Research article Special Issues

Duals of Gelfand-Shilov spaces of type K{Mp} for the Hankel transformation

  • For μ12, and under appropriate conditions on the sequence {Mp}p=0 of weights, the elements, the (weakly, weakly*, strongly) bounded subsets, and the (weakly, weakly*, strongly) convergent sequences in the dual of a space Kμ of type Hankel-K{Mp} can be represented by distributional derivatives of functions and measures in terms of iterated adjoints of the differential operator x1Dx and the Bessel operator Sμ=xμ12Dxx2μ+1Dxxμ12. In this paper, such representations are compiled, and new ones involving adjoints of suitable iterations of the Zemanian differential operator Nμ= xμ+12Dxxμ12 are proved. Prior to this, new descriptions of the topology of the space Kμ are given in terms of the latter iterations.

    Citation: Samuel García-Baquerín, Isabel Marrero. Duals of Gelfand-Shilov spaces of type K{Mp} for the Hankel transformation[J]. AIMS Mathematics, 2024, 9(7): 18247-18277. doi: 10.3934/math.2024891

    Related Papers:

    [1] Songxiao Li, Jizhen Zhou . Essential norm of generalized Hilbert matrix from Bloch type spaces to BMOA and Bloch space. AIMS Mathematics, 2021, 6(4): 3305-3318. doi: 10.3934/math.2021198
    [2] Keun Young Lee, Gwanghyun Jo . The dual of a space of compact operators. AIMS Mathematics, 2024, 9(4): 9682-9691. doi: 10.3934/math.2024473
    [3] Babar Sultan, Mehvish Sultan, Aziz Khan, Thabet Abdeljawad . Boundedness of an intrinsic square function on grand $ p $-adic Herz-Morrey spaces. AIMS Mathematics, 2023, 8(11): 26484-26497. doi: 10.3934/math.20231352
    [4] Gang Wang . Some properties of weaving $ K $-frames in $ n $-Hilbert space. AIMS Mathematics, 2024, 9(9): 25438-25456. doi: 10.3934/math.20241242
    [5] Liu Yang, Ruishen Qian . Volterra integral operator and essential norm on Dirichlet type spaces. AIMS Mathematics, 2021, 6(9): 10092-10104. doi: 10.3934/math.2021586
    [6] Kifayat Ullah, Junaid Ahmad, Hasanen A. Hammad, Reny George . Iterative schemes for numerical reckoning of fixed points of new nonexpansive mappings with an application. AIMS Mathematics, 2023, 8(5): 10711-10727. doi: 10.3934/math.2023543
    [7] Rana Safdar Ali, Saba Batool, Shahid Mubeen, Asad Ali, Gauhar Rahman, Muhammad Samraiz, Kottakkaran Sooppy Nisar, Roshan Noor Mohamed . On generalized fractional integral operator associated with generalized Bessel-Maitland function. AIMS Mathematics, 2022, 7(2): 3027-3046. doi: 10.3934/math.2022167
    [8] Bo Xu . Bilinear $ \theta $-type Calderón-Zygmund operators and its commutators on generalized variable exponent Morrey spaces. AIMS Mathematics, 2022, 7(7): 12123-12143. doi: 10.3934/math.2022674
    [9] Shahid Mubeen, Rana Safdar Ali, Iqra Nayab, Gauhar Rahman, Thabet Abdeljawad, Kottakkaran Sooppy Nisar . Integral transforms of an extended generalized multi-index Bessel function. AIMS Mathematics, 2020, 5(6): 7531-7547. doi: 10.3934/math.2020482
    [10] Suixin He, Shuangping Tao . Boundedness of some operators on grand generalized Morrey spaces over non-homogeneous spaces. AIMS Mathematics, 2022, 7(1): 1000-1014. doi: 10.3934/math.2022060
  • For μ12, and under appropriate conditions on the sequence {Mp}p=0 of weights, the elements, the (weakly, weakly*, strongly) bounded subsets, and the (weakly, weakly*, strongly) convergent sequences in the dual of a space Kμ of type Hankel-K{Mp} can be represented by distributional derivatives of functions and measures in terms of iterated adjoints of the differential operator x1Dx and the Bessel operator Sμ=xμ12Dxx2μ+1Dxxμ12. In this paper, such representations are compiled, and new ones involving adjoints of suitable iterations of the Zemanian differential operator Nμ= xμ+12Dxxμ12 are proved. Prior to this, new descriptions of the topology of the space Kμ are given in terms of the latter iterations.



    As it is well known, Schwartz developed the theory of distributions in the late 1940's; a detailed exposition appears in his monograph [29]. Generalized functions of any kind, as well as their use to solve the Cauchy problem, were introduced by Gelfand and Shilov around 1953. In the period of 1956–58, these two authors published three volumes (in Russian) on the subject, which were translated into English during the 1960's [11,12,13]. Meanwhile, Friedman disseminated the ideas of Gelfand and Shilov in his book [10], enhancing them with more recent applications to differential equations, as well as a more complete treatment of the Cauchy problem.

    Several test function spaces that were derived in the framework of the generalized Fourier transformation belong to the family of Gelfand-Shilov K{Mp} spaces, whose theory was developed in [10,12,13] in connection with the Cauchy problem for various partial differential equations, boundary value problems for elliptic equations, and the problem of eigenfunction expansions for several differential operators. Among them are the spaces DK and S, introduced by Schwartz [29]; the space H (also denoted by K1), as developed by Sebastião e Silva [30] and later studied by Hasumi [16], Zieleźny [41], Sznajder and Zieleźny [34], and, somewhat more recently, by de Sousa Pinto [31]; the spaces denoted by Kp (p>1), as developed by Sznajder and Zieleźny [35]; and the spaces Sα,A and WM,a, developed by Gelfand and Shilov themselves [12,13].

    Most of the examples listed above have analogues in the Hankel transformation setting, such as those considered by Zemanian [38,39], Betancor and Marrero [5], Betancor and Rodríguez-Mesa [6,7], Durán [8], van Eijndhoven and Kerkhof [9], Lee [21], Pathak and Sahoo [26], and Pathak and Upadhyay [27]. In order to unify the underlying theory, Marrero introduced [24] and studied [22,23,25] the so-called Hankel-K{Mp} spaces, which were intended to play the same role in the Hankel transformation setting as do the Gelfand-Shilov K{Mp} spaces in the Fourier transformation setting. The study of Hankel-K{Mp} spaces was continued by Arteaga and Marrero [1,2].

    In [10, p. 37], Friedman asserts the following:

    One of the most interesting and important problems in the theory of generalized functions is the problem of finding the structure of generalized functions by expressing them in terms of differential operators acting on functions or on measures.

    Our aim in this paper is threefold: first, we want to briefly review the existing literature on the structure of distributions in spaces of type Hankel-K{Mp}; second, we want to obtain new structural results for these distributions in terms of the Zemanian differential operator Nμ [40, Section 5.3, Equation (3)]; and, third, we want to apply them in the characterization of the bounded subsets and the convergent sequences in the duals of spaces of type Hankel-K{Mp}.

    The paper is organized as follows. In Section 2, the definition and topological properties, along with some examples of Hankel-K{Mp} spaces, are recalled. Section 3 is devoted to reviewing the literature on the structural properties of the dual of a space of type Hankel-K{Mp}. The main results are established in Section 4, where a new description of the topology of a Hankel-K{Mp} space is obtained; then, in Section 5, where such a description is applied to provide new results on the structure, boundedness, and convergence of distributions of type Hankel-K{Mp}.

    Throughout the paper, the standard notation in distribution theory will be used. The letter I will stand for the interval ]0,[ and, unless otherwise stated, μ will be a fixed real parameter not less than 12, while C will represent a suitable positive constant which may vary from line to line.

    Definition 2.1. ([24, Definition 2.1]) Let {Mp}p=0 be a sequence of continuous functions defined on I=]0,[ such that

    1=M0(x)M1(x)M2(x)(xI).

    We say that Kμ is a space of type Hankel-K{Mp}, or just a Hankel-K{Mp} space, provided that Kμ consists of all of the complex-valued functions φC(I) such that

    φμ,,p=max0kpsupxI|Mp(x)(x1D)kxμ12φ(x)|<(pN0).

    Kμ is endowed with the locally convex topology generated by the sequence of norms {μ,,p}p=0. The dual space of Kμ will be denoted by Kμ.

    What follows are examples of test function spaces of type Hankel-K{Mp} arising in connection with the generalized Hankel transformation.

    Example 2.2. Let

    Mp(x)=(1+x2)p(xI, pN0).

    The corresponding Hankel-K{Mp} space is the Zemanian space Hμ [38], [40, Chapter 5].

    Example 2.3. Fix a>0. If the functions {Mp}p=0 are allowed to take on the value , then, with the convention that 0=0, the Zemanian space Bμ,a [39] can be regarded as a Hankel-K{Mp} space upon setting

    Mp(x)={1,0<x<a,xa(pN0).

    Example 2.4. Given α,A>0, define

    Mp(x)=(1+x2)pexp{αeA1α(11p)x1α}(xI, pN0).

