:12)(cid:12)(cid:12)1 − α π(cid:12)(cid:12)(cid:12)o . Proof. Let ρ2 be the function obtained by reflecting ρ1 in the imaginary axis. Then, by symmetry, Mρ1KMρ1 is unitarily equivalent to Mρ2KMρ2, and hence the two operators have the same spectrum. Applying Lemma 9 and Theorem 3 we obtain the desired conclusion. (cid:3) Using results on perturbations by conformal mappings from [19] allows us to handle a general corner of opening α, not only the one coming from the wedge. Lemma 11. Let Ω be a C 2-smooth curvilinear polygon, and let one of its vertex points be aj, with corresponding angle αj. Let ρ be a smooth function on ∂Ω such that ρ(x) = 1 for all x in a neighborhood of aj. Then, if the support of ρ is sufficiently small, it holds that (12) σea(MρKMρ) =nx ∈ R : x ≤(cid:12)(cid:12)(cid:12)1 − αj π(cid:12)(cid:12)(cid:12)o , where K : H 1/2(∂Ω) → H 1/2(∂Ω) is the Neumann -- Poincar´e operator of Ω. Proof. Due to the local nature of the operator MρKMρ we may clearly assume, without loss of generality, that Ω only has a single corner a, of angle α. Similarly, let Uα be a smooth domain with only one corner. We suppose that this corner is at −1, and that Uα is identical to Vα in a neighborhood of −1. Lemma 10 then produces a function χ on ∂Uα such that α On the other hand, the difference KUα − MχKUαMχ is compact in this case, since there is only one corner. Hence, π(cid:12)(cid:12)(cid:12)o . σea(MχKUαMχ) = σess(KVα) =nx ∈ R : x ≤(cid:12)(cid:12)(cid:12)1 − σess(KUα) = σea(KUα) =nx ∈ R : x ≤(cid:12)(cid:12)(cid:12)1 − π(cid:12)(cid:12)(cid:12)o . α Let ϕ : Ω → Uα be a Riemann map such that ϕ(a) = −1. Lemma 4.3 of [19] then shows that ϕ is C 1,b-smooth in Ω, for 0 < b < 1, and [19, Lemma 4.4] then says that σess(K) = σess(KUα). The proof is finished by noting that σess(K) = σea(MρKMρ) by Lemma 9 applied to Ω. (cid:3) Proof of Theorem 7. The statement follows immediately from Lemmas 9, 10 and 11. (cid:3) 4.1. Final Remarks. Since K : H 1/2(∂Ω)/C → H 1/2(∂Ω)/C is similar to a self-adjoint operator, it is completely described by a spectral resolution. In other words, it is a Dunford scalar operator [9]. In the absence of a multiplicity theory (and hence classification) of Dunford scalar operators modulo compact operators, we gather a few observations which might lead to a better framework to explain the phenomena unveiled by the computations ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 13 specific to the NP operator. We keep the notation of the previous sections, but keep in a mind a more general situation. Let K denote the Neumann -- Poincar´e operator acting in the complex Hilbert space H = H 1/2(∂Ω)/C and let Kj denote its localizations (in our case Kj = Mρj KMρj ). We can assume that the supports Fj of the cut- off functions ρj are separated, so that the operator Kj acts on the closed subspace Hj of elements of H having support contained in Fj, for every j. Let Pj denote the orthogonal projection of H onto Hj. The subspaces Hj are not mutually orthogonal due to the non-locality of H, but their operator angles are almost perpendicular in the sense that PjPk is compact for every j 6= k. Denote by eT the class of an operator in the Calkin algebra L(H)/K(H). Thus fPj are mutually orthogonal projections and Moreover, From here we infer that for every polynomial q ∈ C[z] we have fKj =fPjeKfPj. eK = fK1 +fK2 + . . .gKN . q(fKj) =fPjq(eK)fPj. As K itself is a Dunford scalar operator with real spectrum, contained in the compact set σ ⊂ R, we infer kq(fKj )k ≤ kq(eK)k ≤ Ckqk∞,σ, where C is a constant. By the Stone-Weierstrass theorem, we find that ev- ery component fKj admits a continuous functional calculus with continuous functions on σ. In this sense, every fKj is a scalar operator in the Calkin algebra, with real spectrum. Their direct orthogonal sum is the class eK of of the components fKj stack on top of each other. the Neumann -- Poincar´e operator, and in this manner the essential spectra Acknowledgments The first author is grateful to Alexandru Aleman for discussions that proved essential to writing this article. The second author is grateful to Hyeonbae Kang and Habib Ammari for their unconditional interest in the spectral analysis of the NP operator. References 1. M. B. Abrahamse and Thomas L. Kriete, The spectral multiplicity of a multiplication operator, Indiana Univ. Math. J. 22 (1972/73), 845 -- 857. 2. Lars V. Ahlfors, Remarks on the Neumann-Poincar´e integral equation, Pacific J. Math. 2 (1952), 271 -- 280. 3. Habib Ammari, Giulio Ciraolo, Hyeonbae Kang, Hyundae Lee, and Graeme W. Mil- ton, Spectral theory of a Neumann-Poincar´e-type operator and analysis of cloaking due to anomalous localized resonance, Arch. Ration. Mech. Anal. 208 (2013), no. 2, 667 -- 692. 