    The resulting Hankel-K{Mp} space is the space Hμ,α,A, as introduced by Betancor and Marrero [5].

    Example 2.5. The space χμ, as defined by Betancor and Rodríguez-Mesa [6], is the Hankel-K{Mp} space corresponding to the choice

    Mp(x)=exp(px)(xI, pN0).

    Example 2.6. The space Uμ,M,a developed by Pathak and Upadhyay [27] is also of type Hankel-K{Mp}, as can be seen upon setting

    Mp(x)=exp{M[a(11p)x]}(xI, pN0),

    where a>0,

    M(x)=x0v(ξ)dξ(xI),

    and the function v=v(ξ) is continuous and increasing on [0,[, with v(0)=0 and v()=.

    In the previous examples, the sequence of weights {Mp}p=0 satisfies at least one of the conditions in the following definition.

    Definition 2.7. The sequence {Mp}p=0 is said to satisfy condition () for =O, A, M, N, P, provided that the following hold:

    (O) The limit limx0+Mp(x) (pN0) exists.

    (A) Given r,pN0, there exist sN0 and brp>0 such that

    Mr(x)Mp(x)brpMs(x)(xI).

    (M) Each Mp (pN0) is quasi-monotonic, that is, there exists Cp>0 such that

    Mp(x)CpMp(y)(x,yI, xy).

    (N) For every pN0, there exists rN0, r>p such that the function

    mpr(x)=Mp(x)Mr(x)(xI)

    lies in L1(I) and satisfies

    limxmpr(x)=0.

    (P) Given pN0, there exists rN0, r>p for which

    limxMp(x)Mr(x)=0.

    By imposing appropriate combinations of the above conditions on the weights {Mp}p=0, Hankel-K{Mp} spaces can be endowed with interesting properties as topological vector spaces. It should be remarked that conditions similar to those in Definition 2.7 have been considered by several authors [12,14,15,19,20,32,33,37] in order to develop a suitable theory of the Gelfand-Shilov spaces of type K{Mp}.

    In fact, under adequate assumptions on the weights {Mp}p=0, a space Kμ of type Hankel-K{Mp} can be nuclear [36, Definition Ⅲ.50.1], Schwartz [17, Definition 3.15.1], Montel [17, Definition 3.9.1], and reflexive. We excerpt from [24] some of the results available along these lines, as well as those that reveal the relationship between Kμ and the Zemanian spaces Hμ and Bμ (see Examples 2.2 and 2.3).

    Proposition 2.8. ([24, Proposition 4.1]) The space Kμ is Fréchet. If the sequence {Mp}p=0 satisfies (O) and (P), then Kμ is Schwartz, Montel, and reflexive.

    Proposition 2.9. ([24, Proposition 4.3]) Assume that {Mp}p=0 satisfies conditions (O) or (M). Then, the injection BμKμ is continuous. If, additionally, {Mp}p=0 satisfies (P), then Bμ is dense in Kμ.

    Proposition 2.10. ([24, Proposition 4.4]) Assume that {Mp}p=0 satisfies (M) and (N).

    (i) For every pN0, there exist rN0, r>p, and Cpr>0 such that xCprMr(x) (xI).

    (ii) If {Mp}p=0 satisfies (A) as well, then KμHμ with a continuous embedding.

    Corollary 2.11. ([24, Corollary 4.5]) If {Mp}p=0 satisfies conditions (A), (M), and (N), then BμKμHμ with a continuous embedding. Moreover, Bμ is dense in Kμ and Kμ is dense in Hμ.

    Proposition 2.12. ([24, Proposition 4.6]) Under (A), (M), and (N), the topology of Kμ is compatible with any one of the families of norms {μ,q,p}p=0 (qR, 1q<), where

    φμ,q,p={pk=00|Mp(x)(x1D)kxμ12φ(x)|qdx}1q(φKμ).

    In this case, Kμ is nuclear, Schwartz, Montel, and reflexive.

    Again, by imposing appropiate conditions on the weights {Mp}p=0, the strong dual Kμ,b of a Hankel-K{Mp} space Kμ can be made nuclear, Schwartz, bornological [17, Definition 3.7.1], complete, Montel, and reflexive. Additionally, a wide range of structural results and characterizations of the bounded subsets and convergent sequences in Kμ are made available.

    Proposition 3.1. ([24, Proposition 5.1]) Let Kμ be a Hankel-K{Mp} space with a strong dual Kμ,b.

    (i) If the sequence {Mp}p=0 satisfies (O) and (P), then Kμ,b is complete, bornological, Schwartz, Montel, and reflexive.

    (ii) If {Mp}p=0 satisfies (A),(M), and (N), then Kμ,b is complete, bornological, Schwartz, Montel, reflexive, and nuclear.

    Next, we show that the functionals in the dual Kμ of a space Kμ of type Hankel-K{Mp} can be expressed as distributional derivatives of integrable functions and measures. The highest order of the differential operators that provide such representations is uniform over bounded subsets of Kμ,b. Furthermore, the convergence to zero of a sequence in this space is determined by the convergence to zero, in their respective spaces, of the functions or measures representing the terms of the sequence.

    In Propositions 3.2 and 3.3 below, the results on boundedness and convergence will be stated for the strong topology of Kμ, but, it must be kept in mind that, under the same conditions, they are equally valid if the weak or weak* topologies are instead considered on that space. Indeed, by making appropriate assumptions on the weights, Kμ becomes Montel, and, hence, reflexive (Propositions 2.8 and 2.12). Thus, the weak and weak* topologies of Kμ coincide, whereas the weak* and strong sequential convergence are equivalent on this space [36, Proposition Ⅱ.34.6, Corollary 1]. Furthermore, given that Kμ is Fréchet (Proposition 2.8), it is also barrelled [36, Definition Ⅱ.33.1 and Proposition Ⅱ.33.2, Corollary 1], and, in this class of spaces, the weak* and strong topologies share the same bounded sets [36, Theorem Ⅱ.33.2].

    At this point, we are in a position to state the first result on representation, boundedness, and convergence in the dual of Kμ. To this end, let C(I) denote the space of all of the functions fC[0,[ such that

    limxf(x)=0,

    normed with

    f=supxI|f(x)|.

    Its dual C(I) consists of all of the regular, complex Borel measures σ on [0,[, with the total variation norm |σ|.

    Proposition 3.2. Assume that {Mp}p=0 satisfies (O) and (P).

    (i) [24, Proposition 5.2] A linear functional f belongs to Kμ if, and only if, there exist pN0 and σkC(I) (kN0, 0kp) satisfying

    f=pk=0xμ12(Dx1)k[Mp(x)σk].

    (ii) [24, Proposition 5.4] A set BKμ,b is bounded if, and only if, each fB admits the representation

    f=pk=0xμ12(Dx1)k[Mp(x)σk,f],

    with σk,fC(I) (kN0, 0kp) such that

    pk=00d|σk,f|C,

    where pN0 and C>0 do not depend on fB.

    (iii) [24, Proposition 5.4] A sequence {fj}j=0 converges to zero in Kμ,b if, and only if, each fj admits the representation

    fj=pk=0xμ12(Dx1)k[Mp(x)σk,j](jN0),

    with σk,jC(I) (kN0, 0kp) such that pN0 does not depend on j and

    limjpk=00d|σk,j|=0.

    Proposition 3.3. Assume that {Mp}p=0 satisfies (A), (M), and (N).

    (i) [24, Proposition 5.3] A linear functional f belongs to Kμ if, and only if, for every q, 1<q, there exists pN0 such that f can be written as

    f=pk=0xμ12(Dx1)k[Mp(x)gk(x)],

    with gkLq(I) (kN0, 0kp).

    (ii) [24, Proposition 5.5] A set BKμ,b is bounded if, and only if, given q, 1<q, there exist pN0, C>0, and, for each fB, functions gk,fLq(I) (kN0, 0kp) such that

    f=pk=0xμ12(Dx1)k[Mp(x)gk,f(x)],

    with

    pk=0gk,fqC.

    (iii) [24, Proposition 5.5] A sequence {fj}j=0 converges to zero in Kμ,b if, and only if, for every q, 1<q, there exist pN0 and gk,jLq(I) (kN0, 0kp) such that

    fj=pk=0xμ12(Dx1)k[Mp(x)gk,j(x)](jN0),

    with

    limjpk=0gk,jq=0.

    Starting from Proposition 3.3, and adapting a technique of Kamiński [18], Marrero [25] obtained the results on structure, boundedness, and convergence in Kμ, labeled below as Theorems 3.4, 3.5, and 3.6, respectively. Unlike Proposition 3.3, the aforementioned theorems have the advantage of allowing the elements of the dual to be expressed as the distributional derivative of a single continuous function under the same differential operator.

    Theorem 3.4. ([25, Theorem 2.4]) Assume that {Mp}p=0 satisfies conditions (A), (M), and (N). Then, the following statements are equivalent:

    (i) The functional f lies in Kμ.

    (ii) There exist k,pN0 and a function F, continuous on I, such that

    f=xμ12(Dx1)kF(x) (3.1)

    and

    M1pFLq(I) (3.2)

    for any q, 1q.

    (iii) There exist k,pN0 and a function F, continuous on I and satisfying (3.1), such that (3.2) holds for some q, 1q.

    (iv) There exist k,pN0 and a function F, continuous on I and satisfying (3.1), such that (3.2) holds for q=.

    Now, we shall state a characterization of boundedness in Kμ.

    Theorem 3.5. ([25, Theorem 2.5]) Assume that {Mp}p=0 satisfies conditions (A), (M), and (N). Then, the following four statements are equivalent:

    (i) The set BKμ is (weakly, weakly*, strongly) bounded.

    (ii) There exist k,pN0, C>0, and, for each fB, a function gf, continuous on I, such that

    f=xμ12(Dx1)kgf(x) (3.3)

    and

    M1pgfqC (3.4)

    for any q, 1q.

    (iii) There exist k,pN0, C>0, and, for each fB, a function gf, continuous on I and satisfying (3.3), such that (3.4) holds for some q, 1q.

    (iv) There exist k,pN0, C>0, and, for each fB, a function gf, continuous on I and satisfying (3.3), such that (3.4) holds for q=.

    Finally, convergence in Kμ is described.

    Theorem 3.6. ([25, Theorem 2.6]) Assume that {Mp}p=0 satisfies conditions (A), (M), and (N). Then, the following statements are equivalent:

    (i) The sequence {fj}j=0 converges (weakly, weakly*, strongly) to zero in Kμ.

    (ii) There exist k,pN0 and Fj, continuous on I, such that

    fj=xμ12(Dx1)kFj(x)(jN0) (3.5)

    and

    limjM1pFjq=0 (3.6)

    for any q, 1q.

    (iii) There exist k,pN0 and Fj, continuous on I and satisfying (3.5), such that (3.6) holds for some q, 1q.

    (iv) There exist k,pN0 and Fj, continuous on I and satisfying (3.5), such that (3.6) holds for q=.

    (v) There exist k,pN0, C>0, and functions Fj, continuous on I and satisfying (3.5), such that

    M1pFjC(jN0)

    and limjFj(x)=0 for almost all xI.

    Arteaga and Marrero [1] have shown that the topology of a Hankel-K{Mp} space can be generated by families of norms of type Lq (1q) that involve the Bessel operator.

    Definition 3.7. For 1q<, consider the following families of norms on Kμ:

    |φ|μ,q,r=rk=0{0|Mr(x)xμ12Skμφ(x)|qdx}1q|φ|μ,,r=max0krsupxI|Mr(x)xμ12Skμφ(x)|(φKμ, rN0),

    where

    Sμ=xμ12Dx2μ+1Dxμ12=D24μ214x2

    is the Bessel operator.

    Theorem 3.8. ([1, Proposiciones 2.5 and 2.6]) Under conditions (A), (M), and (N) on the weights {Mp}p=0, any one of the families of norms {||μ,q,r}r=0 (1q) generates the usual topology of Kμ.

    Remark 3.9. Theorem 3.8 widens the class of spaces of type Hankel-K{Mp} and shows how some of them, treated independently in the literature, actually coincide. In fact, for a>0, let

    Mp(x)=exp{M(a[11p]x)}(xI, pN0),

    where MC2[0,[ satisfies that M(0)=M(0)=0, M()=, and M(x)>0 (xI); also, let Kμ be the corresponding Hankel-K{Mp} space. Then, we get that Kμ=xμ+12WeM,a, where WeM,a is the space introduced by van Eijndhoven and Kerkhof [9]. For the space Uqμ,M,a (1q) developed by Pathak and Upadhyay [27], the algebraic and topological identification

    Kμ=Uqμ,M,a=Uμ,M,a=xμ+12WeM,a(1q<)

    follows; see [7] as well.

    More recently, on the basis of Theorem 3.8, and by combining techniques from [24,25], Arteaga and Marrero [2] found new representations for the elements, the bounded subsets, and the convergent sequences in Kμ, this time with the operator Sμ instead of the operator x1D (see Sections 3.2 and 3.3). Their results are summarized below, beginning with the structure of the distributions in Kμ.

    Theorem 3.10. ([2, Theorem 4.1]) Assume that the sequence of weights {Mp}p=0 satisfies conditions (A), (M), and (N). The following statements are equivalent:

    (i) The functional f lies in Kμ.

    (ii) For each q, 1<q, there exist rN0 and fkLq(I) (kN0, 0kr) such that

    f=rk=0Skμ[Mr(x)xμ12fk(x)].

    (iii) There exist k,pN0 and FC(I) such that f=Skμxμ12F(x), with M1pFLq(I) for all q, 1q.

    (iv) There exist k,pN0 and FC(I) such that f=Skμxμ12F(x), with M1pFLq(I) for some q, 1q.

    (v) There exist k,pN0 and FC(I) such that f=Skμxμ12F(x), with M1pFL(I).

    Next, boundedness in Kμ is characterized.

    Theorem 3.11. ([2, Theorem 5.1]) Assume that the sequence of weights {Mp}p=0 satisfies conditions (A), (M), and (N). The following five statements are equivalent:

    (i) The set BKμ is (weakly, weakly*, strongly) bounded.

    (ii) Given q, 1<q, there exist rN0, C>0, and, for each fB, functions gf,iLq(I) (iN0, 0ir) such that

    f=ri=0Siμ[Mr(x)xμ12gf,i(x)],

    with ri=0gf,iqC.

    (iii) There exist k,pN0, C>0, and, for each fB, a function gfC(I) such that f=Skμxμ12gf(x), with M1pgfqC for every q, 1q.

    (iv) There exist k,pN0, C>0, and, for each fB, a function gfC(I) such that f=Skμxμ12gf(x), with M1pgfqC for some q, 1q.

    (v) There exist k,pN0, C>0, and, for each fB, a function gfC(I) such that f=Skμxμ12gf(x), with M1pgfC.

    We close this section with the corresponding characterization of sequential convergence in Kμ.

    Theorem 3.12. ([2, Theorem 6.1]) Under conditions (A), (M), and (N) on the sequence of weights {Mp}p=0, the following statements are equivalent:

    (i) The sequence {fn}n=0 converges (weakly, weakly*, strongly) to zero in Kμ.

    (ii) For each q, 1<q, there exist rN0 and fn,iLq(I) (n,iN0, 0ir) such that

    fn=ri=0Siμ[Mr(x)xμ12fn,i(x)](nN0),

    with limnri=0fn,iq=0.

    (iii) There exist k,pN0 and FnC(I) such that

    fn=Skμxμ12Fn(x)(nN0),

    with limnM1pFnq=0 for all q, 1q.

    (iv) There exist k,pN0 and FnC(I) such that

    fn=Skμxμ12Fn(x)(nN0),

    with limnM1pFnq=0 for some q, 1q.

    (v) There exist k,pN0 and FnC(I) such that

    fn=Skμxμ12Fn(x)(nN0),

    with limnM1pFn=0.

    (vi) There exist k,pN0, C>0, and FnC(I) such that

    fn=Skμxμ12Fn(x),

    with M1pFnC (nN0) and limnFn(x)=0 for almost all xI.

    First, we shall show that, for μ12 and every p, 1p, the families of norms {ρμp,r}r=0 generate the usual topology of a space Kμ of type Hankel-K{Mp}, where, for rN0 and φKμ,

    ρμ,r(φ)=max0krsupxI|Mr(x)Tμ,kφ(x)|,ρμp,r(φ)=max0kr{0|Mr(x)Tμ,kφ(x)|pdx}1p(1p<). (4.1)

    Here,

    Tμ,k=Nμ+k1Nμ(kN), (4.2)

    where Nμ=xμ+12Dxμ12 denotes the Zemanian operator [40, pp. 135ff] and Tμ,0 is the identity operator. The differential operators in (4.2) are interesting because of their symmetric behavior in the presence of the Hankel transformation; in fact, for an appropriate order of this transformation, it exchanges the order of the powers of the variable and the order of the differential operator defined in (4.2), an extremely useful operational rule that is very similar to its Fourier counterpart [40, Proof of Theorem 5.4-1].

    This new description for the topology of Kμ was motivated by [4, Theorem 3.3], where a similar result was established for the Zemanian space Hμ (Example 2.2). The validity of this result requires assuming conditions (O), (A), (M), and (N) apply to the sequence of weights {Mp}p=0 (Definition 2.7), which is a hypothesis that will be maintained throughout the entire section, although, for the sake of simplicity, it will not be made explicit on all occasions.