14 KARL-MIKAEL PERFEKT AND MIHAI PUTINAR 4. Habib Ammari, Giulio Ciraolo, Hyeonbae Kang, Hyundae Lee, and Kihyun Yun, Spectral analysis of the Neumann-Poincar´e operator and characterization of the stress concentration in anti-plane elasticity, Arch. Ration. Mech. Anal. 208 (2013), no. 1, 275 -- 304. 5. Habib Ammari, Youjun Deng, Hyeonbae Kang, and Hyundae Lee, Reconstruction of inhomogeneous conductivities via the concept of generalized polarization tensors, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire 31 (2014), no. 5, 877 -- 897. MR 3258359 6. Habib Ammari and Hyeonbae Kang, Reconstruction of small inhomogeneities from boundary measurements, Lecture Notes in Mathematics, vol. 1846, Springer-Verlag, Berlin, 2004. 7. S. Bergman and M. Schiffer, Kernel functions and conformal mapping, Compositio Math. 8 (1951), 205 -- 249. 8. Torsten Carleman, Uber das Neumann-Poincar´esche problem fur ein gebiet mit ecken, Almquist & Wiksells, Uppsala, 1916. 9. Nelson Dunford and Jacob T. Schwartz, Linear operators. Part III: Spectral operators, Interscience Publishers [John Wiley & Sons, Inc.], New York-London-Sydney, 1971, With the assistance of William G. Bade and Robert G. Bartle, Pure and Applied Mathematics, Vol. VII. 10. Stephan Ramon Garcia, Emil Prodan, and Mihai Putinar, Mathematical and physical aspects of complex symmetric operators, J. Phys. A 47 (2014), no. 35, 353001, 54. 11. Johan Helsing, Hyeonbae Kang, and Mikyoung Lim, Classification of spectrum of the neumann-poincar´e operator on planar domains with corners by resonance: A numer- ical study, preprint (2016). 12. Hyeonbae Kang, Mikyoung Lim, and Sanghyeon Yu, Spectral resolution of the Neumann-Poincar´e operator on intersecting disks and analysis of plasmon resonance, arXiv:1501.02952 [math.AP] (2015). 13. Rainer Kress, Linear integral equations, third ed., Applied Mathematical Sciences, vol. 82, Springer, New York, 2014. 14. Samuel Krushkal, Fredholm eigenvalues of Jordan curves: geometric, variational and computational aspects, Analysis and mathematical physics, Trends Math., Birkhauser, Basel, 2009, pp. 349 -- 368. 15. A.-W. Maue, Zur Formulierung eines allgemeinen Beugungsproblems durch eine In- tegralgleichung, Z. Physik 126 (1949), 601 -- 618. MR 0032455 (11,293i) 16. Irina Mitrea, On the spectra of elastostatic and hydrostatic layer potentials on curvi- linear polygons, J. Fourier Anal. Appl. 8 (2002), no. 5, 443 -- 487. 17. Yoshihisa Miyanishi and Takashi Suzuki, Eigenvalues and eigenfunctions of double layer potentials, arXiv:1501.03627 [math.SP] (2015). 18. Edward Nelson, Topics in dynamics. I: Flows, Mathematical Notes, Princeton Uni- versity Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1969. 19. Karl-Mikael Perfekt and Mihai Putinar, Spectral bounds for the Neumann-Poincar´e operator on planar domains with corners, J. Anal. Math. 124 (2014), 39 -- 57. 20. Mihai Putinar and Harold S. Shapiro, The Friedrichs operator of a planar domain, Complex analysis, operators, and related topics, Oper. Theory Adv. Appl., vol. 113, Birkhauser, Basel, 2000, pp. 303 -- 330. MR 1771771 (2001g:47049) 21. Johann Radon, Gesammelte Abhandlungen. Band 1, Verlag der Osterreichischen Akademie der Wissenschaften, Vienna; Birkhauser Verlag, Basel, 1987, With a fore- word by Otto Hittmair, Edited and with a preface by Peter Manfred Gruber, Edmund Hlawka, Wilfried Nobauer and Leopold Schmetterer. 22. Marius Tucsnak and George Weiss, Observation and control for operator semigroups, Birkhauser Advanced Texts: Basler Lehrbucher. [Birkhauser Advanced Texts: Basel Textbooks], Birkhauser Verlag, Basel, 2009. ESSENTIAL SPECTRUM OF THE NEUMANN -- POINCAR´E OPERATOR 15 23. W. L. Wendland, On the double layer potential, Analysis, partial differential equations and applications, Oper. Theory Adv. Appl., vol. 193, Birkhauser Verlag, Basel, 2009, pp. 319 -- 334. 24. Stephan Werner, Spiegelungskoeffizient und Fredholmscher Eigenwert fur gewisse Poly- gone, Ann. Acad. Sci. Fenn. Math. 22 (1997), no. 1, 165 -- 186. 25. S. Zaremba, Les fonctions fondamentales de M. Poincar´e et la m´ethode de Neumann pour une fronti`ere compos´ee de polygones curvilignes, Journal de Math´ematiques Pures et Appliqu´ees 10 (1904), 395 -- 444. Department of Mathematical Sciences, Norwegian University of Science and Technology (NTNU), NO-7491 Trondheim, Norway E-mail address: [email protected] Mathematics Department, University of California, Santa Barbara, Ca 93106, and School of Mathematics & Statistics, Newcastle University New- castle upon Tyne, NE1 7RU, United Kingdom E-mail address: [email protected], [email protected]
1211.3168
1
1211
2012-11-13T23:56:48
Unfolding Large Biomolecules
[ "physics.bio-ph" ]
The conformational dynamics of biomolecules drives the chemistry of life. We propose trapping large biomolecular ions in a Paul trap to probe their dynamics and that of their surrounding solvent cage.