    The proof of this first result imitates that of [4, Theorem 3.3]. Roughly speaking, it consists of introducing a new space of test functions, which we will denote by Sμ, and whose definition is formally analogous to that of Kμ, but with the operator Tμ,k in place of the operator (x1D)kxμ12 (kN0); this is followed by proving, with the aid of the open mapping theorem [28, Corollary 2.12], that, actually, Sμ=Kμ (Theorem 4.4). Once this is done, it is not difficult to infer that the families of norms defined in (4.1) are equivalent over Sμ (Proposition 4.6). At this point, we will apply techniques that are analogous to those used in the proof of the results in Section 3 in order to find representations of the elements, the (weakly, weakly*, strongly) bounded subsets, and the (weakly, weakly*, strongly) convergent sequences in the dual space Sμ=Kμ, this time as distributional derivatives induced by the adjoint of the operator Tμ,k (kN0). Thus, Theorems 5.4–5.7 below generalize and improve, in a sense that will be specified in due course, their analogues for the space Hμ in [4].

    We must emphasize that, although the ideas presented in this section are not entirely new, the results obtained in the general context of Hankel-K{Mp} spaces have not appeared previously in the literature.

    Given μR, denote by Sμ the vector space of all smooth, complex-valued functions φ=φ(x) defined on I=]0,[ such that ωμp,k(φ)<, where

    ωμp,k(φ)=supxI|Mp(x)Tμ,kφ(x)|(p,kN0)

    and Tμ,k (kN0) is as defined above.

    A direct computation shows that, for kN0,

    Tμ,k=xk+μ+12(x1D)kxμ12=bμk,0xk+bμk,1xk+1D++bμk,kDk, (4.3)

    where the coefficients bμk,j (0jk) are appropriate constants, with bμk,k=1.

    The family Ω={ωμp,k}p,k=0 is a countable family of seminorms. This family is separating, because {ωμp,0}p=0 are norms. Consequently, Ω makes Sμ into a countably multinormed space.

    A family of norms P={ρμ,r}r=0 that is equivalent to Ω, with the property that ρμ,rρμ,s (r,sN0), is obtained by setting

    ρμ,r=max0krωμr,k(rN0).

    Proposition 4.1. The space Sμ (μR) is Fréchet.

    Proof. Let {φn}n=0 be a Cauchy sequence in Sμ; we want to prove that it converges in Sμ.

    By (4.3), we have

    Dk=Tμ,k(bμk,0xk++bμk,k1x1Dk1). (4.4)

    Proceeding by induction on k, we find that, for each compact KI, there exist constants cμk,j (0jk) such that

    supxK|Dkφ(x)|cμk,0ωμ0,0(φ)+cμk,1ωμ0,1(φ)++cμk,kωμ0,k(φ)(φSμ).

    Since ωμ0,k(φnφm)n,m0, the sequence {Dkφn}n=0 is uniformly Cauchy on compact subsets of I for all kN0. Thus, there exists φC(I) such that

    Dkφn(x)nDkφ(x)(kN0)

    uniformly over compact subsets of I. This φ is the limit of {φn}n=0 in Sμ. Indeed, for any p,kN0 and every ε>0, there exists Nμp,k=Nμp,k(ε)N0 such that

    |Mp(x)Tμ,k(φnφm)(x)|<ε(xI, n,mNμp,k).

    Taking the limit as n, it follows that

    ωμp,k(φφm)ε(mNμp,k).

    On the other hand, there exists Bμp,k>0, independent of m, such that

    ωμp,k(φm)Bμp,k(m,p,kN0).

    Thus,

    ωμp,k(φ)Bμp,k+ε(p,kN0).

    This shows that φSμ and {φn}n=0 converges to φ in Sμ, as asserted.

    Proposition 4.2. For μ12, the inclusion KμSμ holds. Moreover, the embedding KμSμ is continuous.

    Proof. Let pN0. Proposition 2.10(i), along with condition (A) on the weights, yields r,sN0, s>r>p+μ+12, and C>0 such that

    |Mp(x)Tμ,kφ(x)|=|Mp(x)xk+μ+12(x1D)kxμ12φ(x)|C|Mp(x)Mr(x)(x1D)kxμ12φ(x)|C|Ms(x)(x1D)kxμ12φ(x)|(xI, kN0, 0kp)

    whenever φKμ. According to Definition 2.1, this means that

    ρμ,p(φ)Cφμ,,s<(φKμ),

    which completes the proof.

    Proposition 4.3. Suppose that φSμ. Then, φKμ if, and only if, the following limits exist:

    limx0+(x1D)kxμ12φ(x)(kN0). (4.5)

    In other words,

    Kμ={φSμ:there existlimx0+(x1D)kxμ12φ(x) (kN0)}={φSμ:(x1D)kxμ12φ(x)=O(1)asx0+ (kN0)}. (4.6)

    Proof. Note that the existence of a limit is a stronger condition than boundedness near the origin; thus, the first set on the right-hand side of (4.6) is contained in the second set.

    Under (A), (M), and (N), we have that KμHμ (Proposition 2.10(ii)). Hence, if φKμ, then φSμ (by Proposition 4.2) and the limits given by (4.5) exist [40, Lemma 5.2-1].

    Finally, suppose that φSμ and (x1D)kxμ12φ(x) is bounded near zero for all kN0. Since (O) holds, Mp(x)(x1D)kxμ12φ(x) is also bounded near zero for any p,kN0. On the other hand,

    |Mp(x)(x1D)kxμ12φ(x)|=|Mp(x)xkμ12Tμ,kφ(x)||Mp(x)Tμ,kφ(x)|ωμp,k(φ)<(1x<, p,kN0).

    This proves that φKμ.

    Theorem 4.4. For μ12, the inclusion SμHμ holds with a continuous embedding. Consequently, Sμ=Kμ both algebraically and topologically.

    Proof. To show that SμHμ and the inclusion map SμHμ is continuous, note that, given pN0, Proposition 2.10(i) and condition (A) on the weights yield rN0, r>p, and C>0, for which xpCMr(x) (xI). Therefore,

    supxI|xpTμ,kφ(x)|CsupxI|Mr(x)Tμ,kφ(x)|Cρμ,r(φ)<(φSμ, kN0, 0kp),

    and it suffices to invoke [4, Theorem 3.3]. In particular, the limits given by (4.5) exist whenever φSμ [40, Lemma 5.2-1], and, from Proposition 4.3, it follows that Sμ=Kμ.

    This equality is also topological, as it can be deduced from Propositions 2.8, 4.1, and 4.2 by applying the open mapping theorem [28, Corollary 2.12].

    Once it is proved that SμHμ, Proposition 4.5 is an immediate consequence of [4, Theorem 3.3 and Proposition 2.3]. Just for the sake of completeness, a direct proof is included.

    Proposition 4.5. Let μ12. Every φSμ is bounded. For any kN0, Dkφ is rapidly decreasing at infinity; in particular, Sμ is a dense proper subspace of L1(I).

    Proof. Every φSμ is bounded because ωk0,0(φ)<.

    To prove that Dkφ (kN0) is rapidly decreasing at infinity, we proceed by induction on k, bearing in mind (4.4) and the fact that, given mN0, Proposition 2.10(i), along with condition (A), yields rN0, r>m, and C>0 such that xmCMr(x) (xI). The inductive scheme is as follows.

    k=0: xmφ(x) (mN0) is bounded for xI because

    |xmφ(x)|C|Mr(x)φ(x)|=Cωμr,0(φ)<.

    Therefore, φ(x) is rapidly decreasing at infinity.

    k=1: xmDφ(x) (mN0) is bounded for x1 because, by (4.4), we have

    xmDφ(x)=xm Tμ,1φ(x)ωμr,1(φ)<bμ1,0xm1φ(x)m1: step 0m=0: φ bounded.

    Therefore, Dφ(x) is rapidly decreasing at infinity.

    k=2: xmD2φ(x) (mN0) is bounded for x1 because, by (4.4), we have

    xmD2φ(x)=xm Tμ,2φ(x)ωμr,2(φ)<bμ2.0xm2φ(x)m2: step 0m=0,1: φ boundedb2,1xm1Dφ(x)m1: step 1m=0: Dφ bounded.

    Therefore, D2φ(x) is rapidly decreasing at infinity.

    Assuming that the statement

    xmDnφ(x) (mN0) is bounded for x1

    holds true for all nN0 with 0nk, we prove it for k+1.

    k+1: xmDk+1φ(x) (mN0) is bounded for x1 because, by (4.4), we have

    xmDk+1φ(x)=xm Tμ,k+1φ(x)ωμr,k+1(φ)<bμk+1,0xmk1φ(x)mk+1: ind. hyp.0mk: φ boundedbμk+1,1xmkDφ(x)mk: ind. hyp.0mk1: Dφ boundedbμk+1,kxm1Dkφ(x)m1: ind. hyp.m=0: Dkφ bounded.

    Therefore, Dk+1φ(x) is rapidly decreasing at infinity and the induction is complete.

    Finally, it is clear that

    D(I)SμL1(I),

    which implies that Sμ is dense in L1(I).

    Proposition 4.6. For μ12 and any p, 1p, the families of norms {ρμp,r}r=0 generate the topology of Sμ, where, for rN0 and φSμ,

    ρμ,r(φ)=max0krsupxI|Mr(x)Tμ,kφ(x)|,ρμp,r(φ)=max0kr{0|Mr(x)Tμ,kφ(x)|pdx}1p(1p<).

    Proof. Fix 1<p< and φSμ. Given rN0, condition (N) yields sN0, s>r such that

    0Mr(x)Ms(x)dx<. (4.7)

    Since

    0Mr(x)Ms(x)1(xI), (4.8)

    we may write the following:

    ρμp,r(φ)=max0kr{0|Mr(x)Tμ,kφ(x)|pdx}1p=max0kr{0|Ms(x)Tμ,kφ(x)|p(Mr(x)Ms(x))pdx}1pmax0kssupxI|Ms(x)Tμ,kφ(x)|{0Mr(x)Ms(x)dx}1p={0Mr(x)Ms(x)dx}1pρμ,s(φ).

    A new application of (4.7) and (4.8), in combination with Hölder's inequality, leads to

    ρμ1,r(φ)=max0kr0|Mr(x)Tμ,kφ(x)|dx=max0kr0|Ms(x)Tμ,kφ(x)|Mr(x)Ms(x)dxmax0ks{0|Ms(x)Tμ,kφ(x)|pdx}1p{0(Mr(x)Ms(x))qdx}1q={0Mr(x)Ms(x)dx}1qρμp,s(φ).

    Here, q=p(p1)1 denotes the conjugate exponent of p.

    Finally, given k0N0 and φSμ, the function (x1D)kxμ12φ(x) is rapidly decreasing at infinity because SμHμ (Theorem 4.4). Hence, for φSμ and r,kN0 with 0kr, by using (M), we find that

    |Mr(x)Tμ,kφ(x)|=|Mr(x) xk+μ+12(x1D)kxμ12φ(x)|=|Mr(x) xk+μ+12xt(t1D)k+1tμ12φ(t)dt|Cx|Mr(t) t(k+1)+μ+12(t1D)k+1tμ12φ(t)|dtC0|Mr(t)Tμ,k+1φ(t)|dt(xI).

    Consequently,

    ρμ,r(φ)Cρμ1,r+1(φ).

    The proof is thus complete.

    Now that Proposition 4.6 has been proved, we intend to apply techniques similar to those employed in the proof of the results in Section 3 in order to characterize the elements, the (weakly, weakly*, strongly) bounded subsets, and the (weakly, weakly*, strongly) convergent sequences in the dual space Kμ=Sμ (μ12), this time as distributional derivatives induced by the adjoint of the operator Tμ,k (kN0). The main results correspond to Theorems 5.4–5.5, 5.6, and 5.7, respectively. Prior to deriving these results, three auxiliary lemmas must be established.

    Lemma 5.1. Let μ12, and let FC(I) be such that there exists pN0 with M1pFLq(I) (1q). Then, for each kN0, there exist pkN0, pkp, and a function FkC(I) satisfying

    xμ12(Dx1)kxk+μ+12Fk(x)=F(x)(xI) (5.1)

    and

    M1pkFkLq(I)(1q). (5.2)

    Proof. We proceed by induction on k. The result is obvious if k=0. Suppose that, given kN0, k1, there exist pkN0, pkp, and a function FkC(I) satisfying (5.1) and (5.2). Use (N) to find n,tN0, n>t>pk, such that

    0Mpk(x)Mt(x)dx<, (5.3)
    0Mt(x)Mn(x)dx<. (5.4)

    The inductive hypotheses, along with (M) and (5.3), allow us to write the following:

    1Mt(x)x0|Fk(ξ)|dξC0|Fk(ξ)Mt(ξ)|dξ=C0|Fk(ξ)Mpk(ξ)|Mpk(ξ)Mt(ξ)dξCsupξI|Fk(ξ)Mpk(ξ)|0Mpk(ξ)Mt(ξ)dξ=CsupξI|Fk(ξ)Mpk(ξ)|. (5.5)

    The function

    ˜Fk(x)=xkμ12x0Fk(ξ)ξk+μ+12dξ(xI)

    is continuous, and, by (5.1), we have

    xμ12(Dx1)k+1x(k+1)+μ+12˜Fk(x)=xμ12(Dx1)kxk+μ+12Fk(x)=F(x)(xI).

    Furthermore, using (5.5), it follows that

    |˜Fk(x)Mt(x)|=1Mt(x)|xkμ12x0Fk(ξ)ξk+μ+12dξ|1Mt(x)x0|Fk(ξ)|dξCsupξI|Fk(ξ)Mpk(ξ)|(xI).

    Because n>t, given (5.2), it clearly follows that

    supxI|˜Fk(x)Mn(x)|supxI|˜Fk(x)Mt(x)|CsupξI|Fk(ξ)Mpk(ξ)|<.

    On the other hand, if 1q<, then, given (5.2), and with the aid of (5.4), we obtain

    0|˜Fk(x)Mn(x)|qdx=0|˜Fk(x)Mt(x)|q(Mt(x)Mn(x))qdxCsupξI|Fk(ξ)Mpk(ξ)|q0Mt(x)Mn(x)dx=CsupξI|Fk(ξ)Mpk(ξ)|q<.

    To complete the induction, it suffices to take pk+1=n and Fk+1=˜Fk.

    Lemma 5.2. Let μ12, and let M be a family of functions in C(I) with the property that

    supFMM1pFqA(1q)

    for certain pN0 and A>0. Then, given kN0, there exist pkN0, pkp, Ck>0, and, for each FM, a function gk,FC(I) such that

    xμ12(Dx1)kxk+μ+12gk,F(x)=F(x)(xI) (5.6)

    and

    supFMM1pkgk,FqCk(1q). (5.7)

    Proof. The result holds trivially for k=0. Proceeding by induction, fix kN0, k1. Let pkN0, pkp, and Ck>0, and, for each FM, let gk,FC(I) satisfy (5.6) and (5.7). As in the proof of Lemma 5.1, for every FM, we construct a function ˜gk,FC(I) such that

    xμ12(Dx1)k+1x(k+1)+μ+12˜gk,F(x)=F(x)(xI),supxI|˜gk,F(x)Mn(x)|CsupξI|gk,F(ξ)Mpk(ξ)|,

    and

    0|˜gk,F(x)Mn(x)|qdxCsupξI|gk,F(ξ)Mpk(ξ)|q(1q<)

    for some nN0, n>pk, where the positive constant C does not depend on F. To complete the proof, it suffices to choose pk+1=n and gk+1,F=˜gk,F, as well as to take into account the inductive hypotheses.

    The next result can be analogously established.

    Lemma 5.3. Let μ12, and let {Fj}j=0 be a sequence of functions in C(I) for which there exists pN0 with

    limjM1pFjq=0(1q).

    Then, for each kN0, there exist pkN0, pkp, and Fk,jC(I) (jN0) such that

    xμ12(Dx1)kxk+μ+12Fk,j(x)=Fj(x)(xI, jN0)

    and

    limjM1pkFk,jq=0(1q).

    At this point, we are in a position to give several representations of the elements in the dual space of Sμ in terms of the adjoint of the operator Tμ,k (kN0). This will be done in Theorems 5.4 and 5.5. The proof of Theorem 5.4 uses a fairly standard method, which can be traced back to [36] and was already used in [3,4,24], but we include it for the sake of completeness.

    Theorem 5.4. Let μ12. A linear functional f belongs to Sμ if, and only if, for every q, 1<q, there exist rN0 and measurable functions gk, with M1rgkLq(I) (kN0, 0kr), such that

    f=xμ12rk=0(Dx1)kxk+μ+12 gk(x). (5.8)

    Proof. Fix q, 1<q, and let p, 1p<, be the conjugate exponent of q.

    If fSμ, then, by Proposition 4.6, there exist rN0 and C>0 such that

    |f,φ|Cρμp,r(φ)(φSμ). (5.9)

    Let us denote by Γp the direct sum of r+1 copies of Lp(I), normed by

    (fk)0krp=max0krfkp, (5.10)

    and by Ξq the direct sum of r+1 copies of Lq(I), normed by

    (fk)0krq=rk=0fkq.

    Consider the following injective map:

    F:SμΓpφF(φ)=(Mr(x)Tμ,kφ(x))0kr,

    and define on F(Sμ)Γp the linear functional L by means of the formula

    L,F(φ)=f,φ.

    In view of (5.9) and (5.10), L is continuous, and its norm is, at most, C. We continue to denote by L a norm-preserving Hahn-Banach extension of this functional up to Γp. The Riesz representation (fk)0krΞq of L over Γp satisfies

    L=rk=0fkqC (5.11)

    and

    f,φ=L,F(φ)=rk=00fk(x)Mr(x)xk+μ+12(x1D)kxμ12φ(x)dx=rk=0fk(x),Mr(x)xk+μ+12(x1D)kxμ12φ(x)=rk=0xμ12(Dx1)kxk+μ+12Mr(x)fk(x),φ=xμ12rk=0(Dx1)kxk+μ+12Mr(x)fk(x),φ(φSμ).

    Setting gk=(1)kfkMr (kN0, 0kr), we obtain (5.8).

    Conversely, if f is given by (5.8), then Hölder's inequality implies that

    |f,φ|rk=00|M1r(x)gk(x)||Mr(x)xk+μ+12(x1D)kxμ12φ(x)|dxρμp,r(φ)rk=0M1rgkq(φSμ),

    which yields that fSμ.