physics.bio-ph
physics
Unfolding Large Biomolecules Erik W. Streed1,2 1Centre for Quantum Dynamics, Griffith University, Brisbane QLD, 4111 Australia 2Institute for Glycomics, Griffith University, Gold Coast QLD, 4222 Australia Abstract Summary (35 words) The conformational dynamics of biomolecules drives the chemistry of life. We propose trapping large biomolecular ions in a Paul trap to probe their dynamics and that of their surrounding solvent cage. Keywords- ion trapping; molecular structure; I. INTRODUCTION The functionality of biological molecules such as nucleic acids, proteins, and carbohydrates are driven by both their chemical composition and conformation. We propose levitating single large isolated biomolecules in an ion trap and using this uniquely adaptable gaseous phase environment to optically investigate their higher-order structure. This approach combines techniques from mass spectroscopy for biomolecule manipulation with single molecule fluorescence imaging approaches for structure detection. Increasing the net charge on the ion will also foster denaturing through Coulomb repulsion. Adding and removing single charges can thus measure the energetics processes. folding of reversibility and Environmental conditions ranging from gas-phase to fully solvated can be probed by varying the number of water molecules adhered to the biomolecular ion. Varying the temperature gives information regarding the entropy of the molecular states. For ions in the gas-phase or not fully solvated the lower end of the workable temperature range is extended beyond the aqueous freezing point. Ion trapping is a well-established technique for mass spectroscopy and high precision quantum measurements. Charged particles ranging from single electrons [1] to macroscopic dust particles [2] can be trapped for up to months at a time. Mass spectroscopy of large biomolecular ions relies on the capability to charge the particles without disruption of their delicate internal structure. Electrospray ionization (ESI) and matrix-assisted laser desorption/ionization (MALDI) are two commonly used soft ionization techniques that are used to load large, complex biomolecules such as double stranded DNA [3] or whole virus particles [4] into charged particle mass spectrometers. Above the micron level, macroscopic particles such as lycopodium pollen spores [5] can be directly ionized without the use of a solid (MALDI) or liquid (ESI) host medium. Friction with the walls of a plastic syringe when being puffed into a trap provides sufficient tribo-electric charging to remove tens of thousands of electrons per particle [2,5]. Advances in trapped atomic ion fluorescence imaging have recently demonstrated high collection efficiency [6] and wavelength scale resolution [7] with large numerical aperture optics. Existing super-resolution single molecule and fluorescence techniques from wet condensed in-vitro type environments can thus be adapted to optically monitor changes in the conformation within the ion trap environment. The value of the charge-to-mass ratio can also be measured with high precision through detection of the ion’s micromotion or via resonant parametric excitation. For larger biomolecules with an appreciable optical scattering cross-section, absorption imaging is also an intriguing option to detect conformal changes. DNA folded into chromosomes is host to numerous additional atomic ions and organic molecules that impact the structure’s stability. Ref [8] calculated that absorption imaging at 260 nm should be capable of detecting the unwrapping of 100 nm and 30 nm DNA nucleosomes. Within an ion-trapping environment the additional molecules involved in the folding stablisation process could either be captured separately during the unwrapping process or the sequence of their mass loss quantitatively measured. In addition the lack of surrounding solution minimizes the complications from absorption and scattering. Spectra Imaging Counting Electrospray Ionziation Loading N2 -- -- + + ++++ +++ + + + High NA Optics Yb+ Yb+ e- e- e- Loading Gate ---- m H u i g u h c a v v a w c u o u eGun L m Linear Quadrupole Ion Trap RF e- ö ö t Sympathetic Cooling/Heating Fluorphore Excitation Molecular Beam RF RF Figure 1. Biomolecular ion loading and trapping apparatus. Biomolecules (red) in solution (blue) are electrospray ionised in a low vacuum chamber against a counter-propaging stream of dry nitrogen (green). The ions are focused through a differential pumping aperture for transient loading into a linear quadrupole ion trap with a time gated voltage inversion on the end-ring electrode. Ions are manipulated through interaction with molecular beams, electron scattering, and laser excitation. Fluorescence collected with high Numerical Aperture (NA) optics can be analysed in the time, wavelength, or spatial domains. Sympathetic thermalisation with trapped atomic ions (e.g. Yb+ or Ba+) can provide indirect laser cooling or heating in a high vacuum environment without the complication of varying a buffer gas temperature. Fig. 1 depicts an apparatus investigation of for biomolecular folding in an ion trap. The four key components of this device are the mechanisms for loading, confinement, manipulation, and detection of the ions. The mechanics of loading and confining biomolecular ions is well-established technology and will not be covered further. Approaches to manipulation of the biomolecular ion’s internal states and fluorescence detection are discussed in subsequent sections. II. MANIPULATION OF BIOMOLECULAR IONS Changing the shape of a biomolecule allows us to investigate the processes which govern its structure and dynamics. X-ray crystallography can provide a detailed information about the electron distribution of a crystalised molecule, but it does so for a static configuration and lacks information about the dynamic flexing and other conformational changes the molecule can undergo. These are particularly important when a conformation change is associated with a biologically relevant process such as drug binding or an enzymatic catalysis. The transformation from physiologically active native state to a detectably denatured variant depends on the mechanisms available to precipitate such a change. In solution this change is accomplished by varying parameters including pH, temperature, solvent composition, and concentration of other ion species with response measured through solubility, optical or NMR spectra, optical activity, or reactivity. The ensemble nature of these measurements can make it challenging to interpret the results in the context of biomolecular structure, such as salting-in and salting-out curves for solubility. In mass spectroscopy the fragmentation pattern from the disruption of covalent bonds can give clues to the higher order structure, however it is dependent on the electron rearrangement behavior during the fragmentation process. With a single biomolecular ion confined in a trap, the charge state can be changed a single electron at a time. The Coulomb force from the distribution of the net charge competes with the internal forces of interest binding the molecule in a particular set of conformations. Fig. 2 illustrates how the hierarchy of structures would be ideally unraveled through this Coulombic mechanis. In the large charge and high temperature limit mass spectroscopy style fragmentation of the biomolecular ion into multiple species occurs at a rate related to the amount of Coulomb repulsion, available thermal energy, and relative favourability of the fragmented structure. Since the net charge of the molecule is only changed in single electron steps, this provides a known self-calibration for the effect of the change in charge. Electrons can be removed selectively through photo-ionisation or less discriminately with low energy electron scattering collisions. Deliberate driving of the ion’s motion in a buffer gas filled trap environment has been observed to increase the charge through the tribo-electric effect [2], while an ion near rest will have its net charge quenched through buffer gas collisions. These later processes offer a much gentler alternative to charging the ion through the use of collisional friction interactions. The success of this process depends highly on the availability to add or remove charge without adverse effects to the overall structure. If the change in charge comes from disrupting a covalent bond, the primary structure will be altered. Unlike in aqueous solution, the difference between Lewis and Brøstend-Lowry acids and bases has an important distinction. In trapping Lewis type systems the donation or acceptance of additional electrons is easily reversible. The donation or acceptance of additional protons (H+) in a Brøstend-Lowry type dynamic may not be either possible (acceptance) or reversible (donation) due to loss when a proton is liberated, either as a charged particle or a neutral. This is not surprising given that multiple protonation states are readily seen as a side effect of electrospray loading [9]. + + + Low Charge Quaternary Complex Ionize Deionize + + + + Ionize Deionize? + + + + + Ionize + Deionize Tertiary Structure Secondary Structure + + + + ++ + + + + + + Highly Charged Primary Structure Figure 2. Charge denaturing via Coulomb repulsion. Higher-order structure is unraveled as the Coulombic energy exceeds the binding energies. No biomolecule functions in isolation. Understanding the impact of the surrounding environment is thus a crucial aspect to unraveling its dynamics. Water molecules prefer proximity to the surface of a highly soluble biomolecule rather than bulk water. This affinity results in the biomolecule moving within a loose “cage” of solvent layers surrounding it. A particularly dramatic instance of the importance of solvation is found in green fluorescent protein. While the fluorophore dipole is buried deep inside this protein, it does not exhibit fluorescence outside of solution. As mentioned above, aqueous solution also provides a reservoir of protons and other species that influence the internal equilibria. + + :O :O :O :O :O :O :O :O + + Droplet Vacuum Humidity + :O + :O :O :O + Vacuum+Heat? + Humidity + + + + Solvent Cage Solvent Free Figure 3. Solvent cage manipulation. Exchange of water molecules with the background gas can either precipitate a droplet or render a dehydrated ion. Within an ion trap environment the level of solvation can be manipulated by varying the partial pressure of water vapour above or below saturation. Fig. 3 illustrates the inter- conversion stages between a solvated droplet and a solvent free biomolecule. Higher humidity will promote condensation of water molecules on the biomolecular ion, leading to the same mechanics that are observed in droplet formation in the highly solvated limit. Reducing the available water vapour will promote evaporation. The more tightly bound final layers of water in the solvent cage may require an extended duration to evaporate off unless additional energy is added to liberate them. Desorption could also be selectively encouraged by illuminating the molecule/solvent system with infrared lasers in the O-H stretching water absorption bands at 1.95 or 2.9 µm. The later wavelength was successfully used for IR laser based MALDI from water ice [10], a situation similar to the semi- crystalised solvent cage state. Complicating matters the presence of water layers around the biomolecule will likely act as a shield against charge manipulation techniques. Conversely the presence of water provides a substrate acid/base chemistry. III. FLUORESCENCE DETECTION Single molecule fluorescence techniques can provide a sensitive and non-destructive measure of the internal dynamics of large biomolecules within an ion trap. Single fluorophore systems are easier to realise [9] but have demanding absolute localisation requirements to be suitable for super-resolution imaging [10] that may not be feasible for trapped biomolecular ion experiments. Wavelength-scale resolution trapped atomic ions have realised fluorescence imaging with one nanometer accuracy (FIONA), however this was achieved at a temperature of a few milli-Kelvin (fitting of Fig 2d from [8]). Biomolecular systems with multiple fluorophores are less ubiquitous, but greatly simplify the technical demands on ion trapping and detection. Förster Resonance Energy Transfer (FRET) enables precision proximity measurements at the 4-7 nm range [11] for fluorophore pairs with an overlap in their emission and excitation spectra. Greater distances between fluorophore pairs with spectrally separable emission can be resolved through colocalisation techniques at the tens or hundreds of nanometer level. FRET provides a rapid measure of the distance between fluorophores in the ratio of their relative emission rates. When the two fluorophores are sufficiently close emissions from the higher wavelength “donor” fluorophore are absorbed and re- emitted by the lower wavelength “acceptor” fluorophore. The effective detection range occurs when the donor only fractionally converts into an acceptor photon. The dipole emission and absorption patterns also make FRET to changes in orientation of the fluorophores. This makes FRET highly sensitive, though complicated in precisely disentangling distance from orientation effects without additional polarisation information. An advantage of