    Theorem 5.5. For μ12, the following statements are equivalent:

    (i) The functional f lies in Sμ.

    (ii) There exist k,pN0 and a function FC(I) such that

    f=xμ12(Dx1)kxk+μ+12F(x), (5.12)

    with

    M1pFLq(I) (5.13)

    for all q, 1q.

    (iii) There exist k,pN0 and a function FC(I) satisfying (5.12) such that (5.13) holds for some q, 1q.

    (iv) There exist k,pN0 and a function FC(I) satisfying (5.12) such that (5.13) holds for q=.

    Proof. (i)(ii) If fSμ, Theorem 5.4 yields that pN0 and measurable functions Gi (iN0, 0ip) exist such that

    f=xμ12pi=0(Dx1)ixi+μ+12Gi(x) (5.14)

    and

    M1pGiL(I)(iN0, 0ip). (5.15)

    Apply (N) to get n,tN0, n>t>p, for which

    0Mp(x)Mt(x)dx<, (5.16)
    0Mt(x)Mn(x)dx<. (5.17)

    Fix iN0, 0ip. By (M) and (5.16), we obtain

    1Mt(x)x0|Gi(ξ)|dξC0|Gi(ξ)Mp(ξ)|Mp(ξ)Mt(ξ)dξCsupξI|Gi(ξ)Mp(ξ)|0Mp(ξ)Mt(ξ)dξ=CsupξI|Gi(ξ)Mp(ξ)|(xI). (5.18)

    The function

    ˜Gi(x)=xiμ12x0Gi(ξ)ξi+μ+12dξ(xI)

    is continuous and satisfies

    xμ12(Dx1)i+1x(i+1)+μ+12˜Gi(x)=xμ12(Dx1)ixi+μ+12Gi(x)(xI).

    Using (5.18), we may write

    |˜Gi(x)|Mt(x)=1Mt(x)|xiμ12x0Gi(ξ)ξi+μ+12dξ|1Mt(x)x0|Gi(ξ)|dξCsupξI|Gi(ξ)Mp(ξ)|(xI). (5.19)

    As n>t,

    supxI|˜Gi(x)Mn(x)|supxI|˜Gi(x)Mt(x)|CsupξI|Gi(ξ)Mp(ξ)|<.

    Furthermore, taking into account (5.19) and (5.17), we find that

    0|˜Gi(x)Mn(x)|qdx=0|˜Gi(x)Mt(x)|q(Mt(x)Mn(x))qdxCsupξI|Gi(ξ)Mp(ξ)|q0(Mt(x)Mn(x))qdxCsupξI|Gi(ξ)Mp(ξ)|q0Mt(x)Mn(x)dx=CsupξI|Gi(ξ)Mp(ξ)|q<

    whenever 1q<. Now, applying Lemma 5.1 with μ+i+112 instead of μ12, we obtain a function FiC(I) and a non-negative integer sin such that

    x(μ+i+1)12(Dx1)pix(pi)+(μ+i+1)+12Fi(x)=˜Gi(x)(xI),

    with

    M1siFiLq(I)(1q).

    It follows that

    xμ12(Dx1)ixi+μ+12Gi(x)=xμ12(Dx1)i+1x(i+1)+μ+12˜Gi(x)=xμ12(Dx1)i+1x(μ+i+1)+12[x(μ+i+1)12(Dx1)pix(pi)+(μ+i+1)+12Fi(x)]=xμ12(Dx1)i+1(Dx1)pix(p+1)+μ+12Fi(x)=xμ12(Dx1)p+1x(p+1)+μ+12Fi(x)(xI).

    Set

    F=pi=0Fi,m=max0ipsi.

    Then, FC(I). Using (5.14) and (5.15), we have

    f=xμ12pi=0(Dx1)ixi+μ+12Gi(x)=xμ12(Dx1)p+1x(p+1)+μ+12pi=0Fi(x)=xμ12(Dx1)p+1x(p+1)+μ+12F(x)

    and

    M1mFLq(I)(1q).

    It is thus proved that (i) implies (ii).

    (ii)(iii) This is obvious.

    (iii)(iv) Suppose that there exist k,pN0 and a function FC(I) satisfying (5.12) and (5.13) for some q, 1q, and apply (N) to find tN0, t>p, such that

    0Mp(x)Mt(x)dx<. (5.20)

    The function

    ˜F(x)=xkμ12x0F(ξ)ξk+μ+12dξ(xI)

    is continuous, with

    xμ12(Dx1)k+1x(k+1)+μ+12˜F(x)=xμ12(Dx1)kxk+μ+12F(x)=f.

    By choosing n>t and using (M), we obtain

    supxI|˜F(x)Mn(x)|CsupxI|xkμ12x0F(ξ)Mp(ξ)Mp(ξ)Mt(ξ)Mt(ξ)Mn(ξ)ξk+μ+12dξ|C0|F(ξ)Mp(ξ)|Mp(ξ)Mt(ξ)dξ.

    If q=1, (5.13) and the fact that t>p imply that

    supxI|˜F(x)Mn(x)|C0|F(ξ)Mp(ξ)|dξ<.

    If q=, conditions (5.13) and (5.20) yield

    supxI|˜F(x)Mn(x)|CsupξI|F(ξ)Mp(ξ)|0Mp(ξ)Mt(ξ)dξ<.

    Finally, if 1<q<, then (5.13), Hölder's inequality, and (5.20) lead to

    supxI|˜F(x)Mn(x)|C{0|F(ξ)Mp(ξ)|qdξ}1q{0(Mp(ξ)Mt(ξ))qdξ}1qC{0|F(ξ)Mp(ξ)|qdξ}1q{0Mp(ξ)Mt(ξ)dξ}1q<.

    Here, q denotes the conjugate exponent of q. Therefore, (5.12) and (5.13) hold with k+1 instead of k, ˜F instead of F, and n instead of p, when q=. This establishes (iv).

    (iv) (i) To complete the proof, assume that there exist k,pN0 and FC(I) such that

    f=xμ12(Dx1)kxk+μ+12F(x),

    with

    M1pFL(I).

    This representation of f guarantees that fSμ, as it can be deduced from the following estimate:

    |f,φ|=|(1)k0F(x)xk+μ+12(x1D)kxμ12φ(x)dx|supxI|F(x)Mp(x)| supxI|Mr(x)xk+μ+12(x1D)kxμ12φ(x)| 0Mp(x)Mr(x)dxCωμr,k(φ)(φSμ), (5.21)

    where r>p is chosen according to (N). Thus, (iv) implies (i), and we are done.

    Next, we shall characterize boundedness in Sμ. Recall that, under (A), (M), and (N), Sμ is reflexive (Theorem 4.4 and Proposition 2.12), which means that the weak and weak* topologies of Sμ coincide. Furthermore, because Sμ is a Fréchet space (Proposition 4.1), it is barrelled [36, Definition Ⅱ.33.1 and Proposition Ⅱ.33.2, Corollary 1], and, in this class of spaces, the weak* and strong topologies share the same bounded sets [36, Theorem Ⅱ.33.2].

    Theorem 5.6. For μ12, the following five statements are equivalent:

    (i) The set BSμ is (weakly, weakly*, strongly) bounded.

    (ii) Given q, 1<q, there exist rN0, C>0, and, for each fB, measurable functions gk,f (kN0, 0kr) such that

    f=xμ12rk=0(Dx1)kxk+μ+12gk,f(x),

    with

    rk=0M1rgk,fqC.

    (iii) There exist k,pN0, C>0, and, for each fB, a function gfC(I) such that

    f=xμ12(Dx1)kxk+μ+12gf(x), (5.22)

    with

    M1pgfqC (5.23)

    for any q, 1q.

    (iv) There exist k,pN0, C>0, and, for each fB, a function gfC(I) satisfying (5.22) such that (5.23) holds for some q, 1q.

    (v) There exist k,pN0, C>0, and, for each fB, a function gfC(I) satisfying (5.22) such that (5.23) holds for q=.

    Proof. Since Sμ is barrelled, a subset B of Sμ is (weakly, weakly*, strongly) bounded if, and only if, it is equicontinuous [36, Proposition Ⅱ.33.1], which means that (5.9) holds for rN0 and C>0 independent of fB. Now, (ii) can be obtained by using the same argument as in the proof of Theorem 5.4, taking into account condition (5.11) on the norms of the representing functions.

    If (ii) holds, then, in particular, there exist pN0, A>0, and, for each fB, measurable functions Gi,f (iN0, 0ip) such that

    f=xμ12pi=0(Dx1)ixi+μ+12Gi,f(x),

    with

    pi=0M1pGi,fA.