this approach is that imaging isn’t required, only collection and spectral separation of the fluorescence. This makes the common-mode motion of the biomolecular ion unimportant to detection, so long as the overall excursion does not exceed the collection area of the detector. The time domain variations in the emission ratio provide information on the roto-vibrational modes linking the two groups. Colocalisation is a more flexible but less sensitive super- resolution imaging method of measuring distances between fluorophore pairs. For each fluorophore an exposure with multiple photons is acquired and a centroid location calculated. The technique relies on imaging with a sufficient degree of chromatic correction, to prevent aberrations such as lateral colour from distorting the signals. The sensitivity to separation measurement depends on the imaging system resolution and the number of photons that can contribute to estimating the centroid location. The relative distances between the fluorophore locations can then be compared. Motion of the ion below the optical resolution limit does not appreciably degrade the centroid accuracy so long as the resolution is the dominant contribution to the RMS width of the imaged spot. The ion motion blurring and the resolution limit are both averaged away in finding the centroid. This is unlike deterministic absolute positioning super-resolution techniques such as STimulated Emission Depletion (STED) or Ground State Depletion (GSD), in which the motion of the ion must be reduced below the position sensitivity. Colocalisation provides a projection of the relative fluorphore positions onto a particular plane. An advantage of the ion trap environment over condensed phase experiments is that the orientation of the biomolecular ion can be controlled by varying the trap parameters, thus facilitating recovery of the three dimensional positioning in colocation imaging of a trapped biomolecule. A biomolecule with a permanent electric dipole will orientate itself along the direction of the residual uncompensated electric field. In the case of no electric dipole or the residual electric compensated to zero the direction of the strongest quadrupole moment will likewise align itself with the weakest axis of confinement, and conversely the weakest quadrupole moment will align itself with the strongest axis of confinement. Selecting biomolecules for preliminary experiments attempts to balance their viability to demonstrate an effect with utility of that effect. Large biomolecules can fluoresce due to either the composition of their constitute components (ex. Phenylalanine, Tyrosine, and Tryptophan amino acids in proteins), the nature of their final functional form (green fluorescent protein), or through the intentional addition of dye groups (PicoGreen™ in DNA). Intrinsic fluorescence is preferable due to the lack of impact on the structure, however the fluorescence dipole may not be located in an interesting location spatially or spectrally and the likelihood of a pair of fluorescent groups suitable for FRET is low. ACKNOWLEDGMENT Funded by Griffith University Strategic Investment in the Physical Sciences. The author thanks John Chodera for his helpful discussions and introductions into the protein folding community. REFERENCES [1] G. Gabrielse, H. Dehmelt, and W. Kells, “Observation of a Relativistic, Bistable Hysteresis in the Cyclotron Motion of a Single Electron” Phys. Rev. Lett. 54, 537 (1985). [2] W. Winter & H. W. Ortjohann “Simple demonstration of storing macroscopic particles in a ‘Paul trap’” Am. J. Phys. 59 807 (1991) [3] F Kirpekar, S. Berkenkamp, F Hillenkamp, “Detection of Double-Stranded DNA by IR- and UV-MALDI Mass Spectroscopy” Anal. Chem. 71, 2334 (1999) [4] M. A. Tito, K. Tars, K Valegard, J. Hajdu, C. V. Robinson “Electrospray Time-of-Flight Mass Spectrometry of the Intact MS2 Virus Capsid” J. Am. Chem. Soc. 122, 3550-3551 (2000) [5] M. Kumph, M. Brownnutt, R. Blatt “Two-Dimensional Arrays of RF Ion Traps with Addressable Interactions” New J. Phys. 13, 073043 (2011) [6] E. W. Streed, B. G. Norton, A. Jechow, T. J. Winehold, D. Kielpinski “Imaging of Trapped Ions with a Microfabricated Optic for Quantum Information Processing” Phys. Rev. Lett. 106, 010502 (2011) [7] A. Jechow, E. W. Streed , B. G. Norton, M. J. Petrasiunas , & D. Kielpinski, “Wavelength-scale imaging of trapped ions using a phase Fresnel lens” Opt. Lett. 36, 1371 (2011) [8] E. W. Streed, A. Jechow, B. G. Norton, D. Kielpinski “Absorption imaging of a single atom” Nat. Comm. 3 933 (2012) [9] D. Offenberg, C. B. Zhang, Ch. Wellers, B. Roth, S. Schiller “Translational cooling and storage of protonated proteins in an ion trap at subkelvin temperatures” Phys. Rev. A 78, 061401R (2008). [10] A. Pirkl, J. Soltwisch, F. Draude, K. Dreisewerd “Infrared matrix-assisted laser desorption/ionization orthogonal-time-of- flight mass spectrometry employing a cooling stage and water ice as a matrix.” Anal. Chem. 84 5669 (2012). [11] H. E. Grecco, P. J. Verveer “FRET in Cell Biology: Still Shining in the Age of Super-Resolution?” Chem. Phys. Chem. 12 484 (2011)
1004.1728
2
1004
2010-05-20T05:36:37
Cavity-water interface is polar
[ "physics.bio-ph", "cond-mat.soft" ]
We present the results of numerical simulations of the electrostatics and dynamics of water hydration shells surrounding Kihara cavities given by a Lennard-Jones (LJ) layer at the surface of a hard-sphere cavity. The local dielectric response of the hydration layer substantially exceeds that of bulk water, with the magnitude of the dielectric constant peak in the shell increasing with the growing cavity size. The polar shell propagates into bulk water to approximately the cavity radius. The statistics of the electrostatic field produced by water inside the cavity follow linear response and approach the prediction of continuum electrostatics with increasing cavity size.
physics.bio-ph
physics
Cavity-water interface is polar Center for Biological Physics, Arizona State University, PO Box 871604, Tempe, AZ 85287-1604 Allan D. Friesen and Dmitry V. Matyushov We present the results of numerical simulations of the electrostatics and dynamics of water hy- dration shells surrounding Kihara cavities given by a Lennard-Jones (LJ) layer at the surface of a hard-sphere cavity. The local dielectric response of the hydration layer substantially exceeds that of bulk water, with the magnitude of the dielectric constant peak in the shell increasing with the growing cavity size. The polar shell propagates into bulk water to approximately the cavity radius. The statistics of the electrostatic field produced by water inside the cavity follow linear response and approach the prediction of continuum electrostatics with increasing cavity size. PACS numbers: 77.22.-d, 87.15.hg, 61.20.Ja, 61.25.Em Keywords: Interfacial electrostatics, hydrophobicity, dewetting, Stokes-shift dynamics, dipole solvation, wa- ter interface Nanoscale interfaces of polar liquids combine strong distortions of the liquid density profile with highly per- turbed long-range electrostatic correlations. The inter- facial density, and the related diffusional dynamics, are governed by short-range packing restrictions and are rel- atively short-ranged. Nevertheless, the density fluctua- tions of the interfacial region are critical for the long- range hydrophobic forces [1, 2] and the related weak dewetting of interfaces of non-polar solutes [3, 4]. In con- trast, electrostatic interactions, and the orientational cor- relations of multipolar moments, are long-ranged. They are assigned macroscopic length-scale in the Maxwell (continuum) electrostatics propagating the effect of par- tial charges at dielectric interfaces on the length-scale of the Coulomb potential. Whether this picture is cor- rect for polar liquids and how the surface polarization is screened by the mobile liquid dipoles remains an open question [5], the resolution of which will define the limits of continuum electrostatics in application to nanoscale interfaces of molecular liquids. Water presents a particular challenge to the prob- lems of interfacial dynamics and thermodynamics since energetically strong hydrogen bonds add a short-range scale competing with long-range electrostatic forces. The properties of hydration layers surrounding nanoscale so- lutes indeed turn out to be unusual. Apart from ubiq- uitous hydrophobic interactions linked to the structure of the water interface [1], measurements of microscopic electrostatics of the protein/water interface have shown some surprising results. Electrostatics on the microscopic scale is traditionally probed by the dynamic and static band-shifts of optical dyes [6]. The corresponding Stokes- shift dynamics at the protein/water interface showed a long exponential decay absent for the free chromophores in solution. This observation has prompted the label of "biological water" for hydration layers of biopolymers [7]. While the cause of this effect is still debated [8, 9], various extent of slowing of the collective Stokes shift dynamics has been universally observed at protein/water interfaces [7, 8]. Further, the hydration shells around proteins were found to carry high local polarity [10], thus linking the slower dynamics to the structural reorganization of wa- ter in the form of a polarized cluster around proteins [11]. Unfortunately, the problem of the protein hydra- tion is inseparable from the complex protein dynamics [12]. Studies excluding this latter component are there- fore necessary to understand the electrostatics of hydra- tion layers when the solute size grows to the nanoscale. This is the goal of this report. Here we present extensive Molecular Dynamics (MD) simulations of the structure of water around non-polar solutes (cavities). In contrast to previous active research in this field [2 -- 4, 13], we ask here the following questions: (i) how polar is the interface? (ii) how far into the bulk does the polarity perturbation propagate? and (iii) how are the orientational dipolar dynamics of the hydration layers affected by the solute? The main result of this study is the observation of a significant increase of the local water polarity at the interface, with the region of enhanced polarity extending into the bulk to approxi- mately the cavity radius. The water nanoscale interface was modeled by insert- ing spherical solutes carrying a hard-sphere (HS) core surrounded by a Lennard-Jones (LJ) potential layer. The interactions with the SPC/E oxygen is then given by the Kihara solute-solvent potential φ(r) = 4ǫLJ"(cid:18) σ r − rHS(cid:19)12 −(cid:18) σ r − rHS(cid:19)6# . (1) The LJ well has the width σ = 3 A and the energy ǫLJ for which two values were used: ǫLJ = 0.65 kJ/mol equal to the LJ energy between oxygens of SPC/E water and ǫLJ = 20 kJ/mol close to the energy of hydrogen bonds in bulk water. The cavity size was varied by changing the HS radius rHS in the range 0 -- 12 A. The number of waters in the simulation cell was varied to allow sufficiently large solvation layers, with 4053 and 11845 hydration waters used for the smallest and largest solutes, respectively. The trajectories for analysis were 5 ns long, following 100 -- 500 ps equilibration. Simulations were performed 0 -0.1 -0.2 2I p 0 10 2.0 3.0 0.06 0.03 0 20 R/Å 10 30 40 FIG. 1. The second-order orientational order parameter pI 2 of the water first shell vs the cavity radius R = rHS + σ. The solid diamonds and triangles refer to cavities in water with ǫLJ = 0.65 and 20 kJ/mol, respectively. The open points refer to cavities in the fluid of dipolar hard spheres [5] with the reduced dipole moments (m∗)2 = βm2/σ3 s equal to 2.0 (open squares) and 3.0 (open circles); m is the dipole moment and σs is the hard-sphere diameter of the solvent. The inset shows the orientational parameter pI 1 of the first hydration layer. with cubic periodic boundary conditions, at 273 K and zero pressure, with Berendsen thermostat and barostat and a timestep of 2 fs. Ewald sums with tin foil boundary conditions were used for electrostatic interactions. A spherical solute induces a spherical symmetry The breaking in an otherwise orientational consistent structure of with this imposed symmetry is characterized by the first- and second-order orientational order parame- isotropic liquid. the interface rj · mjE and p2(r) = ters: p1(r) = (N (r))−1DPrj <r (2N (r))−1DPrj <r[3(rj · mj)2 − 1]E. These parameters project the unit dipolar vectors mj within the shell of radius r on the radial direction rj = rj/rj, N (r) is the number of waters within the shell. The first hydration layer is then defined as R ≤ r ≤ R + 1.5 A, R = rHS + σ 1 and pI and the corresponding order parameters are pI 2. The dielectric constant of a cavity-water mixture can be defined in terms of the volume occupied by the cavity relative to the volume of water [14]. The solute-solvent response function describing the interfacial polarization can then be calculated by accounting for dipolar fluc- tuations accumulated within a radial water layer sur- rounding the cavity χ(r) = βh(δM(r))2i/(3V (r)); β = 1/(kBT ) is the inverse temperature. The water dipole moment M(r) in this equation is taken over the radial shell between the spherical cavity and radius r extend- ing from the cavity center to the bulk, V (r) is the shell volume. The limit of infinite dilution yields the bulk di- electric susceptibility of water χ = χ(∞) = (ǫ − 1)/(4π), where ǫ is the water dielectric constant. The fluctuation susceptibility χ(r) determines the di- electric constant ǫ(r) accumulated within the radial layer. For simulations employing periodic boundary conditions with tin-foil boundary around replicas of the simulation cell the connection between the simulated variance of the 2 dipole moment and the dielectric constant is particularly simple [15]: ǫ(r) = 1 + 4πχ(r). The macroscopic dielec- tric constant of water is then ǫ = ǫ(∞). The orientational structure of polar liquids at inter- faces is strongly affected by short-range orientational cor- relations. Unsaturated hydrogen bonds of surface wa- ters produce preferential in-plane orientations of water's dipoles [16] as reflected by the first and second orien- tational order parameters (Fig. 1). The first-order pa- rameter pI 1 is nearly zero pointing to no preferential ra- dial orientation (inset in Fig. 1), while pI 2 is non-zero and negative, in accord with the preferential in-plane orienta- tion of the dipoles. This orientational pattern is specific for water and does not necessarily repeat itself in other polar liquid. For comparison, pI 2 of hard-sphere dipoles at the surface of a spherical cavity [5] passes through a minimum (Fig. 1). Orientational order first grows with increasing cavity size as frustrations of dipolar orienta- tions are released with the growing number of dipoles. However, further increase of the cavity size leads to a weak dewetting of the interface by the puling force of the liquid [3] with the resulting destruction of the interfacial orientational order. In contrast, water preserves its par- allel interfacial order, mostly determined by its hydrogen- bond network and not much effected by the strength of the solute-solvent LJ attraction (Fig. 1). The function ǫ(r) calculated for shells around cavities is compared to the same function calculated from shells around Lorentz's virtual cavity [14] (water molecules be- tween radii R and r) taken from configurations of pure water without cavity inserted. The difference of the two functions shows a sharp peak (Fig. 2a) pointing to an ef- fectively higher polarity of hydration shells around cavi- ties compared to shells in bulk water. In order to distinguish between the orientational and density origins of the peak in the dielectric constant, we have plotted in Figs. 2(b,c) χ(r) defined as above in comparison with the susceptibility normalized to the number of waters in the shell N (r): χN (r) = βρh(δM(r))2i/(3N (r)), where ρ is the number density of bulk water. This comparison shows that the origin of ∆ǫ is a composite effect of changes in both the local density and orientational structure. Even though a significant part of ∆ǫ comes from the increased density in the first solvation layer, particularly at the large solute-solvent LJ attraction (Fig. 2(c)), the effect cannot be cast in terms of N (r) only. Given the long range of the interfacial orientational order one wonders how the dynamics of hydration layers are affected. We have looked at several correlation func- tions. χI (t) = βh(δMI (t) · δMI(0)i/(3V I ) is the time self-correlation function of the dipole moment of the first solvation layer and χ(r, t) is a similar correlation function extended to a layer within the radius r from the cavity's center. In addition, we have calculated the correlation function CE(t) = hδEs(r, t)·δEs(r, 0)i of the electric field 300 200 100 (a) 3 Å ε ∆ 12 Å 7.5 Å ε ε LJ = 0.65 LJ = 20 0 0 80 60 40 20 0 χ π 4 5 10 15 20 25 r/Å (b) ε LJ = 0.65 5 10 virt. 15 r/Å 20 25 200 100 0 ε LJ = 20 (c) V N virt. 20 25 5 10 15 r/Å FIG. 2. Dielectric constant of the hydration layer relative to the dielectric constant of the same layer around a virtual cav- ity for three cavity sizes indicated in the plot (a). The solid and dashed lines refer to ǫLJ = 20 and 0.65 kJ/mol, respec- tively. The response functions defined through the volume of the shell ("V", solid lines) and through the number of shell waters ("N", dashed lines) are shown in (b) for ǫLJ = 0.65 kJ/mol and in (c) for ǫLJ = 20 kJ/mol; R = 7.5 A. χ(r) for the virtual cavity is marked as "virt." 60 s p / 40 E τ 12 Å 7.5 Å τ D 20 0 0 3 Å 10 ε LJ = 0.65, R = 12 Å 20 r/Å 30 FIG. 3. Exponential relaxation time of χI (t) (closed dia- monds), CE(r, t) (open circles), and χ(r, t) (open squares); ǫLJ = 20 kJ/mol. Also shown is the exponential relaxation time of the self-correlation function of the unit vector eI (t) representing the dipole moment of the first-shell waters. Dif- ferent cavity sizes for the r-dependent relaxation times are indicated in the plot. The filled circles refer to R = 12 A and ǫLJ = 0.65 kJ/mol and the horizontal dotted line indicates the Debye relaxation time τD of pure SPC/E water. The dashed lines in the plot are connecting the points. Es(r, t) produced at the cavity's center by the waters within the r-shell. This latter correlation function repre- sents the Stokes shift dynamics of dipolar chromophores [7, 8]. All correlation functions were fitted to a sum of a ballistic Gaussian decay and an exponential tail [6]. Figure 3 presents the compilation of the results for the exponential relaxation time τE. The main observation from the dynamics calculations is a significant growth of the exponential relaxation time of the first solvation layer with the cavity size (filled dia- 3 monds in Fig. 3). Consistent with the results for ∆ǫ(r), this dynamics perturbation propagates into the bulk to at least the distance of the cavity radius. The expo- nential decay time of CE(r, t) retraces the correspond- ing time from χ(r, t). The dipolar dynamics of the first solvation layer are dominated by rotations of the dipole moment MI, instead of its magnitude fluctuations, as is seen from the self-correlation function of the unit vector eI (t) = MI (t)/M I (t) (crosses in Fig. 3). This observa- tion points to a high level of orientational cooperativity in the first solvation layer, which does not decorrelate by individual dipole rotations and instead rotates slowly as a correlated dipolar domain. This dynamical slowing is however seen only for the larger LJ attraction, ǫLJ = 20 kJ/mol, and no effect of the solute on the dynamics is observed when the solute-solvent LJ potential is similar to that in water, ǫLJ = 0.65 kJ/mol (closed points in Fig. 3). This result might help to explain the conflicting liter- ature [17] on the subject of the surface-induced alteration of the liquid dynamics. The outcome seems to be con- trolled by the strength of the solute-solvent interactions and thus the surface composition. Our results partially support Onsager's concept of "in- verted snowball" dynamics [18]. It stipulates that solva- tion dynamics are slower close to a newly created charge compared to more distant layers, ranging between the one-particle (slow) orientational diffusion and the dielec- tric (fast) relaxation of the bulk. The data in Fig. 3 in- deed show a speedup of dipolar relaxation into the bulk. However, this effect strongly depends on both the cavity size and the solute-solvent LJ interaction. It is expected to be essentially absent for typical optical probes [6] con- sistent in size with the smallest cavity studied here. Fur- ther, even for larger cavities, the slow dynamics of the closest solvation layers are almost lost when the hydra- tion layer is grown to the boundaries of the simulation cell. At that point, the relaxation time becomes the De- bye relaxation time of bulk water (Fig. 3). This obser- vation implies that dipolar optical probes placed inside the cavity [7 -- 9] will not pick up the slowing of the closest hydration shells and instead will average the effect out by the electric field contributions from more distant layers. The statistics of dipolar interfacial fluctuation can be probed by the chemical potential of electrostatic solva- tion, i.e. the free energy of interaction of the charges in- side the cavity with the surrounding water solvent. We found that the dipolar field M(r) is Gaussian and thus the linear response approximation should be applicable. The solvation chemical potential of a charge µq or a dipole µd can then be found from the variance of the elec- trostatic potential φs or the electric field Es produced by the solvent at the position of the corresponding multipole [19]: µq = −β(q0)2h(δφs)2i/2, µd = −β(m0)2h(δEs)2i/6. Here, q0 and m0 are, correspondingly, the charge and point dipole within the cavity. The averages in these re- lations do not depend, in linear response, on whether the 0.65 1.6 1.2 20 d* µ 0.006 0.004 0.002 s 〉 E 〈 0.8 0.4 O 4 0 0 2 8 10 4 6 m0/D 8 R/Å 12 d = µdm2 0/R3 of solvating FIG. 4. Chemical potential µ∗ the point dipole m0 at the center of the cavity of radius R. The solid points (ǫLJ = 20 kJ/mol, circles and 0.65 kJ/mol, squares) and obtained from the variance of the water electric field inside the empty cavity, µd ∝ h(δEs)2i. The open trian- gles are from the thermodynamic integration of the average electric field hEsi produced by dipoles of increasing magni- tude (0 < m0 < 10 D) positioned at the cavity center. The average field (eV/D) is a linear function of the dipole magni- tude m0 (inset, R = 12 A, ǫLJ = 0.65 kJ/mol). The dashed line in the inset is the prediction based on the field variance inside the empty cavity (not the best fit). The dotted hori- zontal line is the result of continuum electrostatics given by the Onsager relation ("O"). corresponding multipoles are actually present inside the cavity [19]. They can therefore be calculated from our simulations with empty cavities providing insights into how these solvation free energies scale with the size of the solute and whether the limit of continuum electro- statics is reached. We found noticeable effects of the size of the sim- ulation cell on h(δφs)2i and much smaller size effects on h(δEs)2i. We have therefore chosen to look at the statistics of the field fluctuations and the corresponding chemical potential of dipole solvation. These results are d = µdR3/(m0)2. This summarized in Fig. 4 showing µ∗ dimensionless parameter is expected to approach, with growing cavity size, the size-independent limit of contin- uum electrostatics, given by the Onsager equation [14] µd = (ǫ − 1)/(2ǫ + 1) ≃ 0.5. The values of µd, al- though noticeably higher at intermediate sizes, indeed seem to approach this limit. The open points in Fig. 4 are obtained by thermodynamic integration of average energies of point dipoles placed at the cavity center. A good agreement between µd from the field variance and from the thermodynamic integration, as well as the lin- ear dependence of hEsi vs m0 (inset in Fig. 4), testifies to the validity of the linear response approximation. The chemical potential µd is not strongly affected by the strength of the solute-solvent LJ attraction. The range of ǫLJ values studied here probably covers most of situations of practical interest. Deviations of the elec- tric field variance from the area between the two curves 4 shown in Fig. 4 might therefore be used to identify nonlin- ear solvation typically associated with hydrogen bonding between water and the solute [20]. The results obtained here must have significant impli- cations for self-assembly of nano-sized objects and bio- logical activity of hydrated biopolymers. Polar solvation layers around solutes are expected to screen the inside charges. In the crowded environment of a living cell [21] this screening will reduce interactions between multipo- lar solutes. The high-polarity layer is also characterized by slower dipolar solvation. However, this effect is is not picked up by the dynamics of dipolar probes placed in- side the cavity. A slow relaxation component observed in optical time-resolved spectra [7] therefore needs to be assigned to protein motions pushing the hydration layers [9]. This research was supported by the DOE, Chemi- cal Sciences Division, Office of Basic Energy Sciences (DEFG0207ER15908). [1] P. Ball, Chem. Rev. 108, 74 (2008). [2] D. Chandler, Nature 437, 640 (2005). [3] G. Hummer and S. Garde, Phys. Rev. Lett. 80, 4193 (1998). [4] B. J. Berne, J. D. Weeks, and R. Zhou, Annu. Rev. Phys. Chem. 60, 85 (2009). [5] D. R. Martin and D. V. Matyushov, Europhys. Lett. 82, 16003 (2008). [6] R. Jimenez, et al. Nature 369, 471 (1994). [7] S. K. Pal and A. H. Zewail, Chem. Rev. 104, 2099 (2004). [8] L. Zhang, et al. Proc. Natl. Acad. Sci. 104, 18461 (2007). [9] L. Nilsson and B. Halle, Proc. Natl. Acad. Sci. 102, 13867 (2005). [10] D. N. LeBard and D. V. Matyushov, Phys. Rev. E 78, 061901 (2008). [11] S. Ebbinghaus, et al. Proc. Natl. Acad. Sci. 104, 20749 (2007). [12] H. Frauenfelder, et al. Proc. Natl. Acad. Sci. 106, 5129 (2009). [13] S. Sarupria and S. Garde, Phys. Rev. Lett. 103, 037803 (2009). [14] B. K. P. Scaife, Principles of dielectrics (Clarendon Press, Oxford, 1998). [15] M. Neumann, Mol. Phys. 57, 97 (1986). [16] C. Y. Lee, J. A. McCammon, and P. J. Rossky, J. Chem. Phys. 80, 4448 (1984). [17] F. He, L.-M. Wang, and R. Richert, Eur. Phys. J. 141, 3 (2007). [18] L. Onsager, Can. J. Chem. 55, 1819 (1977). [19] D. Ben-Amotz, F. O. Raineri, and G. Stell, J. Phys. Chem. B 109, 6866 (2005). [20] P. K. Ghorai and D. V. Matyushov, J. Phys. Chem. A 110, 8857 (2006). [21] J. F. Douglas, J. Dudowicz, and K. F. Freed, Phys. Rev. Lett. 103, 135701 (2009).