    Fix iN0, 0ip. The argument in the proof that (i) implies (ii) in Theorem 5.5 yields that nN0, n>p, ˜A>0, and ˜Gi,fC(I) exist such that

    xμ12(Dx1)i+1x(i+1)+μ+12˜Gi,f(x)=xμ12(Dx1)ixi+μ+12Gi,f(x)(xI),
    supxI|˜Gi,f(x)Mn(x)|CsupξI|Gi,f(ξ)Mp(ξ)|˜A,

    and

    0|˜Gi,f(x)Mn(x)|qdxCsupξI|Gi,f(ξ)Mp(ξ)|q˜Aq(1q<),

    where ˜A is independent of fB. According to Lemma 5.2, applied with μ+i+112 instead of μ12, there exist siN0, sin, Ci>0, and, for each fB, a function Fi,fC(I) such that

    x(μ+i+1)12(Dx1)pix(pi)+(μ+i+1)+12Fi,f(x)=˜Gi,f(x)(xI),

    with

    M1siFi,fqCi(1q).

    Therefore,

    xμ12(Dx1)ixi+μ+12Gi,f(x)=xμ12(Dx1)p+1x(p+1)+μ+12Fi,f(x)(xI).

    Setting

    gf=pi=0Fi,f,m=max0ipsi,C=pi=0Ci,

    we find that gfC(I),

    f=xμ12(Dx1)p+1x(p+1)+μ+12gf(x),

    and

    M1mgfqC(1q),

    where mN0 and C>0 do not depend on fB. Thus, (ii) implies (iii).

    It is apparent that (iii) implies (iv).

    To prove that (iv) implies (v), assume that there exist k,pN0, A>0, and, for each fB, a function gfC(I) satisfying

    f=xμ12(Dx1)kxk+μ+12gf(x)

    with

    M1pgfqA

    for some q, 1q. The argument in the proof that (iii) implies (iv) in Theorem 5.5 allows us to find nN0, n>p, ˜A>0, and, for each fB, a function ˜gfC(I) such that

    f=xμ12(Dx1)k+1x(k+1)+μ+12˜gf(x),

    with

    supxI|˜gf(x)Mn(x)|C{0|gf(ξ)Mp(ξ)|qdξ}1q˜A,

    if 1q<, or

    supxI|˜gf(x)Mn(x)|CsupξI|gf(ξ)Mp(ξ)|˜A,

    if q=. This establishes (v).

    Finally, (v) and (5.21), with gf (fB) instead of F, yield (i).

    Next, we shall describe sequential convergence in the dual of Sμ. Under (A), (M), and (N), Sμ is Montel and, hence, reflexive (Theorem 4.4 and Proposition 2.12). Thus, the weak and weak* topologies of Sμ coincide, and weak* and strong sequential convergence are equivalent in this space [36, Proposition Ⅱ.34.6, Corollary 1].

    Theorem 5.7. For μ12, the following statements are equivalent:

    (i) The sequence {fj}j=0 converges (weakly, weakly*, strongly) to zero in Sμ.

    (ii) For each q, 1<q, there exist pN0 and measurable functions gk,j (kN0, 0kp) such that

    fj=xμ12pk=0(Dx1)kxk+μ+12gk,j(x)(jN0),

    with limjpk=0M1pgk,jq=0.

    (iii) There exist k,pN0 and FjC(I) (jN0) such that

    fj=xμ12(Dx1)kxk+μ+12Fj(x)(jN0) (5.24)

    and

    limjM1pFjq=0 (5.25)

    for any q, 1q.

    (iv) There exist k,pN0 and FjC(I) (jN0) satisfying (5.24) and (5.25) for some q, 1q.

    (v) There exist k,pN0 and FjC(I) (jN0) satisfying (5.24) and (5.25) for q=.

    (vi) There exist k,pN0, C>0, and FjC(I) (jN0) satisfying (5.24), with

    M1pFjC(jN0)

    and limjFj(x)=0 for almost all xI.

    Proof. To show that (i) implies (ii), suppose that {fj}j=0 converges to zero in Sμ. Then [24, Equation (5.7)], there exist pN0 and positive constants {Cj}j=0 such that

    |fj,φ|Cjφμ,,p(φSμ, jN0)

    and limjCj=0. On the other hand (Proposition 4.6), given q, 1<q, there exist C>0 and rN0 such that

    φμ,,pCρμq,r(φ)(φSμ).

    It follows that

    |fj,φ|Cjρμq,r(φ)(φSμ, jN0), (5.26)

    where the redefined sequence {Cj}j=0 still satisfies

    limjCj=0. (5.27)

    Now, it suffices to provide the same argument as in the proof of Theorem 5.4, using (5.26) instead of (5.9) and keeping in mind (5.27) in relation to (5.11).

    Parts (iv) and (vi) follow trivially from (iii) and (v), respectively.

    The arguments in the proof of the corresponding results for Theorems 5.5 and 5.6, in combination with Lemma 5.3 instead of Lemmas 5.1 and 5.2, allow us to establish that (ii) implies (iii) and (iv) implies (v); we omit the details.

    Finally, assuming that (vi) holds, to obtain (i), it suffices to choose nN0, n>p, according to (N) and apply Lebesgue's dominated convergence theorem to the following integrals:

    fj,φ=(1)k0Fj(x)Mp(x)Mp(x)Mn(x)Mn(x)Tμ,kφ(x)dx(φSμ, jN0).

    This completes the proof.

    Theorems 5.4–5.7 characterize the structure, boundedness, and convergence in the dual of a wide range of spaces that arise in the context of the Hankel transformation (see Sections 2 and 3.4). For illustrative purposes, in Corollary 5.8, some of those results have been specialized for the Zemanian space Hμ, which, as already mentioned (Example 2.2), is none other than the Hankel-K{Mp} space defined by the sequence of weights {(1+x2)p}p=0.

    The application of Theorem 5.4 to Hμ is as stated in [4, Theorem 5.5]. However, the representation of the elements in Hμ obtained as a consequence of Theorem 5.5 is an improvement upon the result of [4] in three ways. First, it involves the Tμ,k-distributional derivative (kN0) of a single continuous function F. Second, the representing function F is not dependent on the exponent q; in fact, the products of F and the inverses of the weights belong to all classes Lq(I), 1q. Finally, as we have just highlighted, and unlike what occurs in Theorem 5.4, q=1 is included in the range of exponents for which the representation is valid.

    Remarks analogous to the preceding ones can be made regarding Theorem 5.7; particularly, the application of part (ii) to the space Hμ is as stated in [4, Theorem 5.4], while parts (iii) to (v) improve upon this result along the lines described above.

    Reference [4] contains no analogue of Theorem 5.6.

    Corollary 5.8. Let Hμ be the dual space of Hμ. Then, the following holds:

    (i) A functional f belongs to Hμ if, and only if, there exist k,pN0 and a function FC(I) such that

    f=xμ12(Dx1)kxk+μ+12F(x),

    with

    (1+x2)pF(x)Lq(I)(1q).

    (ii) A set BHμ is (weakly, weakly*, strongly) bounded if, and only if, there exist k,pN0, C>0, and, for every fB, a function gfC(I) such that

    f=xμ12(Dx1)kxk+μ+12gf(x),

    with

    (1+x2)pgf(x)qC(1q).

    (iii) A sequence {fj}j=0 converges (weakly, weakly*, strongly) to zero in Hμ if, and only if, there exist k,pN0 and gjC(I) such that

    fj=xμ12(Dx1)kxk+μ+12gj(x)(jN0),

    with

    limj(1+x2)pgj(x)q=0(1q).

    In Friedman's opinion [10, p. 37], one of the most interesting and important problems in the theory of generalized functions or distributions is to find their structure, that is, to express them in terms of differential operators acting on functions or on measures. This paper provides novel results in this direction for the spaces of type Hankel-K{Mp}. Namely, here it was shown that, for μ12, and under certain conditions on the weights {Mp}p=0, the elements, the (weakly, weakly*, strongly) bounded subsets, and the (weakly, weakly*, strongly) convergent sequences in the dual of a space Kμ of type Hankel-K{Mp} can be represented as distributional derivatives of functions and of measures that satisfy good properties, not only in terms of the differential operator x1D and of the Bessel operator Sμ=xμ12Dx2μ+1Dxμ12, but also in terms of suitable iterations Tμ,k (kN0) of the Zemanian differential operator Nμ=xμ+12Dxμ12. To this end, new descriptions for the usual topology of Kμ in terms of the latter iterations were given.

    The operators Tμ,k are defined by Tμ,k=Nμ+k1Nμ for kN, while Tμ,0 is the identity operator. The interest of such operators lies in their symmetric behavior with respect to hμ, the Hankel transformation of order μ. Indeed, given φL1(I), we define

    (hμφ)(y)=0φ(x)(xy)12Jμ(xy)dx(yI),

    where I=]0,[ and Jμ is the Bessel function of the first kind and order μ. For a sufficiently smooth φ with a good boundary behavior, the identity

    (y)mTμ,k(hμφ)(y)=0(x)kTμ,mφ(x)(xy)12Jμ+k+m(xy)dx=hμ+k+m[(x)kTμ,mφ(x)](y)(yI, m,kN0)

    holds [40, Proof of Theorem 5.4-1]. This rule closely resembles its Fourier counterpart and facilitates the operational calculus of the Hankel transformation, particularly when dealing with the Zemanian space Hμ [38,39], which can be regarded as a paradigm for the spaces of type Hankel-K{Mp}. In fact, the specialization to Hμ of the results obtained in our work improves upon previous results from [4]. The findings of the present paper are thus expected to pave the way for a time-frequency analysis of the spaces Kμ.

    Samuel García-Baquerín and Isabel Marrero: Conceptualization, Methodology, Investigation, Writing – Original Draft, Writing – Review & Editing. All authors of this article have contributed equally. Isabel Marrero: Supervision, Project Administration. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Samuel García-Baquerín would like to thank the Departamento de Análisis Matemático and the Instituto de Matemáticas y Aplicaciones (IMAULL) at the Universidad de La Laguna for their hospitality during the preparation of this paper.

    The authors declare they have no conflict of interest.



    [1] C. Arteaga, I. Marrero, Sobre la topología de los espacios Hankel-K{Mp}, Rev. Acad. Canaria Cienc., 21 (2009), 41–50.
    [2] C. Arteaga, I. Marrero, Structure, boundedness, and convergence in the dual of a Hankel-K{Mp} space, Integr. Transf. Spec. F., 32 (2021), 174–190. https://doi.org/10.1080/10652469.2020.1813128 doi: 10.1080/10652469.2020.1813128
    [3] J. J. Betancor, A characterization of Hankel transformable generalized functions, Int. J. Math. Math. Sci., 14 (1991), 269–274. https://doi.org/10.1155/S0161171291000303 doi: 10.1155/S0161171291000303
    [4] J. J. Betancor, I. Marrero, Some linear topological properties of the Zemanian space Hμ, Bull. Soc. Roy. Sci. Liège, 61 (1992), 299–314.
    [5] J. J. Betancor, I. Marrero, New spaces of type Hμ and the Hankel transformation, Integr. Transf. Spec. F., 3 (1995), 175–200. https://doi.org/10.1080/10652469508819075 doi: 10.1080/10652469508819075
    [6] J. J. Betancor, L. Rodríguez-Mesa, Hankel convolution on distribution spaces with exponential growth, Stud. Math., 121 (1996), 35–52. https://doi.org/10.4064/sm-121-1-35-52 doi: 10.4064/sm-121-1-35-52
    [7] J. J. Betancor, L. Rodríguez-Mesa, Characterizations of W-type spaces, Proc. Amer. Math. Soc., 126 (1998), 1371–1379. https://doi.org/10.1090/S0002-9939-98-04219-1 doi: 10.1090/S0002-9939-98-04219-1
    [8] A. J. Durán, Gelfand-Shilov spaces for the Hankel transform, Indag. Math. (N.S.), 3 (1992), 137–151. https://doi.org/10.1016/0019-3577(92)90002-3 doi: 10.1016/0019-3577(92)90002-3
    [9] S. J. L. Van Eijndhoven, M. J. Kerkhof, The Hankel transformation and spaces of type W, Reports on Applied and Numerical Analysis, Department of Mathematics and Computer Science, Eijndhoven University of Technology, 10 (1988).
    [10] A. Friedman, Generalized functions and partial differential equations, Englewood Cliffs, NJ: Prentice-Hall, 1963.
    [11] I. M. Gelfand, G. E. Shilov, Generalized functions, New York: Academic Press, 1 (1964).
    [12] I. M. Gelfand, G. E. Shilov, Generalized functions, New York: Academic Press, 2 (1968).
    [13] I. M. Gelfand, G. E. Shilov, Generalized functions, New York: Academic Press, 3 (1967).
    [14] I. M. Gelfand, N. Y. Vilenkin, Generalized Functions, New York: Academic Press, 4 (1964).
    [15] G. G. Gould, M. R. A. Sardar, A property of the Gelfand-Shilov test-function spaces of type K{Mp}, J. London Math. Soc., 1 (1969), 545–552.
    [16] M. Hasumi, Note on the n-dimensional tempered ultra-distributions, Tôhoku Math. J., 13 (1961), 99–104. https://doi.org/10.2748/tmj/1178244354 doi: 10.2748/tmj/1178244354
    [17] J. Horváth, Topological vector spaces and distributions, I, Reading, MA: Addison-Wesley, 1966.
    [18] A. Kamiński, Remarks on K{Mp}-spaces, Stud. Math., 77 (1984), 499–508. https://doi.org/10.4064/sm-77-5-499-508 doi: 10.4064/sm-77-5-499-508
    [19] A. Kamiński, D. Perišić, S. Pilipović, On the convolution in the Gel'fand-Shilov spaces, Integr. Transf. Spec. F., 4 (1996), 363–376.
    [20] L. Kitchens, C. Swartz, Convergence in the dual of certain K{Mp}-spaces, Colloq. Math., 30 (1974), 149–155. https://doi.org/10.4064/cm-30-1-149-155 doi: 10.4064/cm-30-1-149-155
    [21] W. Y. K. Lee, On spaces of type Hμ and their Hankel transformations, SIAM J. Math. Anal., 5 (1974), 336–348. https://doi.org/10.1137/0505037 doi: 10.1137/0505037
    [22] I. Marrero, A property characterizing Montel Hankel-K{Mp} spaces, Integr. Transf. Spec. F., 12 (2001), 65–76. https://doi.org/10.1080/10652460108819334 doi: 10.1080/10652460108819334
    [23] I. Marrero, Multipliers of Hankel-K{Mp} spaces, Integr. Transf. Spec. F., 12 (2001), 179–200. https://doi.org/10.1080/10652460108819343 doi: 10.1080/10652460108819343
    [24] I. Marrero, Hankel-K{Mp} spaces, Integr. Transf. Spec. F., 13 (2002), 379–401. https://doi.org/10.1080/10652460213522 doi: 10.1080/10652460213522
    [25] I. Marrero, On representation, boundedness and convergence of Hankel-K{Mp} generalized functions, Z. Anal. Anwend., 23 (2004), 731–743. https://doi.org/10.4171/zaa/1219 doi: 10.4171/zaa/1219
    [26] R. S. Pathak, H. K. Sahoo, A generalization of Hμ-spaces and Hankel transforms, Anal. Math., 12 (1986), 129–142. https://doi.org/10.1007/BF02027297 doi: 10.1007/BF02027297
    [27] R. S. Pathak, S. K. Upadhyay, Upμ-spaces and Hankel transform, Integr. Transf. Spec. F., 3 (1995), 285–300. https://doi.org/10.1080/10652469508819085 doi: 10.1080/10652469508819085
    [28] W. Rudin, Functional analysis, 2 Eds., New York: McGraw-Hill, 1991.
    [29] L. Schwartz, Théorie des distributions, I-II, Paris: Hermann, 1978.
    [30] J. Sebastião e Silva, Les fonctions analytiques comme ultra-distributions dans le calcul opérationnel, Math. Ann., 136 (1958), 58–96. https://doi.org/10.1007/BF01350287 doi: 10.1007/BF01350287
    [31] J. de Sousa Pinto, Ultradistributions of exponential type and the Fourier-Carleman transform, Port. Math., 50 (1993), 387–406.
    [32] C. Swartz, The nuclearity of K{Mp} spaces, Math. Nachr., 44 (1970), 193–197. https://doi.org/10.1002/mana.19700440116 doi: 10.1002/mana.19700440116
    [33] C. Swartz, Convolution in K{Mp} spaces, Rocky Mt. J. Math., 2 (1972), 259–263. https://doi.org/10.1216/RMJ-1972-2-2-259 doi: 10.1216/RMJ-1972-2-2-259
    [34] S. Sznajder, Z. Zieleźny, Solvability of convolution equations in K1, P. Am. Math. Soc., 57 (1976), 103–106. https://doi.org/10.1090/S0002-9939-1976-0410367-2 doi: 10.1090/S0002-9939-1976-0410367-2
    [35] S. Sznajder, Z. Zieleźny, Solvability of convolution equations in Kp, p>1, Pac. J. Math., 68 (1976), 539–544.
    [36] F. Tréves, Topological vector spaces, distributions, and kernels, New York: Academic Press, 1967.
    [37] S. Y. Tien, The topologies on the spaces of multipliers and convolution operators on K{Mp} spaces, Ph.D. Thesis, New Mexico State University, Las Cruces, NM, 1973.
    [38] A. H. Zemanian, A distributional Hankel transformation, SIAM J. Appl. Math., 14 (1966), 561–576. https://doi.org/10.1137/0114049 doi: 10.1137/0114049
    [39] A. H. Zemanian, The Hankel transformation of certain distributions of rapid growth, SIAM J. Appl. Math., 14 (1966), 678–690. https://doi.org/10.1137/0114056 doi: 10.1137/0114056
    [40] A. H. Zemanian, Generalized integral transformations, New York: Interscience, 1968.
    [41] Z. Zieleźny, On the space of convolution operators in K1, Stud. Math., 31 (1968), 111–124. https://doi.org/10.4064/sm-31-2-111-124 doi: 10.4064/sm-31-2-111-124
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(949) PDF downloads(37) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog