diff --git "a/4dFRT4oBgHgl3EQfpDe0/content/tmp_files/2301.13612v1.pdf.txt" "b/4dFRT4oBgHgl3EQfpDe0/content/tmp_files/2301.13612v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/4dFRT4oBgHgl3EQfpDe0/content/tmp_files/2301.13612v1.pdf.txt" @@ -0,0 +1,11231 @@ +On vector-valued functions and the +ε-product +Habilitationsschrift +vorgelegt am 31.01.2022 +der Technischen Universität Hamburg +von +Dr. rer. nat. Karsten Kruse, +geboren am 19.11.1984 in Papenburg. +Die Habilitationsschrift wurde in der Zeit von Juli 2020 bis Januar 2022 im +Institut für Mathematik der Technischen Universität Hamburg angefertigt. +arXiv:2301.13612v1 [math.FA] 31 Jan 2023 + +Gutachter: Prof. José Bonet +Prof. Dr. Leonhard Frerick +Prof. Dr. Thomas Kalmes +PD Dr. Christian Seifert +eingereicht: 31. Januar 2022; überarbeitet: 30. Januar 2023 +Tag des Habilitationskolloquiums: 01. Juli 2022 +DOI: 10.15480/882.4898 +ORCID: +0000-0003-1864-4915 +Creative Commons Lizenz: +Diese Arbeit steht unter der Creative Commons Lizenz Namensnennung 4.0 (CC +BY 4.0). Das bedeutet, dass sie vervielfältigt, verbreitet und öffentlich zugänglich +gemacht werden darf, auch kommerziell, sofern dabei stets der Urheber, die Quelle +des Textes und o. g. Lizenz genannt werden. Die genaue Formulierung der Lizenz +kann unter https://creativecommons.org/licenses/by/4.0/legalcode.de auf- +gerufen werden. + +Acknowledgement +It is quite hard to express how grateful I am to the people who helped, in one +way or the other, to finish this thesis which spans a part of my work between 2016 +and 2022. But I will give it a try. +First of all, I am deeply indebted to Marko Lindner and Christian Seifert who +always supported and encouraged me since I joined the TUHH in 2014 and gave +me a home so that I could work on the kind of mathematics I love. I know how +lucky I was to meet you both. +Second, I would like to thank the two people +who taught me probably the most I know about complex analysis and functional +analysis, namely, Andreas Defant and Michael Langenbruch (Oldenburg). Third, I +am utterly grateful to José Bonet and Enrique Jordá (Valencia) for many helpful +suggestions and comments, improving some of the papers this thesis is based on, +as well as enduring the quite abstract setting. +Let us come to the honorable mentions, ... just kidding. In 2015 I was lucky +again because Jan Meichsner joined the TUHH as a PhD student. Despite him +being a physicist and a dispraiser of green cabbage, it was a real pleasure to share +an office, do mathematics or just spend time with him. +Further, let me thank +Dennis Gallaun with whom I spent a lot of effort and gaffer tape setting up the +mobile e-assessment center at the TUHH between 2018 and 2020. +Apart from the people mentioned above I would like to thank my other mathe- +matical co-authors Hans Daduna, Ruslan Krenzler, Felix Schwenninger and Lin +Xie whose work is not physically present in this thesis but whose mathematical +influence or spirit probably is. Furthermore, I am thankful to the whole Institute +of Mathematics of the TUHH, in particular, the group of Applied Analysis for their +support. Moreover, I am grateful to the reviewers of this thesis Leonhard Frerick +(Trier) and Thomas Kalmes (Chemnitz) besides José Bonet and Christian Seifert, +and the anonymous reviewers of the papers it is based on for their work, helpful +comments and corrections. +Finally, I would like to thank my family for their continuous support and my +love Sonja for sharing my mathematical interests, bearing my kind of humour and +so much more which I cannot put into words. +These words only roughly express my gratitude and I hope that the minimum +that remains after reading the acknowledgement is the thought ‘At least, he gave +it a try.’ and a smile. +3 + + +Abstract +This habilitation thesis centres on linearisation of vector-valued functions which +means that vector-valued functions are represented by continuous linear operators. +The first question we face is which vector-valued functions may be represented +by continuous linear operators. +We study this problem in the framework of ε- +products and give sufficient conditions in Chapter 3 and 4 when a space F(Ω,E) +of vector-valued functions on a set Ω coincides (up to an isomorphism) with the +ε-product F(Ω)εE of a corresponding space of scalar-valued functions F(Ω) and +the codomain E which is usually an infinite-dimensional locally convex Hausdorff +space. The ε-product F(Ω)εE is a space of continuous linear operators from the +dual space F(Ω)′ to E. +Once we have a representation of a space F(Ω,E) of vector-valued functions +by an ε-product F(Ω)εE, we have access to the rich theory of continuous linear +operators which allows us to lift results that are known for the scalar-valued case +to the vector-valued case. The whole Chapter 5, which spans more than half of this +thesis, is dedicated to this lifting mechanism. But we should point out that this +is not only about transferring results from the scalar-valued to the vector-valued +case. The results in the vector-valued case encode additional information for the +scalar-valued case as well, e.g. we may deduce from the solvability of a linear partial +differential equation in the vector-valued case affirmative answers on the parameter +dependence of solutions in the scalar-valued case (see Section 5.1). +In Section 5.2 we give a unified approach to handle the problem of extending +functions with values in E, which have weak extensions in F(Ω), to functions in the +vector-valued counterpart F(Ω,E) of F(Ω). We present different extension the- +orems depending on the topological properties of the spaces F(Ω) and E. These +theorems also cover weak-strong principles. In particular, we study weak-strong +principles for continuously partially differentiable functions of finite order in Sec- +tion 5.3 and improve the well-known weak-strong principles of Grothendieck and +Schwartz. We use our results on the extension of vector-valued functions to de- +rive Blaschke’s convergence theorem for several spaces of vector-valued functions +and Wolff’s theorem for the description of dual spaces of several function spaces +F(Ω) in Section 5.4 and 5.5. Starting from the observation that every scalar-valued +holomorphic function has a local power series expansion and that this is still true +for holomorphic functions with values in E if E is locally complete, we develop a +machinery which is based on linearisation and Schauder decomposition to transfer +known series expansions from scalar-valued to vector-valued functions in Section +5.6. Especially, we apply this machinery to derive Fourier expansions for E-valued +Schwartz functions and C∞-smooth functions on Rd that are 2π-periodic in each +variable. The last section of Chapter 5 is devoted to the representation of spaces +F(Ω,E) of vector-valued functions by sequence spaces, which can be used to iden- +tify the coefficient spaces of the series expansions from the preceding section, if one +knows the coefficient space in the scalar-valued case. Furthermore, we give several +new conditions on the Pettis-integrability of vector-valued functions in Appendix +A.2, which are, for instance, needed for the Fourier expansions in Section 5.6. +5 + + +Kurzfassung +Im Mittelpunkt dieser Habilitationsschrift steht die Linearisierung vektorwer- +tiger Funktionen, d. h. vektorwertige Funktionen sollen durch stetige lineare Opera- +toren dargestellt werden. Die erste Frage, der man sich stellen muss, ist, welche +vektorwertigen Funktionen durch stetige lineare Operatoren dargestellt werden kön- +nen. +Wir untersuchen dieses Problem im Rahmen von ε-Produkten und geben +hinreichende Bedingungen in Kapitel 3 und 4 an, wann ein Raum F(Ω,E) von +vektorwertigen Funktionen auf einer Menge Ω mit dem ε-Produkt F(Ω)εE eines +entsprechenden Raums skalarwertiger Funktionen F(Ω) und des Wertebereichs E +(bis auf Isomorphie) übereinstimmt. +Hierbei ist E üblicherweise ein unendlich- +dimensionaler lokalkonvexer Hausdorff Raum. Das ε-Produkt F(Ω)εE ist ein Raum +stetiger linearer Operatoren, die vom Dualraum F(Ω)′ nach E abbilden. +Sobald wir eine Darstellung eines Raums F(Ω,E) von vektorwertigen Funk- +tionen durch ein ε-Produkt F(Ω)εE gewonnen haben, ist es uns möglich die reich- +haltige Theorie der stetigen linearen Operatoren zu nutzen, die es uns erlaubt, +Ergebnisse, die für den skalarwertigen Fall bekannt sind, auf den vektorwertigen +Fall zu übertragen. Das gesamte Kapitel 5, das mehr als die Hälfte dieser Arbeit +einnimmt, widmet sich diesem Übertragungsmechanismus. +Es sei jedoch darauf +hingewiesen, dass es hier nicht nur um die Übertragung von Ergebnissen aus dem +skalarwertigen auf den vektorwertigen Fall geht. Die Ergebnisse im vektorwertigen +Fall beinhalten auch zusätzliche Informationen für den skalarwertigen Fall, z. B. +können wir aus der Lösbarkeit einer linearen partiellen Differentialgleichung im +vektorwertigen Fall Antworten auf die Frage nach der Parameterabhängigkeit der +Lösungen im skalarwertigen Fall ableiten (siehe Abschnitt 5.1). +In Abschnitt 5.2 stellen wir einen einheitlichen Ansatz zur Lösung des Fort- +setzungsproblems von Funktionen mit Werten in E, die schwache Fortsetzungen in +F(Ω) haben, zu Funktionen im vektorwertigen Gegenstück F(Ω,E) von F(Ω) vor. +Wir präsentieren verschiedene Fortsetzungssätze in Abhängigkeit von den topologi- +schen Eigenschaften der Räume F(Ω) und E. Diese Sätze decken auch schwach- +stark Prinzipien ab. +Insbesondere untersuchen wir schwach-stark Prinzipien für +endlich oft stetig partiell differenzierbare Funktionen in Abschnitt 5.3 und verbes- +sern die bekannten schwach-starken Prinzipien von Grothendieck und Schwartz. +Zudem leiten wir von unseren Ergebnissen zur Fortsetzung vektorwertiger Funktio- +nen den Konvergenzsatz von Blaschke für diverse Räume vektorwertiger Funktionen +ab und übertragen den Satz von Wolff auf Dualräume mehrerer Funktionenräume +F(Ω) in den Abschnitten 5.4 und 5.5. +Ausgehend von der Beobachtung, dass +jede skalarwertige holomorphe Funktion eine lokale Potenzreihenentwicklung hat +und dass dies auch für holomorphe Funktionen mit Werten in E gilt, wenn E +lokal vollständig ist, entwickeln wir einen Mechanismus, der auf Linearisierung und +Schauder-Zerlegung basiert, um in Abschnitt 5.6 bekannte Reihenentwicklungen +von skalarwertigen auf vektorwertige Funktionen zu erweitern. Insbesondere wen- +den wir diesen Mechanismus an, um Fourier-Entwicklungen für E-wertige Schwartz- +Funktionen und C∞-glatte Funktionen auf Rd, die 2π-periodisch in jeder Variablen +sind, zu erhalten. +Der letzte Abschnitt von Kapitel 5 ist der Darstellung von +7 + +8 +KURZFASSUNG +Räumen F(Ω,E) vektorwertiger Funktionen durch Folgenräume gewidmet, was +man dazu nutzen kann, die Koeffizientenräume der Reihenentwicklungen aus dem +vorangegangenen Abschnitt zu bestimmen, sofern man den Koeffizientenraum im +skalarwertigen Fall kennt. Außerdem legen wir mehrere neue Bedingungen für die +Pettis-Integrierbarkeit von vektorwertigen Funktionen in Anhang A.2 dar, die z. B. +für die Fourier-Entwicklungen in Abschnitt 5.6 benötigt werden. + +Contents +Acknowledgement +3 +Abstract +5 +Kurzfassung +7 +Chapter 1. +Introduction +11 +Chapter 2. +Notation and preliminaries +17 +Chapter 3. +The ε-product for weighted function spaces +21 +3.1. +ε-into-compatibility +21 +3.2. +ε-compatibility +28 +Chapter 4. +Consistency +37 +4.1. +The spaces AP(Ω,E) and consistency +37 +4.2. +Further examples of ε-products +41 +4.3. +Riesz–Markov–Kakutani representation theorems +59 +Chapter 5. +Applications +67 +5.1. +Lifting the properties of maps from the scalar-valued case +67 +5.2. +Extension of vector-valued functions +74 +5.2.1. +Extension from thin sets +76 +5.2.2. +Extension from thick sets +92 +5.3. +Weak-strong principles for differentiability of finite order +104 +5.4. +Vector-valued Blaschke theorems +110 +5.5. +Wolff type results +114 +5.6. +Series representation of vector-valued functions via Schauder +decompositions +117 +5.6.1. +Schauder decomposition +120 +5.6.2. +Examples of Schauder decompositions +123 +5.7. +Representation by sequence spaces +130 +Appendices +Appendix A. +Compactness of closed absolutely convex hulls and Pettis- +integrals +139 +A.1. +Compactness of closed absolutely convex hulls +139 +A.2. +The Pettis-integral +142 +List of Symbols +147 +Index +151 +Bibliography +153 +9 + + +CHAPTER 1 +Introduction +This work is dedicated to a classical topic, namely, the linearisation of weighted +spaces of vector-valued functions. The setting we are interested in is the following. +Let F(Ω) be a locally convex Hausdorff space of functions from a non-empty set Ω +to a field K and E be a locally convex Hausdorff space over K. The ε-product of +F(Ω) and E is defined as the space of linear continuous operators +F(Ω)εE ∶= Le(F(Ω)′ +κ,E) +equipped with the topology of uniform convergence on equicontinuous subsets of +the dual F(Ω)′ which itself is equipped with the topology of uniform convergence +on absolutely convex compact subsets of F(Ω). Suppose that the point-evaluation +functionals δx, x ∈ Ω, belong to F(Ω)′ and that there is a locally convex Hausdorff +space F(Ω,E) of E-valued functions on Ω such that the map +S∶F(Ω)εE → F(Ω,E), u �→ [x ↦ u(δx)], +(1) +is well-defined. The main question we want to answer reads as follows. When is +F(Ω)εE a linearisation of F(Ω,E), i.e. when is S an isomorphism? +In [15, 16, 17] Bierstedt treats the space CV(Ω,E) of continuous functions on +a completely regular Hausdorff space Ω weighted with a Nachbin-family V and its +topological subspace CV0(Ω,E) of functions that vanish at infinity in the weighted +topology. He derives sufficient conditions on Ω, V and E such that the answer +to our question is affirmative, i.e. S is an isomorphism. +Schwartz answers this +question for several weighted spaces of k-times continuously partially differentiable +functions on Ω = Rd like the Schwartz space in [158, 159] for quasi-complete E +with regard to vector-valued distributions. +Grothendieck treats the question in +[83], mainly for nuclear F(Ω) and complete E. In [99, 100, 101] Komatsu gives a +positive answer for ultradifferentiable functions of Beurling or Roumieu type and +sequentially complete E with regard to vector-valued ultradistributions. For the +space of k-times continuously partially differentiable functions on open subsets Ω +of infinite dimensional spaces equipped with the topology of uniform convergence +of all partial derivatives up to order k on compact subsets of Ω sufficient conditions +for an affirmative answer are deduced by Meise in [129]. For holomorphic functions +on open subsets of infinite dimensional spaces a positive answer is given in [52] +by Dineen. Bonet, Frerick and Jordá show in [30] that S is an isomorphism for +certain closed subsheaves of the sheaf C∞(Ω,E) of smooth functions on an open +subset Ω ⊂ Rd with the topology of uniform convergence of all partial derivatives +on compact subsets of Ω and locally complete E which, in particular, covers the +spaces of harmonic and holomorphic functions. +An important application of linearisation is within the field of partial differen- +tial equations. Let E be a linear space of functions on a set U and P(∂)∶C∞(Ω) → +C∞(Ω) a linear partial differential operator with C∞-smooth coefficients where +C∞(Ω) ∶= C∞(Ω,K). We call the elements of U parameters and say that a family +(fλ)λ∈U in C∞(Ω) depends on a parameter w.r.t. E if the map λ ↦ fλ(x) is an +element of E for every x ∈ Ω. The question of parameter dependence is whether for +11 + +12 +1. INTRODUCTION +every family (fλ)λ∈U in C∞(Ω) depending on a parameter w.r.t. E there is a family +(uλ)λ∈U in C∞(Ω) with the same kind of parameter dependence which solves the +partial differential equation +P(∂)uλ = fλ, +λ ∈ U. +In particular, it is the question of Ck-smooth (holomorphic, distributional, etc.) +parameter dependence if E is the space Ck(U) of k-times continuously partially +differentiable functions on an open set U ⊂ Rd (the space O(U) of holomorphic +functions on an open set U ⊂ C, the space of distributions D(V )′ on an open set +V ⊂ Rd where U = D(V ), etc.). The question of parameter dependence w.r.t. E has +an affirmative answer for several locally convex Hausdorff spaces E due to tensor +product techniques and splitting theory. Indeed, the answer is affirmative if the +topology of E is stronger than the topology of pointwise convergence on U and +P(∂)E∶C∞(Ω,E) → C∞(Ω,E) +is surjective where P(∂)E is the version of P(∂) for E-valued functions. The oper- +ator P(∂)E is surjective if its version P(∂) for scalar-valued functions is surjective, +for instance, if P(∂) is elliptic, and E is a Fréchet space. This is a consequence +of Grothendieck’s theory of tensor products [83], the nuclearity of C∞(Ω) and the +isomorphism C∞(Ω,E) ≅ C∞(Ω)εE for locally complete E. Thanks to the splitting +theory of Vogt for Fréchet spaces [173] and of Bonet and Domański for PLS-spaces +[54] we even have in case of an elliptic P(∂) that P(∂)E for d > 1 is surjective if +E ∶= F ′ +b where F is a Fréchet space satisfying the condition (DN) or if E is an ultra- +bornological PLS-space having the property (PA) since kerP(∂) has the property +(Ω). In particular, these three results cover the cases that E = Ck(U), O(U) or +D(V )′. Of course, this technique to answer the question of parameter dependence +is not restricted to linear partial differential operators or the space C∞(Ω). +Another application of linearisation lies in the problem of extending a vector- +valued function f∶Λ → E from a subset Λ ⊂ Ω to a locally convex Hausdorff space +E if the scalar-valued functions e′ ○ f are extendable for each continuous linear +functional e′ from certain linear subspaces G of E′ under the constraint of preserving +the properties, like holomorphy, of the scalar-valued extensions. This problem was +considered, among others, by Grothendieck [82, 83], Bierstedt [17], Gramsch [77], +Grosse-Erdmann [79, 81], Arendt and Nikolski [6, 7, 8], Bonet, Frerick, Jordá and +Wengenroth [30, 69, 70, 92, 93]. Even the simple case Λ = Ω and G = E′ is interesting +and an affirmative answer is called a weak-strong principle. +Our goal is to give a unified and flexible approach to linearisation which is able +to handle new examples and covers the already known examples. +Organisation of this thesis +After fixing some notions and preliminaries on locally convex Hausdorff spaces, +continuous linear operators and continuously partially differentiable functions in +Chapter 2, we study the problem of linearisation in Chapter 3. In Section 3.1 we +introduce our standard example of spaces F(Ω,E) that we consider. Namely, spaces +of functions FV(Ω,E) from Ω to E which are subspaces of sections of domains of +linear operators T E on EΩ, and whose topology is generated by a family of weight +functions V. These spaces cover many examples of classical spaces of functions +appearing in analysis like the mentioned ones and an example of the operators T E +are the partial derivative operators. Then we exploit the structure of our spaces +to describe a sufficient condition, which we call consistency, on the interplay of the +pairs of operators (T E,T K) and the map S such that S becomes an isomorphism +into, i.e. an isomorphism to its range (see Theorem 3.1.12). + +ORGANISATION OF THIS THESIS +13 +In Section 3.2 we tackle the problem of surjectivity of S. In our main Theorem +3.2.4 and its Corollary 3.2.5 we give several sufficient conditions on the pairs of +operators (T E,T K) and the spaces involved such that S∶FV(Ω)εE → FV(Ω,E) is +an isomorphism. Looking at the pair of partial differential operators (P(∂)E,P(∂)) +considered above, these conditions allow us to express P(∂)E as P(∂)E = S ○ +(P(∂)εidE) ○ S−1 where P(∂)εidE is the ε-product of P(∂) and the identity idE +on E. Hence it becomes obvious that the surjectivity of P(∂)E is equivalent to +the surjectivity of P(∂)εidE. This is used in [105, 109, 112, 116, 119] in the case +of the Cauchy–Riemann operator P(∂) = ∂ on spaces of smooth functions with +exponential growth. +In Chapter 4 we take a closer look at the notion of consistency of (T E,T K). In +Section 4.1 we characterise several properties of the functions S(u) for u ∈ FV(Ω)εE +that are inherited from the elements of FV(Ω). +Section 4.2 is devoted to several concrete examples of spaces of vector-valued +functions that may be linearised by S and which we use for our applications in the +forthcoming sections and chapters. +In Section 4.3 we answer in several cases the question whether given a con- +tinuous linear functional T K on F(Ω) there is always a continuous linear operator +T E on F(Ω,E) such that (T E,T K) is consistent. This is closely related to Riesz– +Markov–Kakutani theorems for T K, which we transfer to the vector-valued case. +Chapter 5 is dedicated to applications of linearisation. In Section 5.1 we come +back to our problem of parameter dependence. We show in our main Theorem 5.1.2 +of this section how to use linearisations to transfer properties like injectivity, sur- +jectivity or bijectivity from a map T K∶F1(Ω1) → F2(Ω2) to the corresponding map +T E∶F2(Ω1,E) → F2(Ω2,E) if the pair (T E,T K) is consistent under suitable as- +sumptions on the spaces involved. Besides the problem of parameter dependence +for (hypo)elliptic linear partial differential operators (see Corollary 5.1.3), we de- +duce a vector-valued version of the Borel–Ritt theorem (see Theorem 5.1.4) from +this main theorem and give sufficient conditions under which the Fourier transfor- +mation FC∶Sµ(Rd) → Sµ(Rd) on the Beurling–Björck space is still an isomorphism +in the vector-valued case and may be decomposed as FE = S ○ (FCεidE) ○ S−1 (see +Theorem 5.1.5). +In Section 5.2 we present a general approach to the extension problem consid- +ered above for a large class of function spaces F(Ω,E) if the map S is an isomor- +phism into. The spaces we treat are of the kind that F(Ω) belongs to the class of +semi-Montel, Fréchet–Schwartz or Banach spaces, or that E is a semi-Montel space. +Apart from linearisation and consistency, the main ingredient of this approach is to +view the set Λ ⊂ Ω from which we want to extend our functions as a set of function- +als {δx ∣ x ∈ Λ}. This view allows us to generalise the extension problem in Question +5.2.1 by swapping this set of functionals by other functionals, which opens up new +possibilities in applications that we explore in Section 5.3, Section 5.4, Section 5.5 +and Section 5.7. In the extension problem we always have to balance the sets Λ +from which we extend our functions and the subspaces G ⊂ E′ with which we test. +The case of ‘thin’ sets Λ and ‘thick’ subspaces G is handled in Section 5.2.1 with +main theorems Theorem 5.2.15, Theorem 5.2.20 and Theorem 5.2.29, the converse +case of ‘thick’ sets Λ and ‘thin’ subspaces G is handled in Section 5.2.2 with main +theorems Theorem 5.2.52, Theorem 5.2.63 and Theorem 5.2.69. +In Section 5.3 we consider weak-strong principles for continuously partially +differentiable functions of finite order. For locally complete E it is well-known that +a function f belongs to C∞(Ω,E) if and only if e′ ○ f ∈ C∞(Ω) for all e′ ∈ E′ (see +e.g. [30, Theorem 9, p. 232]). If k ∈ N0, then it is still true that f ∈ Ck(Ω,E) implies +e′ ○ f ∈ Ck(Ω) for all e′ ∈ E′. But the converse is not true anymore. Only a weaker + +14 +1. INTRODUCTION +version of this weak-strong principle holds which is due to Grothendieck [82] and +Schwartz [158] (see Theorem 5.3.2). Namely, if k ∈ N0, E is sequentially complete +and f∶Ω → E is such that e′ ○ f ∈ Ck+1(Ω) for all e′ ∈ E′, then f ∈ Ck(Ω,E). Using +the results from Section 5.2, we improve this weaker version of the weak-strong +principle by allowing E to be locally complete, only testing with less functionals +from certain linear subspaces G ⊂ E′ and getting that f does not only belong to +Ck(Ω,E) but that all partial derivatives of order k are actually locally Lipschitz +continuous (see Corollary 5.3.5). If we restrict to semi-Montel spaces E, then even +a ‘full’ weak-strong principle Theorem 5.3.6 holds as in the C∞-case. +In Section 5.4 we derive vector-valued Blaschke theorems like Corollary 5.4.2 +for several function spaces. +This generalises results of Arendt and Nikolski [7] +for bounded holomorphic functions and Frerick, Jordá and Wengenroth [70] for +bounded functions in the kernel of a hypoelliptic linear partial differential operator. +These are results of the form: given a bounded net (fι)ι∈I in some space F1(Ω,E) +of Banach-valued functions which converges pointwise on a certain subset of Ω there +is a limit f ∈ F1(Ω,E) of this net w.r.t. a weaker topology of a linear superspace +F2(Ω,E) of F1(Ω,E). In Blaschke’s classical convergence theorem [38, Theorem +7.4, p. 219] we have E = C, F1(Ω,E) is the space of bounded holomorphic functions +on the open unit disc D ⊂ C, F2(Ω,E) is the space of holomorphic functions on D +and the weaker topology is the topology of compact convergence. +In Section 5.5 we present Wolff type descriptions of the dual space of several +function spaces F(Ω) using linearisation (see Theorem 5.5.1). Wolff’s theorem [183, +p. 1327] (cf. [81, Theorem (Wolff), p. 402]) phrased in a functional analytic way +(see [70, p. 240]) says: if Ω ⊂ C is a domain, then for each µ ∈ O(Ω)′ there are a +sequence (zn)n∈N which is relatively compact in Ω and a sequence (an)n∈N in the +space ℓ1 of absolutely summable sequences such that µ = ∑∞ +n=1 anδzn. +In Section 5.6 we derive a general result for Schauder decompositions of the +ε-product FεE for locally convex Hausdorff spaces F and E if F has an equicon- +tinuous Schauder basis (see Theorem 5.6.1). In combination with linearisation and +consistency this can be used for F = F(Ω) to lift series representations like the +power series expansion of holomorphic functions from scalar-valued functions to +vector-valued functions (see Corollary 5.6.5). We present several examples in Sec- +tion 5.6.2, for instance, Fourier expansions in the Schwartz space S(Rd,E) and in +the space C∞ +2π(Rd,E) of functions in C∞(Rd,E) that are 2π-periodic in each vari- +able. In particular, we combine these expansions for locally complete E with the +results from Section 5.1 to identify the coefficient spaces of the Fourier expansions +in S(Rd,E) and C∞ +2π(Rd,E) (see Theorem 5.6.13 and Theorem 5.6.14). +In Section 5.7 an application of our extension results from Section 5.2 is given to +represent function spaces F(Ω,E) by sequence spaces if one knows such a represen- +tation for F(Ω) (see Theorem 5.7.1). As examples we treat the space O(DR(0),E) +of E-valued holomorphic functions on the disc DR(0) ⊂ C around 0 with radius +0 < R ≤ ∞ and the multiplier space OM(R,E) of the Schwartz space for locally +complete E (see Corollary 5.7.2, Corollary 5.7.3 and Remark 5.7.4). +The first section Appendix A.1 of the Appendix A is devoted to the question +when the closure of an absolutely convex hull of a set is compact in a locally convex +Hausdorff space E and Appendix A.2 to the related question of Pettis-integrability +of an E-valued function. + +CONCERNING ORIGINALITY +15 +Concerning originality +We note that some parts of chapters or sections are based on our papers and +preprints. +● Chapter 3, Section 4.1 and Section 4.2 are based on our paper Weighted +spaces of vector-valued functions and the ε-product [110] and its extended +preprint [106]. Furthermore, Section 4.2 contains results from Sections 3 +and 6 of our accepted preprint Extension of weighted vector-valued func- +tions and sequence space representation [115] and our paper Extension of +weighted vector-valued functions and weak–strong principles for differen- +tiable functions of finite order [117] and its extended preprint [120]. +● Section 5.1 generalises some results of our papers Surjectivity of the ∂- +operator between weighted spaces of smooth vector-valued functions [116] +and Parameter dependence of solutions of the Cauchy–Riemann equation +on weighted spaces of smooth functions [112] and its extended preprint +[108]. +● Section 5.2, Section 5.3, Section 5.4, Section 5.5 and Section 5.7 are based +on our accepted preprint [115] and our paper [117] (and its extended +preprint [120]). +● Section 5.6 is based on our paper Series representations in spaces of vector- +valued functions via Schauder decompositions [114]. +Moreover, the introduction Chapter 1 and Chapter 2 on notation and pre- +liminaries are based on the corresponding sections in our papers and preprints +[106, 110, 112, 114, 115, 116, 117, 120]. However, not all of the results given in this +thesis are already contained in our preprints or papers. +In Chapter 3 the new, i.e. not contained in our preprints or papers, results are +Corollary 3.2.5 (ii), Example 3.2.7 e)+f), Example 3.2.9 and Corollary 3.2.10. +In Section 4.2 the new examples and results are Example 4.2.2, Corollary +4.2.3, Example 4.2.11, Example 4.2.13, Proposition 4.2.14, Example 4.2.16, Ex- +ample 4.2.22 which extends [107, Proposition 3.17 a), p. 244] of our paper The +approximation property for weighted spaces of differentiable function [107], Propo- +sition 4.2.25 which extends [114, Proposition 4.8, p. 370] from sequentially complete +E to locally complete E, Example 4.2.26 and Example 4.2.28 (ii). All the results +of Section 4.3 are new except for Definition 4.3.1 which is [115, 2.2 Definition, p. 4] +(and also not a result). +The main theorem of Section 5.1, Theorem 5.1.2, is new even though special +cases appeared in [112, 116]. Theorem 5.1.4 and Theorem 5.1.5 are new as well. +Corollary 5.4.3 extends [120, 7.3 Corollary, p. 22] from metric spaces with finite +diameter to arbitrary metric spaces. Theorem 5.6.13 and Theorem 5.6.14 b) extend +[114, Theorem 4.9, p. 371–372] and [114, Theorem 4.11, p. 375] from sequentially +complete E to locally complete E. Corollary 5.7.2 is new in the sense that there is +only a sketch how to prove it in [115, p. 31]. +The results of Appendix A are also new except for Proposition A.1.1, Propo- +sition A.1.6, which are contained in [106, 5.2 Proposition, p. 24] and [106, 3.13 +Lemma d), p. 10], and Lemma A.2.2 which is [114, Lemma 4.7, p. 369]. + + +CHAPTER 2 +Notation and preliminaries +Basics of topology +We equip the spaces Rd, d ∈ N, and C with the usual Euclidean norm ∣ ⋅ ∣, +denote by Br(x) ∶= {w ∈ Rd ∣ ∣w − x∣ < r} the ball around x ∈ Rd and by Dr(z) ∶= +{w ∈ C ∣ ∣w − z∣ < r} the disc around z ∈ C with radius r > 0. Furthermore, for a +subset M of a topological space (X,t) we denote the closure of M by M and the +boundary of M by ∂M. If we want to emphasize that we take the closure in X +resp. w.r.t. the topology t, then we write M +X resp. M +t. For a subset M of a vector +space X we denote by ch(M) the circled hull, by cx(M) the convex hull and by +acx(M) the absolutely convex hull of M. If X is a topological vector space, we +write acx(M) for the closure of acx(M) in X. +Locally convex Hausdorff spaces and continuous linear operators +By E we always denote a non-trivial, i.e. E ≠ {0}, locally convex Hausdorff +space over the field K = R or C equipped with a directed fundamental system of +seminorms (pα)α∈A and, in short, we write that E is an lcHs. If E = K, then we set +(pα)α∈A ∶= {∣ ⋅ ∣}. +By XΩ we denote the set of maps from a non-empty set Ω to a non-empty +set X, by χK we mean the characteristic function of K ⊂ Ω, by C(Ω,X) the space +of continuous functions from a topological space Ω to a topological space X, and +by C0(Ω,X) its subspace of continuous functions that vanish at infinity if X is a +locally convex Hausdorff space. +We denote by L(F,E) the space of continuous linear operators from F to E +where F and E are locally convex Hausdorff spaces. +If E = K, we just write +F ′ ∶= L(F,K) for the dual space and G○ for the polar set of G ⊂ F. If F and E +are linearly topologically isomorphic, we just write that F and E are isomorphic, +in symbols F ≅ E. We denote by Lt(F,E) the space L(F,E) equipped with the +locally convex topology t of uniform convergence on the finite subsets of F if t = σ, +on the absolutely convex, compact subsets of F if t = κ, on the absolutely convex, +σ(F,F ′)-compact subsets of F if t = τ, on the precompact (totally bounded) subsets +of F if t = γ and on the bounded subsets of F if t = b. We use the symbols t(F ′,F) +for the corresponding topology on F ′ and t(F) for the corresponding bornology on +F. We say that a subspace G ⊂ F ′ is separating (the points of F) if for every x ∈ F +it follows from y(x) = 0 for all y ∈ G that x = 0. Clearly, this is equivalent to G +being σ(F ′,F)-dense in F ′. For details and notions on the theory of locally convex +spaces not explained in this thesis see [68, 89, 131, 138]. +ε-products and tensor products +The so-called ε-product of Schwartz is defined by +FεE ∶= Le(F ′ +κ,E) +(2) +where L(F ′ +κ,E) is equipped with the topology of uniform convergence on equicon- +tinuous subsets of F ′. This definition of the ε-product coincides with the original +17 + +18 +2. NOTATION AND PRELIMINARIES +one by Schwartz [159, Chap. I, §1, Définition, p. 18]. It is symmetric which means +that FεE ≅ EεF. In the literature the definition of the ε-product is sometimes done +the other way around, i.e. EεF is defined by the right-hand side of (2) but due to the +symmetry these definitions are equivalent and for our purpose the given definition +is more suitable. If we replace F ′ +κ by F ′ +γ, we obtain Grothendieck’s definition of the +ε-product and we remark that the two ε-products coincide if F is quasi-complete +because then F ′ +γ = F ′ +κ holds. However, we stick to Schwartz’ definition. +For locally convex Hausdorff spaces Fi, Ei and Ti ∈ L(Fi,Ei), i = 1,2, we define +the ε-product T1εT2 ∈ L(F1εF2,E1εE2) of the operators T1 and T2 by +(T1εT2)(u) ∶= T2 ○ u ○ T t +1, +u ∈ F1εF2, +where T t +1∶E′ +1 → F ′ +1, e′ ↦ e′ ○ T1, is the dual map of T1. If T1 is an isomorphism +and F2 = E2, then T1εidE2 is also an isomorphism with inverse T −1 +1 εidE2 by [159, +Chap. I, §1, Proposition 1, p. 20] (or [89, 16.2.1 Proposition, p. 347] if the Fi are +complete). +As usual we consider the tensor product F ⊗E as a linear subspace of FεE for +two locally convex Hausdorff spaces F and E by means of the linear injection +Θ∶F ⊗ E → FεE, +k +∑ +n=1 +fn ⊗ en �→ [y ↦ +k +∑ +n=1 +y(fn)en]. +(3) +Via Θ the space F ⊗ E is identified with the space of operators with finite rank +in FεE and a locally convex topology is induced on F ⊗ E. We write F ⊗ε E for +F ⊗ E equipped with this topology and F ̂⊗εE for the completion of the injective +tensor product F ⊗εE. For more information on the theory of ε-products and tensor +products see [49, 89, 94]. +Several degrees of completeness +The sufficient conditions for surjectivity of the map S∶F(Ω)εE → F(���,E) +from the introduction, which we derive in the forthcoming, depend on assumptions +on different types of completeness of E. For this purpose we recapitulate some +definitions which are connected to completeness. We start with local completeness. +For a disk D ⊂ E, i.e. a bounded, absolutely convex set, the linear space ED ∶= +⋃n∈N nD becomes a normed space if it is equipped with the gauge functional of +D as a norm (see [89, p. 151]). The space E is called locally complete if ED is a +Banach space for every closed disk D ⊂ E (see [89, 10.2.1 Proposition, p. 197]). We +call a non-empty subset A of an lcHs E locally closed if every local limit point of +A belongs to A. Here, a point x ∈ E is called a local limit point of A if there is a +sequence (xn)n∈N in A that converges locally to x (see [138, Definition 5.1.14, p. +154–155]), i.e. there is a disk D ⊂ E such that (xn) converges to x in ED (see [138, +Definition 5.1.1, p. 151]). The local closure of a subset A of E is defined as the +smallest locally closed subset of E which contains A (see [138, Definition 5.1.18, p. +155]). Moreover, we note that every locally complete linear subspace of E is locally +closed and a locally closed linear subspace of a locally complete space is locally +complete by [138, Proposition 5.1.20 (i), p. 155]. +Moreover, a locally convex Hausdorff space is locally complete if and only if it +is convenient by [104, 2.14 Theorem, p. 20]. In particular, every complete locally +convex Hausdorff space is quasi-complete, every quasi-complete space is sequentially +complete and every sequentially complete space is locally complete and all these +implications are strict. The first two by [89, p. 58] and the third by [138, 5.1.8 +Corollary, p. 153] and [138, 5.1.12 Example, p. 154]. +Now, let us recall the following definition from [181, 9-2-8 Definition, p. 134] and +[175, p. 259]. A locally convex Hausdorff space is said to have the [metric] convex +compactness property ([metric] ccp) if the closure of the absolutely convex hull of + +VECTOR-VALUED CONTINUOUSLY PARTIALLY DIFFERENTIABLE FUNCTIONS +19 +every [metrisable] compact set is compact. Sometimes this condition is phrased +with the term convex hull instead of absolutely convex hull but these definitions +coincide. Indeed, the first definition implies the second since every convex hull of +a set A ⊂ E is contained in its absolutely convex hull. On the other hand, we have +acx(A) = cx(ch(A)) by [89, 6.1.4 Proposition, p. 103] and the circled hull ch(A) of a +[metrisable] compact set A is compact by [153, Chap. I, 5.2, p. 26] [and metrisable +by [34, Chap. IX, §2.10, Proposition 17, p. 159] since D × A is metrisable and +ch(A) = ME(D × A) where ME∶K × E → E is the continuous scalar multiplication +and D ∶= D1(0) the open unit disc], which yields the other implication. +In particular, every locally convex Hausdorff space with ccp has obviously met- +ric ccp, every quasi-complete locally convex Hausdorff space has ccp by [181, 9-2-10 +Example, p. 134], every sequentially complete locally convex Hausdorff space has +metric ccp by [23, A.1.7 Proposition (ii), p. 364] and every locally convex Hausdorff +space with metric ccp is locally complete by [175, Remark 4.1, p. 267]. All these +implications are strict. The second by [181, 9-2-10 Example, p. 134] and the others +by [175, Remark 4.1, p. 267]. For more details on the [metric] convex compactness +property and local completeness see [29, 175]. In addition, we remark that every +semi-Montel space is semi-reflexive by [89, 11.5.1 Proposition, p. 230] and every +semi-reflexive locally convex Hausdorff space is quasi-complete by [153, Chap. IV, +5.5, Corollary 1, p. 144] and these implications are strict as well. Summarizing, we +have the following diagram of strict implications: +semi-Montel ⇒ semi-reflexive +⇓ +complete ⇒ quasi-complete ⇒ sequentially complete ⇒ locally complete +⇓ +⇓ +�⇒ +ccp +⇒ +metric ccp +Vector-valued continuously partially differentiable functions +Since weighted spaces of continuously partially differentiable resp. holomorphic +vector-valued functions will serve as our standard examples, we recall the definition +of the spaces Ck(Ω,E) resp. O(Ω,E). A function f∶Ω → E on an open set Ω ⊂ Rd +to an lcHs E is called continuously partially differentiable (f is C1) if for the n-th +unit vector en ∈ Rd the limit +(∂en)Ef(x) ∶= +lim +h→0 +h∈R,h≠0 +f(x + hen) − f(x) +h +exists in E for every x ∈ Ω and (∂en)Ef is continuous on Ω ((∂en)Ef is C0) for +every 1 ≤ n ≤ d. For k ∈ N a function f is said to be k-times continuously partially +differentiable (f is Ck) if f is C1 and all its first partial derivatives are Ck−1. A +function f is called infinitely continuously partially differentiable (f is C∞) if f is +Ck for every k ∈ N. For k ∈ N∞ ∶= N∪{∞} the functions f∶Ω → E which are Ck form +a linear space which is denoted by Ck(Ω,E). For β ∈ Nd +0 with ∣β∣ ∶= ∑d +n=1 βn ≤ k +and a function f∶Ω → E on an open set Ω ⊂ Rd to an lcHs E we set (∂βn)Ef ∶= f +if βn = 0, and +(∂βn)Ef(x) ∶= (∂en)E⋯(∂en)E +������������������������������������������������������������������������������������������ +βn-times +f(x) +if βn ≠ 0 and the right-hand side exists in E for every x ∈ Ω. Further, we define +(∂β)Ef(x) ∶= ((∂β1)E⋯(∂βd)E)f(x) + +20 +2. NOTATION AND PRELIMINARIES +if the right-hand side exists in E for every x ∈ Ω. If E = K, we often just write +∂βf ∶= (∂β)Kf for β ∈ Nd +0, ∣β∣ ≤ k, and f ∈ Ck(Ω). Furthermore, we define the space +of bounded continuously partially differentiable functions by +C1 +b (Ω,E) ∶= {f ∈ C1(Ω,E) ∣ ∀ α ∈ A ∶ ∣f∣C1 +b (Ω),α ∶= +sup +x∈Ω +β∈Nd +0,∣β∣≤1 +pα((∂β)Ef(x)) < ∞}. +Vector-valued holomorphic functions +A function f∶Ω → E on an open set Ω ⊂ C to an lcHs E over C is called +holomorphic if the limit +(∂1 +C)Ef(z) ∶= +lim +h→0 +h∈C,h≠0 +f(z + h) − f(z) +h +, +z ∈ Ω, +exists in E. +We denote by O(Ω,E) the linear space of holomorphic functions +f∶Ω → E. Defining the vector-valued version of the Cauchy–Riemann operator by +∂ +Ef ∶= 1 +2((∂e1)E + i(∂e2)E)f +for f ∈ C(Ω,E) such that the partial derivatives (∂en)Ef, n = 1,2, exist in E, we +remark that +O(Ω,E) = {f ∈ C(Ω,E) ∣ f ∈ ker∂ +E} = {f ∈ C∞(Ω,E) ∣ f ∈ ker∂ +E} +(4) +by [113, Theorem 6.1, p. 267] if E is locally complete. Further, we set (∂0 +C)Ef ∶= f +and note that the (n + 1)-th complex derivative (∂n+1 +C +)Ef ∶= (∂1 +C)E((∂n +C)Ef) exists +for all n ∈ N0 and f ∈ O(Ω,E) by [79, 2.1 Theorem and Definition, p. 17–18] and +[79, 5.2 Theorem, p. 35] if E is locally complete. If E = C, we often just write +f (n) ∶= (∂n +C)Cf for n ∈ N0 and f ∈ O(Ω) ∶= O(Ω,C). We note that the real and +complex derivatives are related by +(∂β)Ef(z) = iβ2(∂∣β∣ +C )Ef(z), +z ∈ Ω, +(5) +for every f ∈ O(Ω,E) and β = (β1,β2) ∈ N2 +0 by [113, Proposition 7.1, p. 270] if E is +locally complete. + +CHAPTER 3 +The ε-product for weighted function spaces +3.1. ε-into-compatibility +In the introduction we already mentioned that linearisations of spaces of vector- +valued functions by means of ε-products are essential for our approach. +Here, +one of the important questions is which spaces of vector-valued functions can be +represented by ε-products. Let Ω be a non-empty set and E an lcHs. If F(Ω) ⊂ KΩ +is an lcHs such that δx ∈ F(Ω)′ for all x ∈ Ω, then the map +S∶F(Ω)εE → EΩ, u �→ [x ↦ u(δx)], +is well-defined and linear. This leads to the following definition. +3.1.1. Definition (ε-into-compatible). Let Ω be a non-empty set and E an +lcHs. Let F(Ω) ⊂ KΩ and F(Ω,E) ⊂ EΩ be lcHs such that δx ∈ F(Ω)′ for all x ∈ Ω. +We call the spaces F(Ω) and F(Ω,E) ε-into-compatible if the map +S∶F(Ω)εE → F(Ω,E), u �→ [x ↦ u(δx)], +is a well-defined isomorphism into, i.e. to its range. We call F(Ω) and F(Ω,E) +ε-compatible if S is an isomorphism. We write SF(Ω) if we want to emphasise the +dependency on F(Ω). +In this section we introduce the weighted space FV(Ω,E) of E-valued functions +on Ω as a subspace of sections of domains in EΩ of linear operators T E +m equipped +with a generalised version of a weighted graph topology. This space is the role +model for many function spaces and an example for these operators are the partial +derivative operators. Then we treat the question whether FV(Ω,E) and FV(Ω)εE +are ε-into-compatible. This is deeply connected with the interplay of the pair of +operators (T E +m,T K +m) with the map S (see Definition 3.1.7). In our main theorem of +this section we give sufficient conditions such that S∶FV(Ω)εE → FV(Ω,E) is an +isomorphism into (see Theorem 3.1.12). In the next section we provide conditions +such that S becomes surjective (see Theorem 3.2.4). We start with the well-known +example Ck(Ω,E) of k-times continuously partially differentiable E-valued func- +tions to motivate our definition of FV(Ω,E). +3.1.2. Example. Let k ∈ N∞ and Ω ⊂ Rd be open. Consider the space C(Ω,E) +of continuous functions f∶Ω → E with the topology τc of compact convergence, i.e. +the topology given by the seminorms +∥f∥K,α ∶= sup +x∈K +pα(f(x)), +f ∈ C(Ω,E), +for compact K ⊂ Ω and α ∈ A. The usual topology on the space Ck(Ω,E) of k-times +continuously partially differentiable functions is the graph topology generated by +the partial derivative operators (∂β)E∶Ck(Ω,E) → C(Ω,E) for β ∈ Nd +0, ∣β∣ ≤ k, i.e. +the topology given by the seminorms +∥f∥K,β,α ∶= max(∥f∥K,α,∥(∂β)Ef∥K,α), +f ∈ Ck(Ω,E), +21 + +22 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +for compact K ⊂ Ω, β ∈ Nd +0, ∣β∣ ≤ k, and α ∈ A. The same topology is induced by +the directed system of seminorms given by +∣f∣K,m,α ∶= +sup +β∈Nd +0,∣β∣≤m +∥f∥K,β,α = +sup +x∈K +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x)), +f ∈ Ck(Ω,E), +for compact K ⊂ Ω, m ∈ N0, m ≤ k, and α ∈ A and may also be seen as a weighted +topology induced by the family (χK) of characteristic functions of the compact sets +K ⊂ Ω by writing +∣f∣K,m,α = +sup +x∈Ω +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x))χK(x), +f ∈ Ck(Ω,E). +This topology is inherited by linear subspaces of functions having additional prop- +erties like being holomorphic or harmonic. +We turn to the weight functions which we use to define a kind of weighted +graph topology. +3.1.3. Definition (weight function). Let J be a non-empty set and (ωm)m∈M +a family of non-empty sets. We call V ∶= (νj,m)j∈J,m∈M a family of weight functions +on (ωm)m∈M if it fulfils νj,m∶ωm → [0,∞) for all j ∈ J, m ∈ M and +∀ m ∈ M, x ∈ ωm ∃ j ∈ J ∶ 0 < νj,m(x). +(6) +From the structure of Example 3.1.2 we arrive at the following definition of the +weighted spaces of vector-valued functions we want to consider. +3.1.4. Definition. Let Ω be a non-empty set, V ∶= (νj,m)j∈J,m∈M a family +of weight functions on (ωm)m∈M and T E +m∶EΩ ⊃ domT E +m → Eωm a linear map for +every m ∈ M. Let AP(Ω,E) be a linear subspace of EΩ and define the space of +intersections +F(Ω,E) ∶= AP(Ω,E) ∩ ( ⋂ +m∈M +domT E +m) +as well as +FV(Ω,E) ∶= {f ∈ F(Ω,E) ∣ ∀ j ∈ J, m ∈ M, α ∈ A ∶ ∣f∣j,m,α < ∞} +where +∣f∣j,m,α ∶= sup +x∈ωm +pα(T E +m(f)(x))νj,m(x) = +sup +e∈Nj,m(f) +pα(e) +with +Nj,m(f) ∶= {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm}. +Further, we write F(Ω) ∶= F(Ω,K) and FV(Ω) ∶= FV(Ω,K). If we want to empha- +sise dependencies, we write M(E) instead of M, APFV(Ω,E) instead of AP(Ω,E) +and ∣f∣FV(Ω),j,m,α instead of ∣f∣j,m,α. If J, M or A are singletons, we omit the index +j, m resp. α in ∣f∣j,m,α. +Note that ωm need not be a subset of Ω. The space AP(Ω,E) is a placeholder +where we collect additional properties (AP) of our functions not being reflected by +the operators T E +m which we integrated in the topology. However, these additional +properties might come from being in the domain or kernel of additional operators, +e.g. harmonicity means being in the kernel of the Laplacian. But often AP(Ω,E) +can be chosen as EΩ or C(Ω,E). The space FV(Ω,E) is locally convex but need not +be Hausdorff. Since it is easier to work with Hausdorff spaces and a directed family +of seminorms plus the point evaluation functionals δx∶FV(Ω) → K, f ↦ f(x), for +x ∈ Ω and their continuity play a big role, we introduce the following definition. + +3.1. ε-INTO-COMPATIBILITY +23 +3.1.5. Definition (dom-space and T E +m,x). We call FV(Ω,E) a dom-space if it +is a locally convex Hausdorff space, the system of seminorms (∣f∣j,m,α)j∈J,m∈M,α∈A +is directed and, in addition, δx ∈ FV(Ω)′ for every x ∈ Ω if E = K. We define the +point evaluation of T E +m by T E +m,x∶domT E +m → E, T E +m,x(f) ∶= T E +m(f)(x), for m ∈ M and +x ∈ ωm. +3.1.6. Remark. +a) It is easy to see that FV(Ω,E) is Hausdorff if there +is m ∈ M such that ωm = Ω and T E +m = idEΩ since E is Hausdorff. +b) If E = K, then T K +m,x ∈ FV(Ω)′ for every m ∈ M and x ∈ ωm. Indeed, for +m ∈ M and x ∈ ωm there exists j ∈ J such that νj,m(x) > 0 by (6), implying +for every f ∈ FV(Ω) that +∣T K +m,x(f)∣ = +1 +νj,m(x)∣T K +m(f)(x)∣νj,m(x) ≤ +1 +νj,m(x)∣f∣j,m. +In particular, this implies δx ∈ FV(Ω)′ for all x ∈ Ω if there is m ∈ M such +that ωm = Ω and T K +m = idKΩ. +c) Let the family of weight functions V be directed, i.e. +∀ j1,j2 ∈ J,m1,m2 ∈ M ∃ j3 ∈ J, m3 ∈ M, C > 0 ∀ i ∈ {1,2} ∶ +(ωm1 ∪ ωm2) ⊂ ωm3 +and +νji,mi ≤ Cνj3,m3. +Then the system of seminorms (∣f∣j,m,α)j∈J,m∈M,α∈A is directed if V is +directed and additionally it holds with mi, i ∈ {1,2,3}, from above that +∀ f ∈ FV(Ω,E), i ∈ {1,2}, x ∈ ωmi ∶ T E +mi(f)(x) = T E +m3(f)(x), +since the system (pα)α∈A of E is already directed. +We point out that the additional condition in Remark 3.1.6 c) is missing in +[110, Remark 5 c), p. 1516] (resp. [106, 3.5 Remark, p. 6]), which we correct here. +For the lcHs E over K we want to define a natural E-valued version of a dom- +space FV(Ω) = FV(Ω,K). The natural E-valued version of FV(Ω) should be a +dom-space FV(Ω,E) such that there is a canonical relation between the families +(T K +m) and (T E +m). This canonical relation will be explained in terms of their interplay +with the map +S∶FV(Ω)εE → EΩ, u �→ [x ↦ u(δx)]. +Further, the elements of our E-valued version FV(Ω,E) of FV(Ω) should be com- +patible with a weak definition in the sense that e′ ○f ∈ FV(Ω) should hold for every +e′ ∈ E′ and f ∈ FV(Ω,E). +3.1.7. Definition (generator, consistent, strong). Let FV(Ω) and FV(Ω,E) +be dom-spaces such that M ∶= M(K) = M(E). +a) We call (T E +m,T K +m)m∈M a generator for (FV(Ω),E), in short, (FV,E). +b) We call (T E +m,T K +m)m∈M consistent if we have for all u ∈ FV(Ω)εE that +S(u) ∈ F(Ω,E) and +∀ m ∈ M, x ∈ ωm ∶ (T E +mS(u))(x) = u(T K +m,x). +c) We call (T E +m,T K +m)m∈M strong if we have for all e′ ∈ E′, f ∈ FV(Ω,E) that +e′ ○ f ∈ F(Ω) and +∀ m ∈ M, x ∈ ωm ∶ T K +m(e′ ○ f)(x) = (e′ ○ T E +m(f))(x). +More precisely, T K +m,x in b) means the restriction of T K +m,x to FV(Ω) and the +term u(T K +m,x) is well-defined by Remark 3.1.6 b). Consistency will guarantee that +the map S∶FV(Ω)εE → FV(Ω,E) is a well-defined isomorphism into, i.e. ε-into- +compatibility, and strength will help us to prove its surjectivity under some ad- +ditional assumptions on FV(Ω) and E. Let us come to a lemma which describes + +24 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +the topology of FV(Ω)εE in terms of the operators T K +m with m ∈ M. It was the +motivation for the definition of consistency and allows us to consider FV(Ω)εE as +a topological subspace of FV(Ω,E) via S, assuming consistency. +3.1.8. Lemma. Let FV(Ω) be a dom-space. Then the topology of FV(Ω)εE is +given by the system of seminorms defined by +∥u∥j,m,α ∶= sup +x∈ωm +pα(u(T K +m,x))νj,m(x), +u ∈ FV(Ω)εE, +for j ∈ J, m ∈ M and α ∈ A. +Proof. We define the sets Dj,m ∶= {T K +m,x(⋅)νj,m(x) ∣ x ∈ ωm} and Bj,m ∶= {f ∈ +FV(Ω) ∣ ∣f∣j,m ≤ 1} for every j ∈ J and m ∈ M. We claim that acx(Dj,m) is dense +in the polar B○ +j,m with respect to κ(FV(Ω)′,FV(Ω)). The observation +D○ +j,m = {T K +m,x(⋅)νj,m(x) ∣ x ∈ ωm}○ += {f ∈ FV(Ω) ∣ ∀x ∈ ωm ∶ ∣T K +m(f)(x)∣νj,m(x) ≤ 1} += {f ∈ FV(Ω) ∣ ∣f∣j,m ≤ 1} = Bj,m +yields +acx(Dj,m)κ(FV(Ω)′,FV(Ω)) = (Dj,m)○○ = B○ +j,m +by the bipolar theorem. By [89, 8.4, p. 152, 8.5, p. 156–157] the system of seminorms +defined by +qj,m,α(u) ∶= sup +y∈B○ +j,m +pα(u(y)), +u ∈ FV(Ω)εE, +for j ∈ J, m ∈ M and α ∈ A gives the topology on FV(Ω)εE (here it is used that +the system of seminorms (∣ ⋅ ∣j,m) of FV(Ω) is directed). As every u ∈ FV(Ω)εE +is continuous on B○ +j,m, we may replace B○ +j,m by a κ(FV(Ω)′,FV(Ω))-dense subset. +Therefore we obtain +qj,m,α(u) = sup{pα(u(y)) ∣ y ∈ acx(Dj,m)}. +For y ∈ acx(Dj,m) there are n ∈ N, λk ∈ K, xk ∈ ωm, 1 ≤ k ≤ n, with ∑n +k=1 ∣λk∣ ≤ 1 +such that y = ∑n +k=1 λkT K +m,xk(⋅)νj,m(xk). Then we have for every u ∈ FV(Ω)εE +pα(u(y)) ≤ +n +∑ +k=1 +∣λk∣pα(u(T K +m,xk))νj,m(xk) ≤ ∥u∥j,m,α, +thus qj,m,α(u) ≤ ∥u∥j,m,α. On the other hand, we derive +qj,m,α(u) ≥ sup +y∈Dj,m +pα(u(y)) = sup +x∈ωm +pα(u(T K +m,x))νj,m(x) = ∥u∥j,m,α. +□ +Let us turn to a more general version of Example 3.1.2, namely, to weighted +spaces of k-times continuously partially differentiable functions and kernels of linear +partial differential operators in these spaces. +3.1.9. Example. Let k ∈ N∞ and Ω ⊂ Rd be open. We consider the cases +(i) ωm ∶= Mm × Ω with Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ min(m,k)} for all m ∈ N0, or +(ii) ωm ∶= Nd +0 × Ω for all m ∈ N0 and k = ∞, +and let Vk ∶= (νj,m)j∈J,m∈N0 be a directed family of weights on (ωm)m∈N0. +a) We define the weighted space of k-times continuously partially differentiable +functions with values in an lcHs E as +CVk(Ω,E) ∶= {f ∈ Ck(Ω,E) ∣ ∀ j ∈ J, m ∈ N0, α ∈ A ∶ ∣f∣j,m,α < ∞} +where +∣f∣j,m,α ∶= +sup +(β,x)∈ωm +pα((∂β)Ef(x))νj,m(β,x). + +3.1. ε-INTO-COMPATIBILITY +25 +Setting domT E +m ∶= Ck(Ω,E) and +T E +m∶Ck(Ω,E) → Eωm, f �→ [(β,x) ↦ (∂β)Ef(x)], +as well as AP(Ω,E) ∶= EΩ, we observe that CVk(Ω,E) is a dom-space by Remark +3.1.6 and +∣f∣j,m,α = sup +x∈ωm +pα(T E +mf(x))νj,m(x). +b) The space Ck(Ω,E) with its usual topology given in Example 3.1.2 is a special +case of a)(i) with J ∶= {K ⊂ Ω ∣ K compact}, νK,m(β,x) ∶= χK(x), (β,x) ∈ ωm, for +all m ∈ N0 and K ∈ J where χK is the characteristic function of K. In this case we +write Wk ∶= Vk for the family of weight functions. +c) The Schwartz space is defined by +S(Rd,E) ∶= {f ∈ C∞(Rd,E) ∣ ∀ m ∈ N0, α ∈ A ∶ ∣f∣m,α < ∞} +where +∣f∣m,α ∶= +sup +x∈Rd +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x))(1 + ∣x∣2)m/2. +This is a special case of a)(i) with k ∶= ∞, Ω ∶= Rd, J ∶= {1} and ν1,m(β,x) ∶= +(1 + ∣x∣2)m/2, (β,x) ∈ ωm, for all m ∈ N0. +d) The multiplier space for the Schwartz space is defined by +OM(Rd,E) ∶= {f ∈ C∞(Rd,E) ∣ ∀ g ∈ S(Rd), m ∈ N0, α ∈ A ∶ ∥f∥g,m,α < ∞} +where +∥f∥g,m,α ∶= +sup +x∈Rd +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x))∣g(x)∣ +(see [158, 40), p. 97]). This is a special case of a)(i) with k ∶= ∞, Ω ∶= Rd, J ∶= +{j ⊂ S(Rd) ∣ j finite} and νj,1,m(β,x) ∶= maxg∈j ∣g(x)∣, (β,x) ∈ ωm, for all m ∈ N0. +This choice of J guarantees that the family V∞ is directed and does not change the +topology. +e) Let K ∶= {K ⊂ Ω ∣ K compact} and (Mp)p∈N0 be a sequence of positive real +numbers. The space E(Mp)(Ω,E) of ultradifferentiable functions of class (Mp) of +Beurling-type is defined as +E(Mp)(Ω,E) ∶= {f ∈ C∞(Ω,E) ∣ ∀ K ∈ K, h > 0, α ∈ A ∶ ∣f∣(K,h),α < ∞} +where +∣f∣(K,h),α ∶= sup +x∈K +β∈Nd +0 +pα((∂β)Ef(x)) +1 +h∣β∣M∣β∣ +. +This is a special case of a)(ii) with J ∶= K × R>0 and ν(K,h),m(β,x) ∶= χK(x) +1 +h∣β∣M∣β∣ , +(β,x) ∈ ωm, for all (K,h) ∈ J and m ∈ N0 where R>0 ∶= (0,∞). +f) Let K and (Mp)p∈N0 be as in e). The space E{Mp}(Ω,E) of ultradifferentiable +functions of class {Mp} of Roumieu-type is defined as +E{Mp}(Ω,E) ∶= {f ∈ C∞(Ω,E) ∣ ∀ (K,H) ∈ J, α ∈ A ∶ ∣f∣(K,H),α < ∞} +where +J ∶= K × {H = (Hn)n∈N ∣ ∃ (hk)k∈N, hk > 0, hk ↗ ∞ ∀ n ∈ N ∶ Hn = h1 ⋅ ... ⋅ hn} +and +∣f∣(K,H),α ∶= sup +x∈K +β∈Nd +0 +pα((∂β)Ef(x)) +1 +H∣β∣M∣β∣ + +26 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +(see [101, Proposition 3.5, p. 675]). +Again, this is a special case of a)(ii) with +ν(K,H),m(β,x) ∶= χK(x) +1 +H∣β∣M∣β∣ , (β,x) ∈ ωm, for all (K,H) ∈ J and m ∈ N0. +g) Let n ∈ N, βi ∈ Nd +0 with ∣βi∣ ≤ k and ai∶Ω → K for 1 ≤ i ≤ n. We set +P(∂)E∶Ck(Ω,E) → EΩ, P(∂)E(f)(x) ∶= +n +∑ +i=1 +ai(x)(∂βi)E(f)(x) +and obtain the (topological) subspace of CVk(Ω,E) given by +CVk +P (∂)(Ω,E) ∶= {f ∈ CVk(Ω,E) ∣ f ∈ kerP(∂)E}. +Choosing AP(Ω,E) ∶= kerP(∂)E, we see that this is also a dom-space by a). If +P(∂)E is the Cauchy–Riemann operator (and E locally complete) or the Laplacian, +we obtain the weighted space of holomorphic resp. harmonic functions. +Let us show that the generators of these spaces are strong and consistent. In +order to obtain consistency for their generators we have to restrict to directed +families of weights which are locally bounded away from zero on Ω, i.e. +∀ K ⊂ Ω compact, m ∈ N0 ∃ j ∈ J ∀ β ∈ Nd +0, ∣β∣ ≤ min(m,k) ∶ inf +x∈K νj,m(β,x) > 0. +This condition on Vk guarantees that the map I∶CVk(Ω) → CWk(Ω), f ↦ f, is +continuous which is needed for consistency. +3.1.10. Proposition. Let E be an lcHs, k ∈ N∞, Vk be a directed family of +weights which is locally bounded away from zero on an open set Ω ⊂ Rd. +The +generator of (CVk,E) resp. (CVk +P (∂),E) from Example 3.1.9 is strong and consistent +if CVk(Ω) resp. CVk +P (∂)(Ω) is barrelled. +Proof. We recall the definitions from Example 3.1.9. We have ωm ∶= Mm × Ω +with Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ min(m,k)} for all m ∈ N0 or ωm ∶= Nd +0 × Ω for all m ∈ N0. +Further, APCVk(Ω,E) = EΩ, APCVk +P (∂)(Ω,E) = kerP(∂)E, domT E +m ∶= Ck(Ω,E) and +T E +m∶Ck(Ω,E) → Eωm, f �→ [(β,x) ↦ (∂β)Ef(x)], +for all m ∈ N0 and the same with K instead of E. The family (T E +m,T K +m)m∈N0 is a +strong generator for (CVk,E) because +(∂β)K(e′ ○ f)(x) = e′((∂β)Ef(x)), +(β,x) ∈ ωm, +for all e′ ∈ E′, f ∈ CVk(Ω,E) and m ∈ N0 due to the linearity and continuity of +e′ ∈ E′. In addition, e′ ○ f ∈ kerP(∂)K for all e′ ∈ E′ and f ∈ CVk +P (∂)(Ω,E), which +implies that (T E +m,T K +m)m∈N0 is also a strong generator for (CVk +P (∂),E). +For consistency we need to prove that +(∂β)ES(u)(x) = u(δx ○ (∂β)K), +(β,x) ∈ ωm, +for all u ∈ CVk(Ω)εE resp. u ∈ CVk +P (∂)(Ω)εE. This follows from the subsequent +Proposition 3.1.11 b) since FV(Ω) = CVk(Ω) resp. FV(Ω) = CVk +P (∂)(Ω) is barrelled +and Vk locally bounded away from zero on Ω. Thus (T E +m,T K +m)m∈N0 is a consistent +generator for (CVk,E). In addition, we have with P(∂)E from Example 3.1.9 g) +that +P(∂)E(S(u))(x) = +n +∑ +i=1 +ai(x)(∂βi)E(S(u))(x) = u( +n +∑ +i=1 +ai(x)(δx ○ (∂βi)K)) += u(δx ○ P(∂)K) = 0, +x ∈ Ω, +(7) +for every u ∈ CVk +P (∂)(Ω)εE. This yields S(u) ∈ kerP(∂)E for all u ∈ CVk +P (∂)(Ω)εE. +Therefore (T E +m,T K +m)m∈N0 is a consistent generator for (CVk +P (∂),E) as well. +□ + +3.1. ε-INTO-COMPATIBILITY +27 +Let us turn to the postponed part in the proof of consistency. We denote by +CW(Ω) the space of scalar-valued continuous functions on an open set Ω ⊂ Rd +with the topology of uniform convergence on compact subsets, i.e. the weighted +topology given by the family of weights W ∶= W0 ∶= {χK ∣ K ⊂ Ω compact}, and we +set δ(x) ∶= δx for x ∈ Ω. +3.1.11. Proposition. Let Ω ⊂ Rd be open, k ∈ N∞ and FV(Ω) a dom-space. +a) If T ∈ L(FV(Ω),CW(Ω)), then δ ○ T ∈ C(Ω,FV(Ω)′ +γ). +b) If T ∈ L(FV(Ω),CW1(Ω)) and FV(Ω) is barrelled, then +(∂en)FV(Ω)′ +κ(δ ○ T)(x) = lim +h→0 +δx+hen ○ T − δx ○ T +h += δx ○ (∂en)K ○ T, +x ∈ Ω, 1 ≤ n ≤ d, +and δ ○ T ∈ C1(Ω,FV(Ω)′ +κ). +c) If the inclusion I∶FV(Ω) → CWk(Ω), f ↦ f, is continuous and FV(Ω) +barrelled, then S(u) ∈ Ck(Ω,E) and +(∂β)ES(u)(x) = u(δx ○ (∂β)K), +β ∈ Nd +0, ∣β∣ ≤ k, x ∈ Ω, +for all u ∈ FV(Ω)εE. +Proof. a) First, if x ∈ Ω and (xτ)τ∈T is a net in Ω converging to x, then we +observe that +(δxτ ○ T)(f) = T(f)(xτ) → T(f)(x) = (δx ○ T)(f) +for every f ∈ FV(Ω) as T(f) is continuous on Ω. Second, let K ⊂ Ω be compact. +Then there are j ∈ J, m ∈ M and C > 0 such that +sup +x∈K +∣(δx ○ T)(f)∣ = sup +x∈K +∣T(f)(x)∣ ≤ C∣f∣j,m +for every f ∈ FV(Ω). +This means that {δx ○ T ∣ x ∈ K} is equicontinuous in +FV(Ω)′. The topologies σ(FV(Ω)′,FV(Ω)) and γ(FV(Ω)′,FV(Ω)) coincide on +equicontinuous subsets of FV(Ω)′, implying that the restriction (δ ○ T)∣K∶K → +FV(Ω)′ +γ is continuous by our first observation. As δ ○ T is continuous on every +compact subset of the open set Ω ⊂ Rd, it follows that δ ○ T∶Ω → FV(Ω)′ +γ is well- +defined and continuous. +b) Let x ∈ Ω and 1 ≤ n ≤ d. Then there is ε > 0 such that x + hen ∈ Ω for all +h ∈ R with 0 < ∣h∣ < ε. We note that δ ○ T ∈ C(Ω,FV(Ω)′ +κ) by part a), which implies +δx+hen○T −δx○T +h +∈ FV(Ω)′. For every f ∈ FV(Ω) we have +lim +h→0 +δx+hen ○ T − δx ○ T +h +(f) = lim +h→0 +T(f)(x + hen) − T(f)(x) +h += (∂en)KT(f)(x) +in K as T(f) ∈ C1(Ω). Therefore 1 +h(δx+hen ○ T − δx ○ T) converges to δx ○ (∂en)K ○ T +in FV(Ω)′ +σ and thus in FV(Ω)′ +κ by the Banach–Steinhaus theorem as well. +In +particular, we obtain +δx ○ (∂en)K ○ T = lim +h→0 +δx+hen ○ T − δx ○ T +h += (∂en)FV(Ω)′ +κ(δ ○ T)(x) +in FV(Ω)′ +κ. Moreover, δ ○ (∂en)K ○ T ∈ C(Ω,FV(Ω)′ +κ) by part a) as (∂en)K ○ T ∈ +L(FV(Ω),CW(Ω)). Hence we deduce that δ ○ T ∈ C1(Ω,FV(Ω)′ +κ). +c) We prove our claim by induction on the order of differentiation. Let u ∈ +FV(Ω)εE. For β ∈ Nd +0 with ∣β∣ = 0 we get S(u) = u ○ δ ∈ C(Ω,E) from part a) with +T = I. Further, +(∂β)ES(u)(x) = S(u)(x) = u(δx) = u(δx ○ (∂β)K), +x ∈ Ω. + +28 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +Let m ∈ N0, m < k, such that S(u) ∈ Cm(Ω,E) and +(∂β)ES(u)(x) = u(δx ○ (∂β)K), +x ∈ Ω, +(8) +for all β ∈ Nd +0 with ∣β∣ ≤ m. +Let β ∈ Nd +0 with ∣β∣ = m + 1 ≤ k. +Then there is +1 ≤ n ≤ d and ̃β ∈ Nd +0 with ∣̃β∣ = m such that β = en + ̃β. The barrelledness of +FV(Ω) yields that 1 +h(δx+hen ○ (∂ +̃β)K − δx ○ (∂ +̃β)K) converges to δx ○ (∂en)K ○ (∂ +̃β)K +in FV(Ω)′ +κ for every x ∈ Ω by part b) with T ∶= (∂ +̃β)K. Therefore we derive from +δx ○ (∂en)K ○ (∂ +̃β)K = δx ○ (∂β)K by Schwarz’ theorem that +u(δx ○ (∂β)K) = lim +h→0 +1 +h(u(δx+hen ○ (∂ +̃β)K) − u(δx ○ (∂ +̃β)K)) += +(8) lim +h→0 +1 +h((∂ +̃β)ES(u)(x + hen) − (∂ +̃β)ES(u)(x)) += (∂en)E(∂ +̃β)ES(u)(x) +for every x ∈ Ω. Moreover, δ ○ (∂β)K = (∂en)FV(Ω)′ +κ(δ ○ T) ∈ C(Ω,FV(Ω)′ +κ) for +T = (∂ +̃β)K by part b). +Hence we have S(u) ∈ Cm+1(Ω,E) and it follows from +Schwarz’ theorem again that +u(δx ○ (∂β)K) = (∂en)E(∂ +̃β)ES(u)(x) = (∂β)ES(u)(x), +x ∈ Ω. +□ +Part a) of the preceding proposition is just a modification of [16, 4.1 Lemma, p. +198], where FV(Ω) = CV(Ω) is the Nachbin-weighted space of continuous functions +and T = id, and holds more general for kR-spaces Ω (see Lemma 4.1.2). +3.1.12. Theorem. Let (T E +m,T K +m)m∈M be a consistent generator for (FV,E). +Then the map S∶FV(Ω)εE → FV(Ω,E) is an isomorphism into, i.e. the spaces +FV(Ω) and FV(Ω,E) are ε-into-compatible. +Proof. First, we show that S(FV(Ω)εE) ⊂ FV(Ω,E). Let u ∈ FV(Ω)εE. +Due to the consistency of (T E +m,T K +m)m∈M we have S(u) ∈ AP(Ω,E) ∩ domT E +m and +(T E +mS(u))(x) = u(T K +m,x), +m ∈ M x ∈ ωm. +Furthermore, we get by Lemma 3.1.8 for every j ∈ J, m ∈ M and α ∈ A +∣S(u)∣j,m,α = sup +x∈ωm +pα(T E +m(S(u))(x))νj,m(x) = ∥u∥j,m,α < ∞, +(9) +implying S(u) ∈ FV(Ω,E) and the continuity of S. Moreover, we deduce from (9) +that S is injective and that the inverse of S on the range of S is also continuous. +□ +3.1.13. Remark. If J, M and A are countable, then S is an isometry with +respect to the induced metrics on FV(Ω,E) and FV(Ω)εE by (9). +The basic idea for Theorem 3.1.12 was derived from analysing the proof of an +analogous statement for Bierstedt’s weighted spaces CV(Ω,E) and CV0(Ω,E) of +continuous functions already mentioned in the introduction (see [16, 4.2 Lemma, +4.3 Folgerung, p. 199–200] and [17, 2.1 Satz, p. 137]). +3.2. ε-compatibility +Now, we try to answer the natural question. When is S surjective? The strength +of a generator and a weaker concept to define a natural E-valued version of FV(Ω) +come into play to answer the question on the surjectivity of our key map S. Let +FV(Ω) be a dom-space. We define the linear space of E-valued weak FV-functions +by +FV(Ω,E)σ ∶= {f∶Ω → E ∣ ∀ e′ ∈ E′ ∶ e′ ○ f ∈ FV(Ω)}. + +3.2. ε-COMPATIBILITY +29 +Moreover, for f ∈ FV(Ω,E)σ we define the linear map +Rf∶E′ → FV(Ω), Rf(e′) ∶= e′ ○ f, +and the dual map +Rt +f∶FV(Ω)′ → E′⋆, f ′ �→ [e′ ↦ f ′(Rf(e′))], +where E′⋆ is the algebraic dual of E′. Furthermore, we set +FV(Ω,E)κ ∶= {f ∈ FV(Ω,E)σ ∣ ∀ α ∈ A ∶ Rf(B○ +α) relatively compact in FV(Ω)} +where Bα ∶= {x ∈ E ∣ pα(x) < 1} for α ∈ A. Next, we give a sufficient condition for +the inclusion FV(Ω,E) ⊂ FV(Ω,E)σ by means of the family (T E +m,T K +m)m∈M. +3.2.1. Lemma. If (T E +m,T K +m)m∈M is a strong generator for (FV,E), then we have +FV(Ω,E) ⊂ FV(Ω,E)σ and +sup +e′∈B○α +∣Rf(e′)∣j,m = ∣f∣j,m,α +(10) +for every f ∈ FV(Ω,E), j ∈ J, m ∈ M and α ∈ A. +Proof. Let f ∈ FV(Ω,E). +We have e′ ○ f ∈ F(Ω) for every e′ ∈ E′ since +(T E +m,T K +m)m∈M is a strong generator. Moreover, we have +∣Rf(e′)∣j,m = ∣e′ ○ f∣j,m = sup +x∈ωm +∣T K +m(e′ ○ f)(x)∣νj,m(x) += sup +x∈ωm +∣e′(T E +m(f)(x))∣νj,m(x) = +sup +x∈Nj,m(f) +∣e′(x)∣ +(11) +for every j ∈ J and m ∈ M with the set Nj,m(f) from Definition 3.1.4. We note +that Nj,m(f) is bounded in E by Definition 3.1.4 and thus weakly bounded, im- +plying that the right-hand side of (11) is finite. Hence we conclude f ∈ FV(Ω,E)σ. +Further, we observe that +sup +e′∈B○α +∣Rf(e′)∣j,m = ∣f∣j,m,α +for every j ∈ J, m ∈ M and α ∈ A due to [131, Proposition 22.14, p. 256]. +□ +Now, we phrase some sufficient conditions for FV(Ω,E) ⊂ FV(Ω,E)κ to hold +which is one of the key points regarding the surjectivity of S. +3.2.2. Lemma. If (T E +m,T K +m)m∈M is a strong generator for (FV,E) and one of +the following conditions is fulfilled, then FV(Ω,E) ⊂ FV(Ω,E)κ. +a) FV(Ω) is a semi-Montel space. +b) E is a semi-Montel or Schwartz space. +c) ∀ f ∈ FV(Ω,E), j ∈ J, m ∈ M ∃ K ∈ γ(E) ∶ Nj,m(f) ⊂ K. +Proof. Let f ∈ FV(Ω,E). By virtue of Lemma 3.2.1 we already have f ∈ +FV(Ω,E)σ. +a) For every j ∈ J, m ∈ M and α ∈ A we derive from +sup +e′∈B○α +∣Rf(e′)∣j,m = +(10) ∣f∣j,m,α < ∞ +that Rf(B○ +α) is bounded and thus relatively compact in the semi-Montel space +FV(Ω). +c) It follows from (11) that Rf ∈ L(E′ +γ,FV(Ω)). +Further, the polar B○ +α is +relatively compact in E′ +γ for every α ∈ A by the Alaoğlu–Bourbaki theorem. The +continuity of Rf implies that Rf(B○ +α) is relatively compact as well. +b) Let j ∈ J and m ∈ M. The set K ∶= Nj,m(f) is bounded in E by Definition +3.1.4. We deduce that K is already precompact in E by [89, 10.4.3 Corollary, p. +202] if E is a Schwartz space resp. since it is relatively compact if E is a semi-Montel +space. Hence the statement follows from c). +□ + +30 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +Let us turn to sufficient conditions for FV(Ω,E) ≅ FV(Ω)εE. For the lcHs E +we denote by J ∶E → E′⋆, x �→ [e′ ↦ e′(x)], the canonical injection. +3.2.3. Condition. Let (T E +m,T K +m)m∈M be a strong generator for (FV,E). Define +the following conditions: +a) E is complete. +b) E is quasi-complete and for every f ∈ FV(Ω,E) and f ′ ∈ FV(Ω)′ there is +a bounded net (f ′ +τ)τ∈T in FV(Ω)′ converging to f ′ in FV(Ω)′ +κ such that +Rt +f(f ′ +τ) ∈ J (E) for every τ ∈ T . +c) E is sequentially complete and for every f ∈ FV(Ω,E) and f ′ ∈ FV(Ω)′ +there is a sequence (f ′ +n)n∈N in FV(Ω)′ converging to f ′ in FV(Ω)′ +κ such +that Rt +f(f ′ +n) ∈ J (E) for every n ∈ N. +d) E is locally complete and for every f ∈ FV(Ω,E) and f ′ ∈ FV(Ω)′ there +is a sequence (f ′ +n)n∈N in FV(Ω)′ locally converging to f ′ in FV(Ω)′ +κ such +that Rt +f(f ′ +n) ∈ J (E) for every n ∈ N. +e) ∀ f ∈ FV(Ω,E), j ∈ J, m ∈ M ∃ K ∈ τ(E) ∶ Nj,m(f) ⊂ K. +3.2.4. Theorem. Let (T E +m,T K +m)m∈M be a consistent generator for (FV,E) and +let FV(Ω,E) ⊂ FV(Ω,E)κ. If one of the Conditions 3.2.3 is fulfilled, then the +map S∶FV(Ω)εE → FV(Ω,E) is an isomorphism, i.e. FV(Ω) and FV(Ω,E) are +ε-compatible. The inverse of S is given by the map +Rt∶FV(Ω,E) → FV(Ω)εE, f ↦ J −1 ○ Rt +f, +where J ∶E → E′⋆ is the canonical injection and +Rt +f∶FV(Ω)′ → E′⋆, f ′ �→ [e′ ↦ f ′(Rf(e′))], +with Rf(e′) = e′ ○ f. +Proof. Due to Theorem 3.1.12 we only have to show that S is surjective. We +equip J (E) with the system of seminorms given by +pB○α(J (x)) ∶= sup +e′∈B○α +∣J (x)(e′)∣ = pα(x), +x ∈ E, +(12) +for every α ∈ A. Let f ∈ FV(Ω,E). We consider the dual map Rt +f and claim that +Rt +f ∈ L(FV(Ω)′ +κ,J (E)). Indeed, we have +pB○α(Rt +f(y)) = sup +e′∈B○α +∣y(Rf(e′))∣ = +sup +x∈Rf (B○α) +∣y(x)∣ ≤ sup +x∈Kα +∣y(x)∣ +(13) +for all y ∈ FV(Ω)′ where Kα ∶= Rf(B○α). Since FV(Ω,E) ⊂ FV(Ω,E)κ, the set +Rf(B○ +α) is absolutely convex and relatively compact, implying that Kα is absolutely +convex and compact in FV(Ω) by [89, 6.2.1 Proposition, p. 103]. Further, we have +for all e′ ∈ E′ and x ∈ Ω +Rt +f(δx)(e′) = δx(e′ ○ f) = e′(f(x)) = J (f(x))(e′) +(14) +and thus Rt +f(δx) ∈ J (E). +a) Let E be complete and f ′ ∈ FV(Ω)′. Since the span of {δx ∣ x ∈ Ω} is dense +in F(Ω)′ +κ by the bipolar theorem, there is a net (f ′ +τ) converging to f ′ in FV(Ω)′ +κ +with Rt +f(f ′ +τ) ∈ J (E) by (14). As +pB○α(Rt +f(f ′ +τ) − Rt +f(f ′)) ≤ +(13) sup +x∈Kα +∣(f ′ +τ − f ′)(x)∣ → 0, +(15) +for all α ∈ A, we gain that (Rt +f(f ′ +τ)) is a Cauchy net in the complete space J (E). +Hence it has a limit g ∈ J (E) which coincides with Rt +f(f ′) since +pB○α(g − Rt +f(f ′)) ≤ pB○α(g − Rt +f(f ′ +τ)) + pB○α(Rt +f(f ′ +τ) − Rt +f(f ′)) + +3.2. ε-COMPATIBILITY +31 +≤ +(15)pB○α(g − Rt +f(f ′ +τ)) + sup +x∈Kα +∣(f ′ +τ − f ′)(x)∣ → 0 +for all α ∈ A. We conclude that Rt +f(f ′) ∈ J (E) for every f ′ ∈ FV(Ω)′. +b) Let Condition 3.2.3 b) hold and f ′ ∈ FV(Ω)′. Then there is a bounded +net (f ′ +τ)τ∈T in FV(Ω)′ converging to f ′ in FV(Ω)′ +κ such that Rt +f(f ′ +τ) ∈ J (E) for +every τ ∈ T . Due to (13) we obtain that (Rt +f(f ′ +τ)) is a bounded Cauchy net in the +quasi-complete space J (E) converging to Rt +f(f ′) ∈ J (E). +c) Let Condition 3.2.3 c) hold and f ′ ∈ FV(Ω)′. +Then there is a sequence +(f ′ +n)n∈N in FV(Ω)′ converging to f ′ in FV(Ω)′ +κ such that Rt +f(f ′ +n) ∈ J (E) for every +n ∈ N. Again (13) implies that (Rt +f(f ′ +n)) is a Cauchy sequence in the sequentially +complete space J (E) which converges to Rt +f(f ′) ∈ J (E). +d) Let Condition 3.2.3 d) hold and f ′ ∈ FV(Ω)′. Then there is an absolutely +convex, bounded subset D ⊂ FV(Ω)′ +κ and a sequence (f ′ +n)n∈N in FV(Ω)′ converging +to f ′ in (FV(Ω)′ +κ)D such that Rt +f(fn) ∈ J (E) for every n ∈ N. Let r > 0 and +f ′ +n − f ′ +k ∈ rD. Then Rt +f(f ′ +n − f ′ +k) ∈ r(Rt +f(D) ∩ J (E)), implying +{r > 0 ∣ f ′ +n − f ′ +k ∈ rD} ⊂ {r > 0 ∣ Rt +f(f ′ +n − f ′ +k) ∈ r(Rt +f(D) ∩ J (E)) +J (E)}. +Setting B ∶= Rt +f(D) ∩ J (E) +J (E), we derive +qB(Rt +f(f ′ +n − f ′ +k)) ≤ qD(f ′ +n − f ′ +k) +where qB and qD are the gauge functionals of B resp. D. The set Rt +f(D)∩J (E) is +absolutely convex as the intersection of two absolutely convex sets and it is bounded +by (13) and the boundedness of D. So B, being the closure of a disk, is a disk as +well. Since (f ′ +n) is a Cauchy sequence in (FV(Ω)′ +κ)D, we conclude that (R′ +f(f ′ +n)) +is a Cauchy sequence in J (E)B. The set B is a closed disk in the locally complete +space J (E) and hence a Banach disk by [89, 10.2.1 Proposition, p. 197]. Thus +J (E)B is a Banach space and (Rt +f(f ′ +n)) has a limit g ∈ J (E)B. The continuity +of the canonical injection J (E)B ↪ J (E) implies that (Rt +f(f ′ +n)) converges to g in +J (E) as well. As in a) we obtain that Rt +f(f ′) = g ∈ J (E). +e) Let Condition 3.2.3 e) be fulfilled. Let f ∈ FV(Ω,E) and e′ ∈ E′. For every +f ′ ∈ FV(Ω)′ there are j ∈ J, m ∈ M and C > 0 such that +∣Rt +f(f ′)(e′)∣ ≤ C∣Rf(e′)∣j,m = +(11) C +sup +x∈Nj,m(f) +∣e′(x)∣ +because (T E +m,T K +m)m∈M is a strong generator. Since there is K ∈ τ(E) such that +Nj,m(f) ⊂ K, we have +∣Rt +f(f ′)(e′)∣ ≤ C sup +x∈K +∣e′(x)∣, +implying Rt +f(f ′) ∈ (E′ +τ)′ = J (E) by the Mackey–Arens theorem. +Therefore we obtain that Rt +f ∈ L(FV(Ω)′ +κ,J (E)). So we get for all α ∈ A and +y ∈ F(Ω)′ +pα((J −1 ○ Rt +f)(y)) = +(12) pB○α(J ((J −1 ○ Rt +f)(y))) = pB○α(Rt +f(y)) ≤ +(13) sup +x∈Kα +∣y(x)∣. +This implies J −1 ○ Rt +f ∈ L(FV(Ω)′ +κ,E) = FV(Ω)εE (as linear spaces) and we gain +S(J −1 ○ Rt +f)(x) = J −1(Rt +f(δx)) = +(14) J −1(J (f(x))) = f(x) +for every x ∈ Ω. Thus S(J −1 ○ Rt +f) = f, proving the surjectivity of S. +□ +Further sufficient conditions for S being a topological isomorphism can be found +in Proposition 5.2.10, Proposition 5.6.6 and Theorem 5.7.1. In particular, we get +the following corollary as a special case of Theorem 3.2.4. + +32 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +3.2.5. Corollary. Let (T E +m,T K +m)m∈M be a strong, consistent generator for +(FV,E). If +(i) FV(Ω) is a semi-Montel space and E complete, or +(ii) FV(Ω) is a Fréchet–Schwartz space and E locally complete, or +(iii) E is a semi-Montel space, or +(iv) ∀ f ∈ FV(Ω,E), j ∈ J, m ∈ M ∃ K ∈ κ(E) ∶ Nj,m(f) ⊂ K, +then FV(Ω) and FV(Ω,E) are ε-compatible, in particluar, FV(Ω,E) ≅ FV(Ω)εE. +Proof. (i) Follows from Lemma 3.2.2 a) and Theorem 3.2.4 with Condition +3.2.3 a). +(ii) If FV(Ω) is a Fréchet–Schwartz space, then we have +span{δx ∣ x ∈ Ω} +lc = span{δx ∣ x ∈ Ω} +FV(Ω)′ +b = span{δx ∣ x ∈ Ω} +FV(Ω)′ +κ = FV(Ω)′ +by [30, Lemma 6 (b), p. 231] and the bipolar theorem where span{δx ∣ x ∈ Ω} +lc is the +local closure of span{δx ∣ x ∈ Ω} in FV(Ω)′ +b. Hence for every f ′ ∈ FV(Ω)′ there is a +sequence (f ′ +n) in the span of {δx ∣ x ∈ Ω} which converges locally to f ′ in FV(Ω)′ +κ. +Due to (14) we know that Rt +f(f ′ +n) ∈ J (E) for every f ∈ FV(Ω,E) and n ∈ N. Since +Fréchet–Schwartz spaces are also semi-Montel spaces, the statement follows from +Lemma 3.2.2 a) and Theorem 3.2.4 with Condition 3.2.3 d). +(iv) Follows from Lemma 3.2.2 c) and Theorem 3.2.4 with Condition 3.2.3 e). +(iii) Is a special case of (iv) since the set K ∶= acx(Nj,m(f)) is absolutely convex +and compact in the semi-Montel space E by [89, 6.2.1 Proposition, p. 103] and [89, +6.7.1 Proposition, p. 112] for every f ∈ FV(Ω,E), j ∈ J and m ∈ M. +□ +3.2.6. Remark. Linearisations of spaces FV(Ω,E)σ of weak E-valued func- +tions, where FV(Ω) need not be a dom-space, are treated in [118]. +Let us apply our preceding results to our weighted spaces of k-times continu- +ously partially differentiable functions on an open set Ω ⊂ Rd with k ∈ N∞. +3.2.7. Example. Let E be an lcHs, k ∈ N∞, Vk be a directed family of weights +which is locally bounded away from zero on an open set Ω ⊂ Rd. +a) CVk(Ω,E) ≅ CVk(Ω)εE if E is a semi-Montel space and CVk(Ω) barrelled. +b) CVk +P (∂)(Ω,E) ≅ CVk +P (∂)(Ω)εE if E is a semi-Montel space and CVk +P (∂)(Ω) +barrelled. +c) CVk(Ω,E) ≅ CVk(Ω)εE if E is complete and CVk(Ω) a Montel space. +d) CVk +P (∂)(Ω,E) ≅ CVk +P (∂)(Ω)εE if E is complete and CVk +P (∂)(Ω) a Montel +space. +e) CVk(Ω,E) ≅ CVk(Ω)εE if E is locally complete and CVk(Ω) a Fréchet– +Schwartz space. +f) CVk +P (∂)(Ω,E) ≅ CVk +P (∂)(Ω)εE if E is locally complete and CVk +P (∂)(Ω) a +Fréchet–Schwartz space. +Proof. The generator of (CVk,E) and (CVk +P (∂),E) is strong and consistent +by Proposition 3.1.10. From Corollary 3.2.5 (iii) we deduce part a) and b), from +(i) part c) and d) and from (ii) part e) and f). +□ +Closed subspaces of Fréchet–Schwartz spaces are also Fréchet–Schwartz spaces +by [131, Proposition 24.18, p. 284]. The space CV∞ +P (∂)(Ω) is closed in CV∞(Ω) if +there is an lcHs Y such that P(∂)∣CV∞(Ω)∶CV∞(Ω) → Y is continuous. For example, +this is fulfilled if the coefficients of P(∂) belong to C(Ω), in particular, if P(∂) ∶= ∆ +or ∂, with Y ∶= (C(Ω),τc) due to V∞ being locally bounded away from zero. The +spaces CVk(Ω) from Example 3.1.9 a)(i) with ωm ∶= Mm × Ω for all m ∈ N0, where +Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ min(m,k)}, are Fréchet spaces and thus barrelled if the J + +3.2. ε-COMPATIBILITY +33 +in Vk ∶= (νj,m)j∈J,m∈N0 is countable by [107, Proposition 3.7, p. 240]. Sufficient +conditions on the weights that guarantee that CV∞(Ω) is a nuclear Fréchet space +and hence a Schwartz space as well can be found in [111, Theorem 3.1, p. 188]. For +the case ωm = Nd +0 × Ω see the references given in [111, p. 1]. +If Vk = Wk, i.e. Ck(Ω,E) is equipped with its usual topology of uniform conver- +gence of all partial derivatives up to order k on compact subsets of Ω, Example 3.2.7 +c)+d) can be improved to quasi-complete E. For Ω = Rd this can be found in [158, +Proposition 9, p. 108, Théorème 1, p. 111] and for general open Ω ⊂ Rd it is already +mentioned in [94, (9), p. 236] (without a proof) that CWk(Ω,E) ≅ CWk(Ω)εE for +k ∈ N∞ and quasi-complete E. For k = ∞ we even have CW∞(Ω,E) ≅ CW∞(Ω)εE +for locally complete E by [30, p. 228]. Our technique allows us to generalise the +first result and to get back the second result. +3.2.8. Example. Let E be an lcHs, k ∈ N∞ and Ω ⊂ Rd open. If k < ∞ and E +has metric ccp, or if k = ∞ and E is locally complete, then +a) CWk(Ω,E) ≅ CWk(Ω)εE, and +b) CWk +P (∂)(Ω,E) ≅ CWk +P (∂)(Ω)εE if CWk +P (∂)(Ω) is closed in CWk(Ω). +Proof. We recall from Example 3.1.9 b) that Wk is the family of weights +given by νK,m(β,x) ∶= χK(x), (β,x) ∈ Mm × Ω, for all m ∈ N0 and compact K ⊂ Ω +where Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ min(m,k)} and χK is the characteristic function of K. +We already know that the generator for (CWk,E) and (CWk +P (∂),E) is strong and +consistent by Proposition 3.1.10 because Wk is locally bounded away from zero +on Ω, and CWk(Ω) and its closed subspace CWk +P (∂)(Ω) are Fréchet spaces. Let +f ∈ CWk(Ω,E), K ⊂ Ω be compact, m ∈ N0 and consider +NK,m(f) = {(∂β)Ef(x)νK,m(β,x) ∣ x ∈ Ω, β ∈ Mm} = {0} ∪ +⋃ +β∈Mm +(∂β)Ef(K). +NK,m(f) is compact since it is a finite union of compact sets. Furthermore, the +compact sets {0} and (∂β)Ef(K) are metrisable by [34, Chap. IX, §2.10, Proposi- +tion 17, p. 159] and thus their finite union NK,m(f) is metrisable as well by [169, +Theorem 1, p. 361] since the compact set NK,m(f) is collectionwise normal and +locally countably compact by [63, 5.1.18 Theorem, p. 305]. If E has metric ccp, +then the set acx(NK,m(f)) is absolutely convex and compact. Thus Corollary 3.2.5 +(iv) settles the case for k < ∞. If k = ∞ and E is locally complete, we observe that +Kβ ∶= acx((∂β)Ef(K)) for f ∈ CW∞(Ω,E) is absolutely convex and compact by +[29, Proposition 2, p. 354]. Then we have +NK,m(f) ⊂ acx( ⋃ +β∈Mm +Kβ) +and the set on the right-hand side is absolutely convex and compact by [89, 6.7.3 +Proposition, p. 113]. Again, the statement follows from Corollary 3.2.5 (iv). +□ +The statement above for k = ∞ follows from Example 3.2.7 e)+f) as well because +CW∞(Ω) and its closed subspaces are Fréchet–Schwartz spaces. In the context of +differentiability on infinite dimensional spaces the preceding example a) remains +true for an open subset Ω of a Fréchet space or DFM-space and quasi-complete +E by [129, 3.2 Corollary, p. 286]. +Like here this can be generalised to E with +[metric] ccp. A special case of example b) is already known to be a consequence of +[30, Theorem 9, p. 232], namely, if k = ∞ and P(∂)K is hypoelliptic with constant +coefficients. In particular, this covers the space of holomorphic functions and the +space of harmonic functions. Holomorphy on infinite dimensional spaces is treated +in [52, Corollary 6.35, p. 332–333] where V = W0, Ω is an open subset of a locally + +34 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +convex Hausdorff k-space and E a quasi-complete locally convex Hausdorff space, +both over C, which can be generalised to E with [metric] ccp in a similar way. +For a second improvement of Example 3.2.7 for k = ∞ to locally complete E +without the condition that CV∞(Ω) resp. CV∞ +P (∂)(Ω) is a Fréchet–Schwartz space +we introduce the following conditions on the family V∞ on (Mm × Ω)m∈N0. We say +that a family V∞ of weights on (Mm × Ω)m∈N0 is C1-controlled if +(i) ∀ j ∈ J, m ∈ N0, β ∈ Mm ∶ νj,m(β,⋅) ∈ C1(Ω), +(ii) ∀ j ∈ J, m ∈ N0, β,γ ∈ Mm,x ∈ Ω ∶ νj,m(β,x) = νj,m(γ,x), +(iii) ∀ j ∈ J, m ∈ N0 ∃ i ∈ J, k ∈ N0, k ≥ m, C > 0 ∀ β ∈ Mm, x ∈ Ω, 1 ≤ n ≤ d ∶ +∣∂enνj,m(β,⋅)∣(x) ≤ Cνi,k(β,x). +We say that family Vk, k ∈ N∞, fulfils condition (V∞) if +∀ m ∈ N0, j ∈ J ∃ n ∈ N≥m, i ∈ J ∀ ε > 0 ∃ K ⊂ Ω compact ∀ β ∈ Mm, x ∈ Ω ∖ K ∶ +νj,m(β,x) ≤ ενi,n(β,x) +where N≥m ∶= {n ∈ N0 ∣ n ≥ m}. Here (V∞) stands for vanishing at infinity and +the condition was introduced in [107, Remark 3.4, p. 239] and for k = 0 in [16, 1.3 +Bemerkung, p. 189]. +3.2.9. Example. Let E be an lcHs and V∞ a directed C1-controlled family of +weights on an open convex set Ω ⊂ Rd which fulfils (V∞). If E is locally complete, +then +a) CV∞(Ω,E) ≅ CV∞(Ω)εE if CV∞(Ω) is barrelled, and +b) CV∞ +P (∂)(Ω,E) ≅ CV∞ +P (∂)(Ω)εE if CV∞ +P (∂)(Ω) is barrelled. +Proof. We already know that the generator for (CV∞,E) and (CV∞ +P (∂),E) is +strong and consistent by Proposition 3.1.10 because V∞ is locally bounded away +from zero on Ω as νj,m(β,⋅) is continuous for all j ∈ J, m ∈ N0 and β ∈ Mm. +Let f ∈ CV∞(Ω,E), j ∈ J, m ∈ N0 and β ∈ Mm. We set g∶Ω → E, g(x) ∶= +(∂β)Ef(x)νj,m(β,x), and note that +(∂en)Eg(x) = (∂β+en)Ef(x)νj,m(β,x) + (∂β)Ef(x)((∂en)Rνj,m(β,⋅))(x), +x ∈ Ω, +for all 1 ≤ n ≤ d. +Since V∞ is directed and C1-controlled there are i1,i2 ∈ J, +k1,k2 ∈ N0, k1 > m, k2 ≥ m, and C1,C2 > 0 such that +pα((∂en)Eg(x)) +≤ pα((∂β+en)Ef(x))νj,m(β,x) + pα((∂β)Ef(x))∣(∂en)Rνj,m(β,⋅)∣(x) +≤ C1pα((∂β+en)Ef(x))νi1,k1(β,x) + C2pα((∂β)Ef(x))νi2,k2(β,x) += C1pα((∂β+en)Ef(x))νi1,k1(β + en,x) + C2pα((∂β)Ef(x))νi2,k2(β,x) +for all 1 ≤ n ≤ d and α ∈ A, which implies +sup +x∈Ω +γ∈Nd +0,∣γ∣≤1 +pα((∂γ)Eg(x)) ≤ ∣f∣j,m,α + C1∣f∣i1,k1,α + C2∣f∣i2,k2,α. +Thus g is (weakly) C1 +b . +Due to (V∞) there are n ∈ N≥m and i ∈ J such that for all ε > 0 there is a +compact set K ⊂ Ω such that for all β ∈ Mm and x ∈ Ω ∖ K we have +νj,m(β,x) ≤ ενi,n(β,x). +Since V∞ is directed, we may assume w.l.o.g. that νj,m(β,x) ≤ νi,n(β,x) for all +x ∈ Ω. This implies that the zeros of νi,n(β,⋅) are zeros of νj,m(β,⋅). We define +h∶Ω → [0,∞) by h(x) ∶= νi,n(β,x)/νj,m(β,x) for x ∈ Ω with νj,m(β,x) ≠ 0 and + +3.2. ε-COMPATIBILITY +35 +h(x) ∶= 1 if νj,m(β,x) = 0. We note that h(x) > 0 for all x ∈ Ω as the zeros of +νi,n(β,⋅) are contained in the zeros of νj,m(β,⋅). It follows that +(∂β)Ef(x)νj,m(β,x)h(x) = (∂β)Ef(x)νi,n(β,x) +for x ∈ Ω with νj,m(β,x) ≠ 0 and (∂β)Ef(x)νj,m(β,x)h(x) = 0 for x ∈ Ω with +νj,m(β,x) = 0. Therefore (∂β)Efνj,m(β,⋅)h is bounded on Ω. Further, +εh(x) = ενi,n(β,x)/νj,m(β,x) ≥ 1 +for x ∈ Ω ∖ K with νj,m(β,x) ≠ 0 because (V∞) is fulfilled. Further, the zeros of +νj,m(β,⋅) are contained in N ∶= {x ∈ Ω ∣ (∂β)Ef(x)νj,m(β,x) = 0}. This yields that +Kβ ∶= acx((∂β)Efνj,m(β,⋅)(Ω)) is absolutely convex and compact by Proposition +A.1.4 and A.1.5. Furthermore, +Nj,m(f) = {(∂β)Ef(x)νj,m(β,x) ∣ x ∈ Ω, β ∈ Mm} ⊂ acx( ⋃ +β∈Mm +Kβ) +and the set on the right-hand side is absolutely convex and compact by [89, 6.7.3 +Proposition, p. 113]. Finally, our statement follows from Corollary 3.2.5 (iv). +□ +For the Schwartz space S(Rd,E) and the multiplier space OM(Rd,E) from +Example 3.1.9 c) and d) an improvement of Example 3.2.7 c) to quasi-complete E +is already known, see e.g. [158, Proposition 9, p. 108, Théorème 1, p. 111]. However, +due to Example 3.2.9 it is even allowed that E is only locally complete. +3.2.10. Corollary. If E is a locally complete lcHs, then S(Rd,E) ≅ S(Rd)εE +and OM(Rd,E) ≅ OM(Rd)εE. +Proof. We start with the Schwartz space. Due to Example 3.2.9 a) and the +barrelledness of the Fréchet space S(Rd) we only need to check that its directed +family V∞ ∶= (ν1,m)m∈N0 of weights given by ν1,m(β,x) ∶= (1 + ∣x∣2)m/2, x ∈ Rd, for +m ∈ N0 and β ∈ Mm is C1-controlled and fulfils (V∞). Obviously, condition (i) and +(ii) are fulfilled. Since +∣∂enν1,m(β,⋅)∣(x) = (m/2)(1 + ∣x∣2)(m/2)−12∣xn∣ ≤ m(1 + ∣x∣2)m/2 = mν1,m(β,x) +for all x ∈ Rd and 1 ≤ n ≤ d, condition (iii) is also fulfilled. Thus V∞ is C1-controlled. +Noting that for every m ∈ N and ε > 0 there is r > 0 such that +(1 + ∣x∣2)m/2 +(1 + ∣x∣2)m = (1 + ∣x∣2)−m/2 ≤ ε +for all x ∉ Br(0), we obtain that +ν1,m(β,x) ≤ εν1,2m(β,x) +for all x ∉ Br(0) and β ∈ Mm. Hence V∞ fulfils condition (V∞). +Now, let us consider the multiplier space. We already know that the generator +for (OM,E) is strong and consistent by Proposition 3.1.10 because OM(R) is a +Montel space, thus barrelled, by [83, Chap. II, §4, n○4, Théorème 16, p. 131] and +its family of weights is continuous on Rd, thus locally bounded away from zero. +Let f ∈ OM(R,E), g ∈ S(Rd), m ∈ N0 and β ∈ Mm. Then (∂β)Ef ∈ OM(R,E) +and hence ((∂β)Ef)g ∈ S(Rd,E), which implies that ((∂β)Ef)g ∈ C1 +b (Rd,E). More- +over, we choose h∶Rd → (0,∞), h(x) ∶= 1 + ∣x∣2. Then ((∂β)Ef)gh is bounded on +Rd and for ε > 0 there is r > 0 such that (1 + ∣x∣2)−1 ≤ ε for all x ∉ Br(0), yielding +that Kβ,g ∶= acx(((∂β)Ef)g(Rd)) is absolutely convex and compact by Proposi- +tion A.1.4 and A.1.5. +Let j ⊂ S(Rd) be finite. +Since for each x ∈ Rd we have +(∂β)Ef(x)maxg∈j ∣g(x)∣ = eiθ(∂β)Ef(x)̃g(x) for some ̃g ∈ j and θ ∈ [0,2π), we get +Nj,m(f) = {(∂β)Ef(x)max +g∈j ∣g(x)∣ ∣ x ∈ Rd, β ∈ Mm} ⊂ acx( +⋃ +β∈Mm,g∈j +Kβ,g). + +36 +3. THE ε-PRODUCT FOR WEIGHTED FUNCTION SPACES +The set on the right-hand side is absolutely convex and compact by [89, 6.7.3 +Proposition, p. 113]. Finally, our statement follows from Corollary 3.2.5 (iv). +□ +For an alternative proof in the case of the Schwartz space we may also use +Example 3.2.7 e) since S(Rd) is a Fréchet–Schwartz space. +Example 3.2.9 can +also be used for an alternative proof of Example 3.2.8 if k = ∞ by observing that +CW∞(Ω,E) = CV∞(Ω,E) for any lcHs E where V∞ ∶= {ν ∈ C∞ +c (Ω) ∣ ν ≥ 0} and +C∞ +c (Ω) is the space of functions in C∞(Ω) with compact support. +Now, we improve Example 3.2.7 for the special case of spaces of ultradifferen- +tiable functions E(Mp)(Ω,E) and E{Mp}(Ω,E) from Example 3.1.9 e) and f) where +ωm ∶= Nd +0 × Ω for all m ∈ N0. For this purpose we recall the following conditions of +Komatsu for the sequence (Mp)p∈N0 (see [99, p. 26] and [101, p. 653]): +(M.0) M0 = M1 = 1, +(M.1) ∀ p ∈ N ∶ M 2 +p ≤ Mp−1Mp+1, +(M.2)’ ∃ A,C > 0 ∀ p ∈ N0 ∶ Mp+1 ≤ ACp+1Mp, +(M.3)’ ∑∞ +p=1 +Mp−1 +Mp < ∞. +3.2.11. Example. Let E be an lcHs, Ω ⊂ Rd open and (Mp)p∈N0 a sequence of +positive real numbers. +a) E(Mp)(Ω,E) ≅ E(Mp)(Ω)εE if E is locally complete. +b) E{Mp}(Ω,E) ≅ E{Mp}(Ω)εE if E is complete or semi-Montel and in both +cases (Mp)p∈N0 fulfils (M.1) and (M.3)’. +c) E{Mp}(Ω,E) ≅ E{Mp}(Ω)εE if E is sequentially complete and (Mp)p∈N0 +fulfils (M.0), (M.1), (M.2)’ and (M.3)���. +Proof. The generator is strong and consistent by Proposition 3.1.10 since the +family of weights given in Example 3.1.9 e) resp. f) is locally bounded away from +zero on Ω and E(Mp)(Ω) is a Fréchet–Schwartz space in a) by [99, Theorem 2.6, +p. 44] whereas E{Mp}(Ω) is a Montel space in b) and c) by [99, Theorem 5.12, p. +65–66]. Hence the statements a) and b) follow from Example 3.2.7. +Let us turn to c). We note that E{Mp}(Ω,E) ⊂ E{Mp}(Ω,E)κ by Lemma 3.2.2 +a) for any lcHs E. +Further, we claim that Condition 3.2.3 c) is fulfilled. +Let +f ′ ∈ E{Mp}(Ω)′. Due to [101, Proposition 3.7, p. 677] there is a sequence (fn)n∈N in +the space D{Mp}(Ω) of ultradifferentiable functions of class {Mp} of Roumieu-type +with compact support which converges to f ′ in E{Mp}(Ω)′ +b. Let f ∈ E{Mp}(Ω,E). +We observe that for every e′ ∈ E′ +∣Rt +f(fn)(e′)∣ = ∣∫ +Ω +fn(x)e′(f(x))dx∣ ≤ λ(supp(fn)) +sup +y∈Kn(f) +∣e′(y)∣ +where λ is the Lebesgue measure, supp(fn) is the support of fn and Kn(f) ∶= +{fn(x)f(x) ∣ x ∈ supp(fn)}. The set Kn(f) is compact and metrisable by [34, +Chap. IX, §2.10, Proposition 17, p. 159] and thus the closure of its absolutely +convex hull is compact in E as the sequentially complete space E has metric ccp. +We conclude that Rt +f(fn) ∈ (E′ +κ)′ = J (E) for every n ∈ N. Therefore Condition +3.2.3 c) is fulfilled, implying statement c) for sequentially complete E by Theorem +3.2.4. +□ +The results a) and b) in this example are new whereas c) is already proved in +[101, Theorem 3.10, p. 678] in a different way. In particular, part a) improves [101, +Theorem 3.10, p. 678] since Komatsu’s conditions (M.0), (M.1), (M.2)’ and (M.3)’ +are not needed and the condition that E is sequentially complete is weakened to +local completeness. We included c) to demonstrate an application of Condition +3.2.3 c). + +CHAPTER 4 +Consistency +4.1. The spaces AP(Ω,E) and consistency +This section is dedicated to the properties of functions which are compatible +with the ε-product in the sense that the space of functions having these properties +can be chosen as the space AP(Ω,E) or ⋂m∈M domT E +m in the Definition 3.1.7 b) of +consistency. This is done in a quite general way so that we are not tied to certain +spaces and have to redo our argumentation, for example, if we consider the same +generator (T E +m,T K +m)m∈M for two different spaces of functions. +Due to the linearity and continuity of u ∈ FV(Ω)εE for a dom-space FV(Ω) +and S(u) = u ○ δ with δ∶Ω → FV(Ω)′, x ↦ δx, these are properties which are purely +pointwise or given by pointwise approximation. Among such properties of func- +tions are continuity by Proposition 4.1.1, Cauchy continuity by Proposition 4.1.3, +uniform continuity by Proposition 4.1.5, continuous extendability by Proposition +4.1.7, continuous differentiability by Proposition 3.1.10, vanishing at infinity by +Proposition 4.1.9 and purely pointwise properties of a function like vanishing on a +set by Proposition 4.1.10. +We collect these properties in propositions and in follow-up lemmas we handle +properties which can be described by compositions of defining operators T E +m1 ○ T E +m2 +like continuous differentiability (of higher order) of Fourier transformations (see +Example 4.2.26). We fix the following notation for this section. For a dom-space +FV(Ω) and linear T∶FV(Ω) → KΩ we set (δ ○ T)(x)(f) ∶= (δx ○ T)(f) ∶= T(f)(x) +for all x ∈ Ω and f ∈ FV(Ω). +4.1.1. Proposition (continuity). Let Ω be a topological Hausdorff space and +FV(Ω) a dom-space such that FV(Ω) ⊂ C(Ω) as a linear subspace. Then S(u) ∈ +C(Ω,E) for all u ∈ FV(Ω)εE if δ ∈ C(Ω,FV(Ω)′ +κ). +Proof. Let u ∈ FV(Ω)εE. Since S(u) = u○δ and δ ∈ C(Ω,FV(Ω)′ +κ), we obtain +that S(u) is in C(Ω,E). +□ +Now, we tackle the problem of the continuity of δ∶Ω → FV(Ω)′ +κ in the proposi- +tion above and phrase our solution in a way such that it can be applied to show the +continuity of the partial derivative (∂β)E(S(u)) as well (see Proposition 3.1.11). +We recall that a topological space Ω is called completely regular if for any non-empty +closed subset A ⊂ Ω and x ∈ Ω ∖ A there is f ∈ C(Ω,[0,1]) such that f(x) = 0 and +f(z) = 1 for all z ∈ A (see [88, Definition 11.1, p. 180]). Examples of completely +regular spaces are uniformisable, particularly metrisable, spaces by [88, Proposition +11.5, p. 181] and locally convex Hausdorff spaces by [65, Proposition 3.27, p. 95]. +A completely regular space Ω is a kR-space if for any completely regular space Y +and any map f∶Ω → Y , whose restriction to each compact K ⊂ Ω is continuous, +the map is already continuous on Ω (see [37, (2.3.7) Proposition, p. 22]). Examples +of kR-spaces are completely regular k-spaces by [63, 3.3.21 Theorem, p. 152]. A +topological space Ω is called k-space (compactly generated space) if it satisfies the +following condition: A ⊂ Ω is closed if and only if A ∩ K is closed in K for every +compact K ⊂ Ω. Every locally compact Hausdorff space is a completely regular +37 + +38 +4. CONSISTENCY +k-space. Further, every sequential Hausdorff space is a k-space by [63, 3.3.20 The- +orem, p. 152], in particular, every first-countable Hausdorff space. Thus metrisable +spaces are completely regular Hausdorff k-spaces. Moreover, the dual space (X′,τc) +with the topology of compact convergence τc is an example of a completely regular +Hausdorff k-space that is neither locally compact nor metrisable by [178, p. 267] if +X is an infinite-dimensional Fréchet space. +We denote by CW(Ω) the space of scalar-valued continuous functions on a +topological Hausdorff space Ω with the topology τc of compact convergence, i.e. the +topology of uniform convergence on compact subsets, which itself is the weighted +topology given by the family of weights W ∶= W0 ∶= {χK ∣ K ⊂ Ω compact}, and +by Cb(Ω) the space of scalar-valued bounded, continuous functions on Ω with the +topology of uniform convergence on Ω. +4.1.2. Lemma. Let Ω be a topological Hausdorff space, FV(Ω) a dom-space +and T∶FV(Ω) → C(Ω) linear. Then δ ○T ∈ C(Ω,FV(Ω)′ +γ) in each of the subsequent +cases: +(i) Ω is a kR-space and T∶FV(Ω) → CW(Ω) is continuous. +(ii) T∶FV(Ω) → Cb(Ω) is continuous. +Proof. First, if x ∈ Ω and (xτ)τ∈T is a net in Ω converging to x, then we +observe that +(δxτ ○ T)(f) = T(f)(xτ) → T(f)(x) = (δx ○ T)(f) +for every f ∈ FV(Ω) as T(f) is continuous on Ω. +(i) Verbatim as in Proposition 3.1.11 a). +(ii) There are j ∈ J, m ∈ M and C > 0 such that +sup +x∈Ω +∣(δx ○ T)(f)∣ = sup +x∈Ω +∣T(f)(x)∣ ≤ C∣f∣FV(Ω),j,m +for every f ∈ FV(Ω). This means that {δx ○T ∣ x ∈ Ω} is equicontinuous in FV(Ω)′, +yielding the statement like before. +□ +The preceding lemma is just a modification of [16, 4.1 Lemma, p. 198] where +FV(Ω) = CV(Ω), the Nachbin-weighted space of continuous functions, and T = id. +Next, we turn to Cauchy continuity. A function f∶Ω → E from a metric space +Ω to an lcHs E is called Cauchy continuous if it maps Cauchy sequences to Cauchy +sequences. We write CC(Ω,E) for the space of Cauchy continuous functions from +Ω to E and set CC(Ω) ∶= CC(Ω,K). +4.1.3. Proposition (Cauchy continuity). Let Ω be a metric space and FV(Ω) +a dom-space such that FV(Ω) ⊂ CC(Ω) as a linear subspace. Then S(u) ∈ CC(Ω,E) +for all u ∈ FV(Ω)εE if δ ∈ CC(Ω,FV(Ω)′ +κ). +Proof. Let u ∈ FV(Ω)εE and (xn) a Cauchy sequence in Ω. Then (δxn) is a +Cauchy sequence in FV(Ω)′ +κ since δ ∈ CC(Ω,FV(Ω)′ +κ). It follows that (S(u)(xn)) is +a Cauchy sequence in E because u is uniformly continuous and u(δxn) = S(u)(xn). +Hence we conclude that S(u) ∈ CC(Ω,E). +□ +For the next lemma we equip the space CC(Ω) with the topology of uniform +convergence on precompact subsets of Ω. +4.1.4. Lemma. Let FV(Ω) be a dom-space and T ∈ L(FV(Ω),CC(Ω)) for a +metric space Ω. Then δ ○ T ∈ CC(Ω,FV(Ω)′ +γ). +Proof. Let (xn) be a Cauchy sequence in Ω. We have (δxn○T)(f) = T(f)(xn) +for every f ∈ FV(Ω), which implies that ((δxn ○ T)(f)) is a Cauchy sequence in +K because T(f) ∈ CC(Ω) by assumption. +Since K is complete, it has a unique +limit T∞(f) ∶= limn→∞(δxn ○ T)(f) defining a linear functional in f. The set N ∶= + +4.1. THE SPACES AP(Ω, E) AND CONSISTENCY +39 +{xn ∣ n ∈ N} is precompact in Ω since Cauchy sequences are precompact. Hence +there are j ∈ J, m ∈ M and C > 0 such that +sup +n∈N +∣(δxn ○ T)(f)∣ = sup +x∈N +∣T(f)(x)∣ ≤ C∣f∣FV(Ω),j,m +for every f ∈ FV(Ω). Therefore the set {δxn○T ∣ n ∈ N} is equicontinuous in FV(Ω)′, +which implies that T∞ ∈ FV(Ω)′ and the convergence of (δxn ○T) to T∞ in FV(Ω)′ +γ +due to the observation in the beginning and the fact that γ(FV(Ω)′,FV(Ω)) and +σ(FV(Ω)′,FV(Ω)) coincide on equicontinuous sets. In particular, (δxn ○ T) is a +Cauchy sequence in FV(Ω)′ +γ. Furthermore, for every x ∈ Ω we obtain from the +choice xn = x for all n ∈ N that δx ○ T ∈ FV(Ω)′. Thus the map δ ○ T∶Ω → FV(Ω)′ +γ +is well-defined and Cauchy continuous. +□ +The subsequent proposition and lemma handle the analogous statements for +uniform continuity. +For a metric space Ω we denote by Cu(Ω,E) the space of +uniformly continuous functions from Ω to E and set Cu(Ω) ∶= Cu(Ω,K). +4.1.5. Proposition (uniform continuity). Let (Ω,d) be a metric space and +FV(Ω) a dom-space such that FV(Ω) ⊂ Cu(Ω) as a linear subspace. Then S(u) ∈ +Cu(Ω,E) for all u ∈ FV(Ω)εE if δ ∈ Cu(Ω,FV(Ω)′ +κ).1 +Proof. Let (zn), (xn) be sequences in Ω with limn→∞ d(zn,xn) = 0 and u ∈ +FV(Ω)εE. Then (δzn −δxn) converges to 0 in FV(Ω)′ +κ because δ ∈ Cu(Ω,FV(Ω)′ +κ). +As a consequence (S(u)(zn) − S(u)(xn)) converges to 0 in E since u is uniformly +continuous and u(δzn −δxn) = S(u)(zn)−S(u)(xn). Hence we conclude that S(u) ∈ +Cu(Ω,E). +□ +For the next lemma we mean by Cbu(Ω) the space of scalar-valued bounded, +uniformly continuous functions equipped with the topology of uniform convergence +on a metric space Ω. +4.1.6. Lemma. Let FV(Ω) be a dom-space and T ∈ L(FV(Ω),Cbu(Ω)) for a +metric space (Ω,d). Then δ ○ T ∈ Cu(Ω,FV(Ω)′ +γ). +Proof. Let (zn) and (xn) be sequences in Ω such that limn→∞ d(zn,xn) = 0. +We have +(δzn ○ T − δxn ○ T)(f) = T(f)(zn) − T(f)(xn) +for every f ∈ FV(Ω), which implies that (δzn ○ T − δxn ○ T)(f) converges to 0 in K +for every f ∈ FV(Ω) because T(f) ∈ Cu(Ω). There exist j ∈ J, m ∈ M and C > 0 +such that +sup +n∈N +∣(δzn ○ T − δxn ○ T)(f)∣ ≤ 2sup +x∈Ω +∣T(f)(x)∣ ≤ 2C∣f∣FV(Ω),j,m +for every f ∈ FV(Ω). Therefore the set {δzn ○ T − δxn ○ T ∣ n ∈ N} is equicontinuous +in FV(Ω)′ and we conclude the statement like before. +□ +Let us turn to continuous extensions. Let X be a metric space and Ω ⊂ X. We +write Cext(Ω,E) for the space of functions f ∈ C(Ω,E) which have a continuous +extension to Ω and set Cext(Ω) ∶= Cext(Ω,K). +4.1.7. Proposition (continuous extendability). Let X be a metric space, Ω ⊂ X +and FV(Ω) a dom-space such that FV(Ω) ⊂ Cext(Ω) as a linear subspace. Then +S(u) ∈ Cext(Ω,E) for all u ∈ FV(Ω)εE if δ ∈ Cext(Ω,FV(Ω)′ +κ). +1Here, we use the symbol u for elements in FV(Ω)εE instead of the usual u to avoid confusion +with the index u of Cu(Ω) resp. Cu(Ω, E). + +40 +4. CONSISTENCY +Proof. Let u ∈ FV(Ω)εE. There is δext ∈ C(Ω,FV(Ω)′ +κ) such that δext = δ on +Ω since δ ∈ Cext(Ω,FV(Ω)′ +κ). Moreover, u○δext ∈ C(Ω,E) and equal to S(u) = u○δ +on Ω, yielding S(u) ∈ Cext(Ω,E). +□ +For the next lemma we equip Cext(Ω) with the topology of uniform convergence +on compact subsets of Ω. +4.1.8. Lemma. Let X be a metric space, Ω ⊂ X, FV(Ω) a dom-space and +T ∈ L(FV(Ω),Cext(Ω)). Then δ ○ T ∈ Cext(Ω,FV(Ω)′ +γ) if FV(Ω) is barrelled. +Proof. From Lemma 4.1.2 (i) we derive that δ○T ∈ C(Ω,FV(Ω)′ +γ). Let x ∈ ∂Ω +and (xn) be a sequence in Ω with xn → x. Then (δxn ○ T) is a sequence in FV(Ω)′ +and +lim +n→∞(δxn ○ T)(f) = lim +n→∞T(f)(xn) =∶ (δext +x +○ T)(f) +in K for every f ∈ FV(Ω), which implies that (δxn○T) converges to δext +x +○T pointwise +on FV(Ω) because T(f) ∈ Cext(Ω). As a consequence of the Banach–Steinhaus +theorem we get (δext +x +○ T) ∈ FV(Ω)′ and the convergence in FV(Ω)′ +γ. +□ +Let FV(Ω,E) be a dom-space, X a set, K a family of sets and π∶⋃m∈M ωm → X +such that ⋃K∈K K ⊂ X. +We say that a function f ∈ ⋂m∈M domT E +m vanishes at +infinity in the weighted topology w.r.t. (π,K) if +∀ ε > 0, j ∈ J, m ∈ M, α ∈ A ∃ K ∈ K ∶ +sup +x∈ωm, +π(x)∉K +pα(T E +m(f)(x))νj,m(x) < ε. +(16) +Further, we set +APπ,K(Ω,E) ∶= {f ∈ ⋂ +m∈M +domT E +m ∣ f fulfils (16)}. +4.1.9. Proposition (vanishing at ∞ w.r.t. to (π,K)). Let (T E +m,T K +m)m∈M be +the generator for (FV,E), let FV(Ω,Y ) ⊂ APπ,K(Ω,Y ) as a linear subspace for +Y ∈ {K,E} and K be closed under taking finite unions. +(i) If for all u ∈ FV(Ω)εE it holds that S(u) ∈ ⋂m∈M dom(T E +m) and +∀ m ∈ M, x ∈ ωm ∶ (T E +mS(u))(x) = u(T K +m,x), +(17) +then S(u) ∈ APπ,K(Ω,E) for all u ∈ FV(Ω)εE. +(ii) If for all e′ ∈ E′ and f ∈ FV(Ω,E) it holds that e′ ○ f ∈ ⋂m∈M dom(T K +m) +and +∀ m ∈ M, x ∈ ωm ∶ T K +m(e′ ○ f)(x) = (e′ ○ T E +m(f))(x), +(18) +then e′ ○ f ∈ APπ,K(Ω) for all e′ ∈ E′ and f ∈ FV(Ω,E). +Proof. (i) We set Bj,m ∶= {f ∈ FV(Ω) ∣ ∣f∣j,m ≤ 1} for j ∈ J and m ∈ M. Let +u ∈ FV(Ω)εE. The topologies σ(FV(Ω)′,FV(Ω)) and κ(FV(Ω)′,FV(Ω)) coincide +on the equicontinuous set B○ +j,m and we deduce that the restriction of u to B○ +j,m is +σ(FV(Ω)′,FV(Ω))-continuous. +Let ε > 0, j ∈ J, m ∈ M, α ∈ A and set Uα,ε ∶= {x ∈ E ∣ pα(x) < ε}. Then there +are a finite set N ⊂ FV(Ω) and η > 0 such that u(f ′) ∈ Uα,ε for all f ′ ∈ VN,η where +VN,η ∶= {f ′ ∈ FV(Ω)′ ∣ sup +f∈N +∣f ′(f)∣ < η} ∩ B○ +j,m +because the restriction of u to B○ +j,m is σ(FV(Ω)′,FV(Ω))-continuous. Since N ⊂ +FV(Ω) is finite, FV(Ω) ⊂ APπ,K(Ω) and K is closed under taking finite unions, +there is K ∈ K such that +sup +x∈ωm +π(x)∉K +∣T K +m(f)(x)∣νj,m(x) < η +(19) + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +41 +for every f ∈ N. It follows from (19) and (the proof of) Lemma 3.1.8 that +Dπ⊄K,j,m ∶= {T K +m,x(⋅)νj,m(x) ∣ x ∈ ωm, π(x) ∉ K} ⊂ VN,η +and thus u(Dπ⊄K,j,m) ⊂ Uα,ε. Therefore we have +sup +x∈ωm +π(x)∉K +pα(T E +m(S(u))(x))νj,m(x) = +(17) +sup +x∈ωm +π(x)∉K +pα(u(T K +m,x))νj,m(x) < ε. +Hence we conclude that S(u) ∈ APπ,K(Ω,E). +(ii) Let ε > 0, f ∈ FV(Ω,E) and e′ ∈ E′. Then there exist α ∈ A and C > 0 such +that ∣e′(x)∣ ≤ Cpα(x) for every x ∈ E. For j ∈ J and m ∈ M there is K ∈ K such that +sup +x∈ωm +π(x)∉K +pα(T E +m(f)(x))νj,m(x) < ε +C +since FV(Ω,E) ⊂ APπ,K(Ω,E). It follows that +sup +x∈ωm +π(x)∉K +∣T K +m(e′ ○ f)(x)∣νj,m(x) = +(18) +sup +x∈ωm +π(x)∉K +∣e′(T E +m(f)(x))∣νj,m(x) < C ε +C = ε, +yielding e′ ○ f ∈ APπ,K(Ω). +□ +The first part of the proof above adapts an idea in the proof of [16, 4.4 Theo- +rem, p. 199–200] where (T E +m,T K +m)m∈M = (idEΩ,idKΩ) which is a special case of our +proposition. +Our last proposition of this section is immediate. For ω ⊂ Ω we set APω(Ω,E) ∶= +{f ∈ EΩ ∣ ∀ x ∈ ω ∶ f(x) = 0} and APω(Ω) ∶= APω(Ω,K). +4.1.10. Proposition (vanishing on a subset). Let ω ⊂ Ω and FV(Ω) a dom- +space such that FV(Ω) ⊂ APω(Ω) as a linear subspace. Then S(u) ∈ APω(Ω,E) +for all u ∈ FV(Ω)εE. +4.2. Further examples of ε-products +In Chapter 3 we dealt with weighted spaces of continuously partially differen- +tiable functions. Now, we treat many examples of weighted spaces FV(Ω,E) of +functions with less regularity on a set Ω with values in a locally convex Hausdorff +space E over the field K. Applying the results of the preceding sections, we give +conditions on E such that FV(Ω) and FV(Ω,E) are ε-compatible, in particular, +that +FV(Ω,E) ≅ FV(Ω)εE +holds. We start with the simplest example of all. Let Ω be a non-empty set and +equip the space EΩ with the topology of pointwise convergence, i.e. the locally +convex topology given by the seminorms +∣f∣K,α ∶= sup +x∈K +pα(f(x))χK(x), +f ∈ EΩ, +for finite K ⊂ Ω and α ∈ A. To prove EN0 ≅ KN0εE for complete E is given as an +exercise in [94, Aufgabe 10.5, p. 259], which we generalise now. +4.2.1. Example. Let Ω be a non-empty set and E an lcHs. Then EΩ ≅ KΩεE. +Proof. The strength and consistency of the generator (idEΩ,idKΩ) is obvious. +Let f ∈ EΩ, K ⊂ Ω be finite and set NK(f) ∶= f(Ω)χK(Ω). Then we have NK(f) = +f(K) ∪ {0} if K ≠ Ω, and NK(f) = f(K) if K = Ω. Thus NK(f) is finite, hence +compact, NK(f) ⊂ acx(f(K)) and acx(f(K)) is a subset of the finite dimensional +subspace span(f(K)) of E. It follows that acx(f(K)) is compact by [89, 6.7.4 +Proposition, p. 113], implying our statement by virtue of Corollary 3.2.5 (iv). +□ + +42 +4. CONSISTENCY +The next example will give us the counterpart of Example 3.2.9 a) on the level +of sequence spaces. Let Ω be a set, E an lcHs and V ∶= (νj)j∈J a directed family of +weights νj∶Ω → [0,∞) on Ω. We set +ℓV(Ω,E) ∶= {f ∈ EΩ ∣ ∀ j ∈ J, α ∈ A ∶ ∣f∣j,α ∶= sup +x∈Ω +pα(f(x))νj(x) < ∞} +and ℓV(Ω) ∶= ℓV(Ω,K). +4.2.2. Example. Let E be an lcHs, (Ω,d) a uniformly discrete metric space, +i.e. there is r > 0 such that d(x,y) ≥ r for all x,y ∈ Ω, x ≠ y, and V ∶= (νj)j∈J a +directed family of weights on Ω such that +∀ j ∈ J ∃ i ∈ J ∀ ε > 0 ∃ K ⊂ Ω compact ∀ x ∈ Ω ∖ K ∶ νj(x) ≤ ενi(x). +(20) +If E is locally complete, then ℓV(Ω,E) ≅ ℓV(Ω)εE. +Proof. Let f ∈ ℓV(Ω,E) and j ∈ J. Then fνj is bounded on Ω by definition +of ℓV(Ω,E). Since (Ω,d) is uniformly discrete, there is r > 0 such that +pα(f(x)νj(x) − f(y)νj(y)) +d(x,y) +≤ 2 +r ∣f∣j,α < ∞, +x,y ∈ Ω, x ≠ y, +for every α ∈ A. Therefore fνj ∈ C[1] +b (Ω,E) where +C[1] +b (Ω,E) ∶= {g ∈ EΩ ∣ ∀α ∈ A ∶ sup +x∈Ω +pα(g(x)) < ∞ and sup +x,y∈Ω +x≠y +pα(g(x) − g(y)) +d(x,y) +< ∞}. +Due to (20) there is i ∈ J such that for all ε > 0 there exists a compact set K ⊂ Ω +such that νj(x) ≤ ενi(x) for all x ∈ Ω∖K. As V is directed, we may assume w.l.o.g. +that νj(x) ≤ νi(x) for all x ∈ Ω. This implies that the zeros of νi are zeros of νj. We +define h∶Ω → [0,∞) by h(x) ∶= νi(x)/νj(x) for x ∈ Ω with νj(x) ≠ 0 and h(x) ∶= 1 +if νj(x) = 0. We observe that h(x) > 0 for all x ∈ Ω as the zeros of νi are contained +in the zeros of νj. It follows that +f(x)νj(x)h(x) = f(x)νi(x) +for x ∈ Ω with νj(x) ≠ 0 and f(x)νj(x)h(x) = 0 for x ∈ Ω with νj(x) = 0. Hence +fνjh is bounded on Ω. Further, +εh(x) = ενi(x)/νj(x) ≥ 1, +for x ∈ Ω ∖ K with νj(x) ≠ 0 because (20) is fulfilled. Moreover, the zeros of νj +are contained in N ∶= {x ∈ Ω ∣ f(x)νj(x) = 0}. This yields that acx(fνj(Ω)) is +absolutely convex and compact by Proposition A.1.4. +So our statement follows +from Corollary 3.2.5 (iv). +□ +Let us apply the preceding result to some known sequence spaces. We recall +that a matrix A ∶= (ak,j)k,j∈N of non-negative numbers is called Köthe matrix if it +fulfils: +(1) ∀ k ∈ N ∃ j ∈ N ∶ ak,j > 0, +(2) ∀ k,j ∈ N ∶ ak,j ≤ ak,j+1. +We note that what we call k is usually called j and vice-versa (see e.g. [131, Def- +inition, p. 326]). But the notation we chose is more in line with the meaning of +j in our Definition 3.1.3 of a weight function and therefore we prefer to keep our +notation consistent. For an lcHs E we define the Köthe space +λ∞(A,E) ∶= {x = (xk) ∈ EN ∣ ∀ j ∈ N, α ∈ A ∶ ∣x∣j,α ∶= sup +k∈N +pα(xk)ak,j < ∞} + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +43 +and the spaces of E-valued rapidly decreasing sequences which we need for some +theorems on Fourier expansions (see Theorem 5.6.13, Theorem 5.6.14) by +s(Ω,E) ∶= {x = (xk) ∈ EΩ ∣ ∀ j ∈ N, α ∈ A ∶ ∣x∣j,α ∶= sup +k∈Ω +pα(xk)(1 + ∣k∣2)j/2 < ∞} +with Ω = Nd, Nd +0, Zd. Further, we set λ∞(A) ∶= λ∞(A,K) and s(Ω) ∶= s(Ω,K). +4.2.3. Corollary. Let E be a locally complete lcHs. +a) If A ∶= (ak,j)k,j∈N is a Köthe matrix such that +∀ j ∈ N ∃ i ∈ N ∀ ε > 0 ∃ K ∈ N ∀ k ∈ N, k > K ∶ ak,j ≤ εak,i, +(21) +then λ∞(A,E) ≅ λ∞(A)εE. +b) s(Ω,E) ≅ s(Ω)εE for Ω = Nd, Nd +0, Zd. +Proof. We observe that N and Ω are uniformly discrete metric spaces if they +are equipped with the metric induced by the absolute value. Further, a set in a +discrete space is compact if and only if it is finite. In case b) we set νj∶Ω → (0,∞), +νj(k) ∶= (1 + ∣k∣2)j/2 for j ∈ N. Then for ε > 0 there is K ∈ N such that +(1 + ∣k∣2)j/2 +(1 + ∣k∣2)j += (1 + ∣k∣2)−j/2 ≤ ε +for all k ∈ Ω with ∣k∣ > K. In both cases the family of weights are directed, in case +a) due to condition (2) of the definition of a Köthe matrix. Hence we can apply +Example 4.2.2 in both cases. +□ +Due to [131, Proposition 27.10, p. 330–331] condition (21) is equivalent to +λ∞(A) being a Schwartz space. +Since λ∞(A) is also a Fréchet space by [131, +Lemma 27.1, p. 326], another way to prove Corollary 4.2.3 a) (and b) as well) is +given by Corollary 3.2.5 (ii). +Our next examples are Favard-spaces. Let E be an lcHs, 0 < γ ≤ 1, Ω a compact +Hausdorff space, ϕ∶[0,∞) × Ω → Ω a continuous semiflow, i.e. +ϕ(t + r,s) = ϕ(t,ϕ(r,s)) +and +ϕ(0,s) = s, +t,r ∈ [0,∞), s ∈ Ω, +and (̃T E +t )t≥0 the induced semigroup given by ̃T E +t ∶C(Ω,E) → C(Ω,E), ̃T E +t (f) ∶= +f(ϕ(t,⋅)). The semigroup (̃T K +t )t≥0 is (equi-)bounded and strongly continuous by +[62, Chap. II, 3.31 Exercises (1), p. 95]. The vector-valued Favard space of order γ +of the semigroup (̃T E +t )t≥0 is defined by +Fγ(Ω,E) ∶= {f ∈ C(Ω,E) ∣ ∀ α ∈ A ∶ +sup +x∈Ω,t>0 +pα(̃T E +t (f)(x) − f(x))t−γ < ∞} +equipped with the system of seminorms given by +∣f∣α ∶= max(sup +x∈Ω +pα(f(x)), sup +x∈Ω,t>0 +pα(̃T E +t (f)(x) − f(x))t−γ), +f ∈ Fγ(Ω,E), +for α ∈ A (see [39, Definition 3.1.2, p. 160] and [39, Proposition 3.1.3, p. 160]). +Further, we set Fγ(Ω) ∶= Fγ(Ω,K). Fγ(Ω,E) is a dom-space, which follows from +the setting ω ∶= [0,∞) × Ω, domT E ∶= C(Ω,E) and T E∶C(Ω,E) → Eω given by +T E(f)(0,x) ∶= f(x) and T E(f)(t,x) ∶= ̃T E +t (f)(x) − f(x), +t > 0, x ∈ Ω, +as well as AP(Ω,E) ∶= EΩ and the weight given by ν(0,x) ∶= 1 and ν(t,x) ∶= t−γ +for t > 0 and x ∈ Ω. +4.2.4. Example. Let E be a semi-Montel space, 0 < γ ≤ 1, Ω a compact Haus- +dorff space, ϕ∶[0,∞) × Ω → Ω a continuous semiflow. Then Fγ(Ω,E) ≅ Fγ(Ω)εE +holds for the Favard space of order γ of the induced semigroup (̃T E +t )t≥0. + +44 +4. CONSISTENCY +Proof. The generator (T E,T K) for (Fγ,E) is consistent by Proposition 4.1.1 +and Lemma 4.1.2 b)(ii). Its strength is clear. Thus our statement follows from +Corollary 3.2.5 (iii). +□ +The space of càdlàg functions on a set Ω ⊂ R with values in an lcHs E is defined +by +D(Ω,E) ∶= {f ∈ EΩ ∣ ∀ x ∈ Ω ∶ +lim +w→x+f(w) = f(x) and lim +w→x−f(w) exists}.2 +Further, we set D(Ω) ∶= D(Ω,K). Due to Proposition A.1.1 the maps given by +∣f∣K,α ∶= sup +x∈Ω +pα(f(x))χK(x), +f ∈ D(Ω,E), +for compact K ⊂ Ω and α ∈ A form a system of seminorms inducing a locally convex +Hausdorff topology on D(Ω,E). +4.2.5. Example. Let E be an lcHs and Ω ⊂ R locally compact. If E is quasi- +complete, then D(Ω)εE ≅ D(Ω,E). +Proof. First, we show that the generator (idEΩ,idKΩ) for (D,E) is strong +and consistent. The strength is a consequence of a simple calculation, so we only +prove the consistency explicitly. +We have to show that S(u) ∈ D(Ω,E) for all +u ∈ D(Ω)εE. Let x ∈ Ω be an accumulation point of [x,∞) ∩ Ω resp. (−∞,x] ∩ Ω, +(xn) be a sequence in Ω such that xn → x+ resp. xn → x−. We have +δxn(f) = f(xn) → f(x) = δx(f), +xn → x+, +and +δxn(f) = f(xn) → lim +n→∞f(xn) =∶ T(f)(x), +xn → x−, +for every f ∈ D(Ω), which implies that (δxn) converges to δx if xn → x+, and to +δx ○ T if xn → x− in D(Ω)′ +σ. +Since Ω is locally compact, there are a compact +neighbourhood U(x) ⊂ Ω of x and n0 ∈ N such that xn ∈ U(x) for all n ≥ n0. Hence +we deduce +sup +n≥n0 +∣δxn(f)∣ ≤ ∣f∣U(x) +for every f ∈ D(Ω). Therefore the set {δxn ∣ n ≥ n0} is equicontinuous in D(Ω)′, +which implies that (δxn) converges to δx if xn → x+ and to δx ○ T if xn → x− in +D(Ω)′ +γ and thus in D(Ω)′ +κ. From +S(u)(x) = u(δx) = lim +n→∞u(δxn) = lim +n→∞S(u)(xn), +xn → x+, +and +u(δx ○ T) = lim +n→∞u(δxn) = lim +n→∞S(u)(xn), +xn → x−, +for every u ∈ D(Ω)εE follows the consistency. Second, let f ∈ D(Ω,E), K ⊂ Ω be +compact and consider NK(f) = f(Ω)χK(Ω). We observe that NK(f) = f(K)∪{0} +if K ≠ Ω, and NK(f) = f(K) if K = Ω. We note that NK(f) ⊂ acx(f(K)) and +acx(f(K)) is absolutely convex and compact by Proposition A.1.1 because E is +quasi-complete. Thus we derive our statement from Corollary 3.2.5 (iv). +□ +We turn to Cauchy continuous functions. Let Ω be a metric space, E an lcHs +and the space CC(Ω,E) of Cauchy continuous functions from Ω to E be equipped +with the system of seminorms given by +∣f∣K,α ∶= sup +x∈K +pα(f(x))χK(x), +f ∈ CC(Ω,E), +for K ⊂ Ω precompact and α ∈ A. +2We note that for x ∈ Ω we only demand limw→x+ f(w) = f(x) if x is an accumulation point of +[x, ∞)∩Ω, and the existence of the limit limw→x− f(w) if x is an accumulation point of (−∞, x]∩Ω. + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +45 +4.2.6. Example. Let E be an lcHs and Ω a metric space. If E is a Fréchet +space or a semi-Montel space, then CC(Ω,E) ≅ CC(Ω)εE. +Proof. The generator (idEΩ,idKΩ) for (CC,E) is consistent by Proposition +4.1.3 with Lemma 4.1.4. Its strength follows from the uniform continuity of every +e′ ∈ E′. First, we consider the case that E is a Fréchet space. Let f ∈ CC(Ω,E), K ⊂ +Ω be precompact and consider NK(f) = f(Ω)χK(Ω). Then NK(f) = f(K) ∪ {0} if +K ≠ Ω, and NK(f) = f(K) if K = Ω. The set f(K) is precompact in the metrisable +space E by [13, Proposition 4.11, p. 576]. Thus we obtain CC(Ω,E) ⊂ CC(Ω,E)κ +by virtue of Lemma 3.2.2 c). Since E is complete, the first part of the statement +follows from Theorem 3.2.4 with Condition 3.2.3 a). If E is a semi-Montel space, +then it is a consequence of Corollary 3.2.5 (iii). +□ +Let (Ω,d) be a metric space, E an lcHs and the space Cbu(Ω,E) of bounded +uniformly continuous functions from Ω to E be equipped with the system of semi- +norms given by +∣f∣α ∶= sup +x∈Ω +pα(f(x)), +f ∈ Cbu(Ω,E), +for α ∈ A. +4.2.7. Example. Let E be an lcHs and (Ω,d) a metric space. If E is a semi- +Montel space, then Cbu(Ω,E) ≅ Cbu(Ω)εE. +Proof. The generator (idEΩ,idKΩ) for (Cbu,E) is consistent by Proposition +4.1.5 with Lemma 4.1.6. It is also strong due to the uniform continuity of every +e′ ∈ E′, yielding our statement by Corollary 3.2.5 (iii). +□ +4.2.8. Remark. If N is equipped with the metric induced by the absolut value, +then Cbu(N,E) = ℓ∞(N,E) where ℓ∞(N,E) is the space of bounded E-valued +sequences. If E is a separable infinite-dimensional Hilbert space, then the map +S∶Cbu(N)εE → Cbu(N,E) is not surjective by [17, 2.8 Beispiel, p. 140] and [94, Satz +10.5, p. 235–236]. Hence one cannot drop the condition that E is a semi-Montel +space in Example 4.2.7. +Let (Ω,d) be a metric space, z ∈ Ω, E an lcHs, 0 < γ ≤ 1 and define the space +of E-valued γ-Hölder continuous functions on Ω that vanish at z by +C[γ] +z (Ω,E) ∶= {f ∈ EΩ ∣ f(z) = 0 and ∀ α ∈ A ∶ ∣f∣α < ∞} +where +∣f∣α ∶= sup +x,w∈Ω +x≠w +pα(f(x) − f(w)) +d(x,w)γ +. +The topological subspace C[γ] +z,0(Ω,E) of γ-Hölder continuous functions that vanish +at infinity consists of all f ∈ C[γ] +z (Ω,E) such that for all ε > 0 there is δ > 0 with +sup +x,w∈Ω +0 0}, and let π∶ω1 → ω1 be the identity. Then +C[γ] +z,0(Ω,E) = C[γ] +z (Ω,E) ∩ APπ,K(Ω,E) with APπ,K(Ω,E) from Proposition 4.1.9 +and the generator (T E +1 ,T K +1 ) for (C[γ] +z,0,E) is strong and consistent by Proposition +4.1.10 for vanishing at z and Proposition 4.1.9 for vanishing at infinity w.r.t. (π,K). +Let f ∈ C[γ] +z,0(Ω,E) and Kδ ∶= {(x,w) ∈ Ω2 ∣ d(x,w) ≥ δ} for δ > 0. For +Nπ⊂Kδ,1,1(f) = {T E +1 (f)(x,w)ν1,1(x,w) ∣ (x,w) ∈ Kδ} = { f(x)−f(w) +d(x,w)γ +∣ (x,w) ∈ Kδ} +we have +Nπ⊂Kδ,1,1(f) ⊂ δ−γ{c(f(x) − f(w)) ∣ x,w ∈ Ω, ∣c∣ ≤ 1} += δ−γ ch(f(Ω) − f(Ω)). +The set f(Ω) is precompact because Ω is precompact and the γ-Hölder continuous +function f is uniformly continuous. It follows that the linear combination f(Ω) − +f(Ω) is precompact and the circled hull of a precompact set is still precompact by +[153, Chap. I, 5.1, p. 25]. Therefore Nπ⊂Kδ,1,1(f) is precompact for every δ > 0, +giving the precompactness of +N1,1(f) = {T E +1 (f)(x,w)ν1,1(x,w) ∣ (x,w) ∈ ω1} +by Proposition A.1.6. Hence statement b) is a consequence of Corollary 3.2.5 (iv), +Proposition A.1.6 and the quasi-completeness of E. +□ +Let Ω be a topological Hausdorff space and V ∶= (νj)j∈J a directed family of +weights νj∶Ω → [0,∞). +The weighted space of continuous functions on Ω with +values in an lcHs E is given by +CV(Ω,E) ∶= {f ∈ C(Ω,E) ∣ ∀ j ∈ J, α ∈ A ∶ ∣f∣j,α < ∞} +where +∣f∣j,α ∶= sup +x∈Ω +pα(f(x))νj(x). +Its topological subspace of functions that vanish at infinity in the weighted topology +is defined by +CV0(Ω,E) ∶= {f ∈ CV(Ω,E) ∣ ∀ j ∈ J, α ∈ A, ε > 0 +∃ K ⊂ Ω compact ∶ ∣f∣Ω∖K,j,α < ε} +where +∣f∣Ω∖K,j,α ∶= sup +x∈Ω∖K +pα(f(x))νj(x). +Further, we define CV(Ω) ∶= CV(Ω,K) and CV0(Ω) ∶= CV(Ω,K). In particular, we +set Cb(Ω,E) ∶= CV(Ω,E), i.e. the space of bounded continuous functions, and have +CV0(Ω,E) = C0(Ω,E) if V ∶= {1}. In [15, 16, 17] Bierstedt studies these spaces in the +case that V is a Nachbin-family which means that the functions νj are upper semi- +continuous for all j ∈ J and directed in the sense that for j1,j2 ∈ J and λ ≥ 0 there +is j3 ∈ J such that λνj1,λνj2 ≤ νj3. Formally this is stronger than our definition +of being directed in Remark 3.1.6 c). The notion U ≤ V for two Nachbin-families + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +47 +means that for every µ ∈ U there is ν ∈ V such that µ ≤ ν. One of his main results +from [17] is the following theorem. +4.2.10. Theorem ([17, 2.4 Theorem (2), p. 138–139]). Let E be a quasi- +complete lcHs, Ω a completely regular Hausdorff space and V a Nachbin-family +on Ω. If +(i) Z ∶= {v∶Ω → R ∣ v constant, v ≥ 0} ≤ V, or +(ii) ̃ +W ∶= {µχK ∣ µ > 0, K ⊂ Ω compact} ≤ V and Ω is a kR-space, +then CV0(Ω,E) ≅ CV0(Ω)εE. +We note that C̃ +W(Ω,E) = CW(Ω,E) with our definition of W = {χK ∣ K ⊂ +Ω compact} from above Lemma 4.1.2. +The only difference is that W is not a +Nachbin-family because it is not directed in the sense of Nachbin-families but in +the sense of Remark 3.1.6 c). We improve this result by strengthening the conditions +on Ω and V which allows us to weaken the assumptions on E. +4.2.11. Example. Let E be an lcHs, Ω a locally compact topological Hausdorff +space and V a directed family of continuous weights on Ω. +(i) If E has ccp, or +(ii) if E has metric ccp and Ω is second-countable, +then CV0(Ω,E) ≅ CV0(Ω)εE. +Proof. We set K ∶= {K ⊂ Ω ∣ K compact} and π∶Ω → Ω, π(x) ∶= x. It follows +from Proposition 4.1.1 combined with Lemma 4.1.2 (i) (continuity) and Proposition +4.1.9 (vanish at infinity w.r.t. (π,K)) that the generator (idEΩ,idKΩ) is strong and +consistent since V is a family of continuous weights and Ω a kR-space due to local +compactness. +Let f ∈ CV0(Ω,E), j ∈ J and consider Nj(f) = (fνj)(Ω). +By Proposition +A.1.3 the set K ∶= acx(Nj(f)) is absolutely convex and compact as fνj ∈ C0(Ω,E), +implying our statement by Corollary 3.2.5 (iv). +□ +4.2.12. Example. Let E be an lcHs and Ω a [metrisable] kR-space. If E has +[metric] ccp, then CW(Ω,E) ≅ CW(Ω)εE. +Proof. First, we observe that the generator (idEΩ,idKΩ) for (CW,E) is con- +sistent by Proposition 4.1.1 and Lemma 4.1.2 b)(i). Its strength is obvious. Let +f ∈ CW(Ω,E), K ⊂ Ω be compact and consider NK(f) = f(Ω)νK(Ω). +Then +NK(f) = f(K) ∪ {0} if K ≠ Ω, and NK(f) = f(K) if K = Ω, which yields that +NK(f) is compact in E. If Ω is even metrisable, then f(K) is also metrisable by +[34, Chap. IX, §2.10, Proposition 17, p. 159] and thus the finite union NK(f) as +well by [169, Theorem 1, p. 361] since the compact set NK(f) is collectionwise +normal and locally countably compact by [63, 5.1.18 Theorem, p. 305]. Further, +acx(NK(f)) is absolutely convex and compact in E if E has ccp resp. if Ω is metris- +able and E has metric ccp. We conclude that CW(Ω,E) ≅ CW(Ω)εE if E has ccp +resp. if Ω is metrisable and E has metric ccp by Corollary 3.2.5 (iv). +□ +Bierstedt also considers closed subspaces of CV(Ω) and CV0(Ω), for instance +subspaces of holomorpic functions on open Ω, and of holomorpic functions on the +inner points of Ω which are continuous on the boundary in [17, 3.1 Bemerkung, p. +141] and [17, 3.7 Satz, p. 144]. +Let Ω ⊂ C be open and bounded and E an lcHs over C. We denote by A(Ω,E) +the space of continuous functions from Ω to an lcHs E which are holomorphic on +Ω and equip A(Ω,E) with the system of seminorms given by +∣f∣α ∶= sup +x∈Ω +pα(f(x)), +f ∈ A(Ω,E), + +48 +4. CONSISTENCY +for α ∈ A. We set A(Ω) ∶= A(Ω,C), J ∶= M ∶= {1} and ν1,1 ∶= 1 on Ω. +4.2.13. Example. Let E be an lcHs and Ω ⊂ C open and bounded. +Then +A(Ω,E) ≅ A(Ω)εE if E has metric ccp. +Proof. The space A(Ω) is a Banach space and hence barrelled. The inclusion +I∶A(Ω) → CW∞ +∂ (Ω) is continuous due to the Cauchy inequality (I is an inclusion +due to the identity theorem). It follows from Proposition 4.1.1, Lemma 4.1.2 b)(i), +Proposition 3.1.11 c) and (4) that the generator (idEΩ,idCΩ) is consistent and as +in Proposition 3.1.10 that it is strong, too. +Let f ∈ A(Ω,E) and N1,1(f) = f(Ω). The set K ∶= acx(N1,1(f)) is absolutely +convex and compact by Proposition A.1.3 since f ∈ C(Ω,E) = C0(Ω,E), implying +our statement by Corollary 3.2.5 (iv). +□ +For quasi-complete E this is already covered by [17, 3.1 Bemerkung, p. 141]. +More general than holomorphic functions, we may also consider kernels of hypoel- +liptic linear partial differential operators in CV(Ω) and CV0(Ω). For an open set +Ω ⊂ Rd, a directed family V ∶= (νj)j∈N of weights νj∶Ω → [0,∞), an lcHs E and a +linear partial differential operator P(∂)E which is hypoelliptic if E = K we define +the space of zero solutions +CVP (∂)(Ω,E) ∶= {f ∈ C∞ +P (∂)(Ω,E) ∣ ∀ j ∈ N, α ∈ A ∶ ∣f∣j,α < ∞}, +where C∞ +P (∂)(Ω,E) is the kernel of P(∂)E in C∞(Ω,E), +∣f∣j,α ∶= sup +x∈Ω +pα(f(x))νj(x), +and its topological subspace +CV0,P (∂)(Ω,E) ∶= CVP (∂)(Ω,E) ∩ CV0(Ω,E). +Further, we set CVP (∂)(Ω) ∶= CVP (∂)(Ω,K) and CV0,P (∂)(Ω) ∶= CV0,P (∂)(Ω,K). We +say that V is locally bounded away from zero on Ω if +∀ K ⊂ Ω compact ∃ j ∈ N ∶ inf +x∈K νj(x) > 0. +This is an extension of the definition of being locally bounded away from zero from +Vk with k ∈ N∞ to the case k = 0 (see Proposition 3.1.10). If V is a Nachbin-family, +this means that ̃ +W ≤ V (see Theorem 4.2.10 (ii)). +4.2.14. Proposition. Let Ω ⊂ Rd be open, V ∶= (νj)j∈N an increasing family +of weights which is locally bounded away from zero on Ω and P(∂)K a hypoelliptic +linear partial differential operator. Then CVP (∂)(Ω) and CV0,P (∂)(Ω) are Fréchet +spaces. +Proof. We note that CVP (∂)(Ω) is metrisable as V is countable. Let (fn) be +a Cauchy sequence in CVP (∂)(Ω). From V being locally bounded away from zero +it follows that for every compact K ⊂ Ω there is j ∈ N such that +sup +x∈K +∣f(x)∣ ≤ sup +z∈K +νj(z)−1 sup +x∈K +∣f(x)∣νj(x) ≤ sup +z∈K +νj(z)−1∣f∣j, +f ∈ CVP (∂)(Ω), +(22) +which means that the inclusion I∶CVP (∂)(Ω) → CWP (∂)(Ω) is continuous. Thus +(fn) is also a Cauchy sequence in CWP (∂)(Ω) and has a limit f there as CWP (∂)(Ω) +is complete due to the hypoellipticity of P(∂)K. Let j ∈ N, ε > 0 and x ∈ Ω. Then +there is mj,ε,x ∈ N such that for all m ≥ mj,ε,x it holds that +∣fm(x) − f(x)∣ < +ε +2νj(x) + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +49 +if νj(x) ≠ 0. Further, there is mj,ε ∈ N such that for all n,m ≥ mj,ε it holds that +∣fn − fm∣j < ε +2. +Hence for n ≥ mj,ε we choose m ≥ max(mj,ε,mj,ε,x) and derive +∣fn(x)−f(x)∣νj(x) ≤ ∣fn(x)−fm(x)∣νj(x)+∣fm(x)−f(x)∣νj(x) < ε +2+ +ε +2νj(x)νj(x) = ε. +It follows that ∣fn − f∣j ≤ ε and ∣f∣j ≤ ε + ∣fn∣j for all n ≥ mj,ε, implying the +convergence of (fn) to f in CVP (∂)(Ω). Therefore CVP (∂)(Ω) is a Fréchet space. +CV0,P (∂)(Ω) is a closed subspace of CVP (∂)(Ω) and so a Fréchet space as well. +□ +Due to the proposition above the spaces CVP (∂)(Ω) and CV0,P (∂)(Ω) are closed +subspaces of CV(Ω) resp. CV0(Ω). Hence we have the following consequence of +Theorem 4.2.10 (ii), [17, 2.12 Satz (1), p. 141] and [17, 3.1 Bemerkung, p. 141]. +4.2.15. Corollary. Let E be an lcHs, Ω ⊂ Rd open, V a Nachbin-family on +Ω which is locally bounded away from zero and P(∂)K a hypoelliptic linear partial +differential operator. +a) CVP (∂)(Ω,E) ≅ CVP (∂)(Ω)εE if E is a semi-Montel space. +b) CV0,P (∂)(Ω,E) ≅ CV0,P (∂)(Ω)εE if E is quasi-complete. +Like before we may improve this result by strengthening the conditions on V +and CVP (∂)(Ω) resp. CV0,P (∂)(Ω) which allows us to weaken the assumptions on +E. +4.2.16. Example. Let E be an lcHs, Ω ⊂ Rd open, V ∶= (νj)j∈N an increasing +family of weights which is locally bounded away from zero on Ω and P(∂)K a +hypoelliptic linear partial differential operator. +a) CVP (∂)(Ω,E) ≅ CVP (∂)(Ω)εE if E is complete and CVP (∂)(Ω) a semi- +Montel space. +b) CVP (∂)(Ω,E) ≅ CVP (∂)(Ω)εE if E is locally complete and CVP (∂)(Ω) a +Schwartz space. +c) CV0,P (∂)(Ω,E) ≅ CV0,P (∂)(Ω)εE if E has metric ccp and νj ∈ C(Ω) for all +j ∈ N. +d) CV0,P (∂)(Ω,E) ≅ CV0,P (∂)(Ω)εE if E is locally complete and CV0,P (∂)(Ω) +a Schwartz space. +Proof. Let F stand for CVP (∂) or CV0,P (∂). The space F(Ω) is a Fréchet space +and hence barrelled by Proposition 4.2.14. The inclusion I∶F(Ω) → CWP (∂)(Ω) is +continuous since V is locally bounded away from zero on Ω. The hypoellipticity +of P(∂)K (see e.g. [70, p. 690]) yields that CWP (∂)(Ω) = CW∞ +P (∂)(Ω) as locally +convex spaces. Thus the inclusion I∶F(Ω) → CW∞ +P (∂)(Ω) is continuous. It follows +from Proposition 3.1.11 c) that the generator (idEΩ,idKΩ) is consistent if F = +CVP (∂), and combined with Proposition 4.1.9 (vanish at infinity w.r.t. (π,K)) if +F = CV0,P (∂) where K and π are chosen as in Example 4.2.11. The strength of the +generator follows as in Proposition 3.1.10 and, if F = CV0,P (∂), in combination with +Proposition 4.1.9 b). This proves part a), b) and d) due to Corollary 3.2.5 (i) and +(ii). +Let us turn to part c). Let f ∈ CV0,P (∂)(Ω,E), j ∈ N and Nj(f) ∶= (fνj)(Ω). +The set K ∶= acx(Nj(f)) is absolutely convex compact by Proposition A.1.3 as +fνj ∈ C0(Ω,E), implying our statement by Corollary 3.2.5 (iv). +□ +At least for some weights and operators P(∂) we can show that CVP (∂)(Ω,E) +coincides with a corresponding space CV∞ +P (∂)(Ω,E) from Example 3.1.9 if E is +locally complete. + +50 +4. CONSISTENCY +4.2.17. Proposition. Let E be a locally complete lcHs, Ω ⊂ Rd and P(∂)K +a hypoelliptic linear partial differential operator. Then we have CWP (∂)(Ω)εE ≅ +CWP (∂)(Ω,E) and CWP (∂)(Ω,E) = CW∞ +P (∂)(Ω,E) as locally convex spaces. +Proof. We already know that +SCW∞ +P (∂)(Ω)∶CW∞ +P (∂)(Ω)εE → CW∞ +P (∂)(Ω,E) +is an isomorphism by Example 3.2.8 b). +The hypoellipticity of P(∂)K (see e.g. +[70, p. 690]) yields that CWP (∂)(Ω)εE = CW∞ +P (∂)(Ω)εE. Thus SCWP (∂)(Ω)(u) = +SCW∞ +P (∂)(Ω)(u) ∈ C∞ +P (∂)(Ω,E) for all u ∈ CWP (∂)(Ω)εE. In particular, we obtain +that +SCWP (∂)(Ω)∶CWP (∂)(Ω)εE → CW∞ +P (∂)(Ω,E) +is an isomorphism. From Proposition 3.1.11 c) and Theorem 3.1.12 with (T E,T K) ∶= +(idEΩ,idKΩ) we deduce that +SCWP (∂)(Ω)∶CWP (∂)(Ω)εE → CWP (∂)(Ω,E) +is an isomorphism into, and from +SCWP (∂)(Ω)(CWP (∂)(Ω)εE) = C∞ +P (∂)(Ω,E) +that CWP (∂)(Ω,E) = CW∞ +P (∂)(Ω,E) as locally convex spaces, which proves our +statement. +□ +Hence the topology τc of compact convergence induced by C(Ω,E) and the +usual topology from Example 3.1.2 induced by C∞(Ω,E) coincide on CP (∂)(Ω,E) if +P(∂)K is hypoelliptic and E locally complete by Proposition 4.2.17. In particular, +we have +(O(Ω,E),τc) = +(4) CW∂(Ω,E) = CW∞ +∂ (Ω,E) +(23) +if E is locally complete. For more interesting weights than W we introduce the +following condition. +4.2.18. Condition. Let V ∶= (νj)j∈N be an increasing family of continuous +weights on Rd. Let there be r∶Rd → (0,1] and for any j ∈ N let there be ψj ∈ L1(Rd), +ψj > 0, and N ∋ Im(j) ≥ j and Am(j) > 0, m ∈ {1,2,3}, such that for any x ∈ Rd: +(α.1) supζ∈Rd, ∥ζ∥∞≤r(x) νj(x + ζ) ≤ A1(j)infζ∈Rd, ∥ζ∥∞≤r(x) νI1(j)(x + ζ), +(α.2) νj(x) ≤ A2(j)ψj(x)νI2(j)(x), +(α.3) νj(x) ≤ A3(j)r(x)νI3(j)(x). +Here, ∥ζ∥∞ ∶= sup1≤n≤d ∣ζn∣ for ζ = (ζn) ∈ Rd. The preceding condition is a +special case of [111, Condition 2.1, p. 176] with Ω ∶= Ωn ∶= Rd for all n ∈ N. If V +fulfils Condition 4.2.18 and we set V∞ ∶= (νj,m)j∈N,m∈N0 where νj,m∶{β ∈ Nd +0 ∣ ∣β∣ ≤ +m} × Rd → [0,∞), νj,m(β,x) ∶= νj(x), then CV∞(Rd) and its closed subspace +CV∞ +P (∂)(Rd) for P(∂) with continuous coefficients are nuclear by [111, Theorem +3.1, p. 188] in combination with [111, Remark 2.7, p. 178–179] and Fréchet spaces +by [107, Proposition 3.7, p. 240]. +4.2.19. Proposition. Let E be a locally complete lcHs, V ∶= (νj)j∈N an in- +creasing family of continuous weights on Rd and V∞ defined as above. If V ful- +fils Condition 4.2.18, then CV∂(C) and CV∆(Rd) are nuclear Fréchet spaces and +CV∂(C,E) = CV∞ +∂ (C,E) and CV∆(Rd,E) = CV∞ +∆(Rd,E) as locally convex spaces. +Proof. Let P(∂) ∶= ∂ (d ∶= 2 and K ∶= C) or P(∂) ∶= ∆. First, we show that +CVP (∂)(Rd) = CV∞ +P (∂)(Rd) as locally convex spaces, which implies that CVP (∂)(Rd) +is a nuclear Fréchet space as CV∞ +P (∂)(Rd) is such a space. Let f ∈ CV∂(C), j ∈ N, + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +51 +m ∈ N0, z ∈ C and β ∶= (β1,β2) ∈ N2 +0. Then it follows from ∥ ⋅ ∥∞ ≤ ∣ ⋅ ∣ and Cauchy’s +inequality that +∣∂βf(z)∣νj(z) = +(5) ∣iβ2∂∣β∣ +C f(z)∣νj(z) ≤ +∣β∣! +r(z)∣β∣ +sup +∣w−z∣=r(z) +∣f(w)∣νj(z) +≤ +(α.3)∣β∣!C(j,∣β∣) +sup +∣w−z∣=r(z) +∣f(w)∣νB3(j)(z) +≤ +(α.1)∣β∣!C(j,∣β∣)A1(B3(j)) +sup +∣w−z∣=r(z) +∣f(w)∣νI1B3(j)(w) +≤ ∣β∣!C(j,∣β∣)A1(B3(j))∣f∣CV∂(C),I1B3(j) +where C(j,∣β∣) ∶= A3(j)A3(I3(j))⋯A3((B3 − 1)(j)) and B3 − 1 is the (∣β∣ − 1)-fold +composition of I3. Choosing k ∶= max∣β∣≤m I1B3(j), it follows that +∣f∣CV∞ +∂ (C),j,m ≤ sup +∣β∣≤m +∣β∣!C(j,∣β∣)A1(B3(j))∣f∣CV∂(C),k < ∞ +and thus f ∈ CV∞ +∂ (C) and CV∂(C) = CV∞ +∂ (C) as locally convex spaces. +In the +case P(∂) = ∆ an analogous proof works due to Cauchy’s inequality for harmonic +functions, i.e. for all f ∈ CV∆(Rd), j ∈ N, x ∈ Rd and β ∈ Nd +0 it holds that +∣∂βf(x)∣νj(x) ≤ ( d∣β∣ +r(x)) +∣β∣ +sup +∣w−x∣ 0 +∃ K ⊂ Ω compact ∶ ∣f∣Ω∖K,j,m,α < ε} + +52 +4. CONSISTENCY +where +∣f∣Ω∖K,j,m,α ∶= sup +x∈Ω∖K +β∈Mm +pα((∂β)Ef(x))νj,m(β,x). +Further, we define its subspace CVk +0,P (∂)(Ω,E) ∶= {f ∈ CVk +0(Ω,E) ∣ f ∈ kerP(∂)E} +where +P(∂)E∶Ck(Ω,E) → EΩ, P(∂)E(f)(x) ∶= +n +∑ +i=1 +ai(x)(∂βi)E(f)(x), +with n ∈ N, βi ∈ Nd +0 such that ∣βi∣ ≤ k and ai∶Ω → K for 1 ≤ i ≤ n. +4.2.21. Remark. If Vk fulfils condition (V∞) from Example 3.2.9, then we have +CVk +0(Ω,E) = CVk(Ω,E) (see [107, Remark 3.4, p. 239]). +So CWk(Ω,E), S(Rd,E) and OM(Rd,E) are concrete examples of spaces +CVk +0(Ω,E) (see Corollary 3.2.10). +We present the counterpart for differentiable +functions to Bierstedt’s Theorem 4.2.10 for the space CV0(Ω,E) of continuous func- +tions from a completely regular Hausdorff space Ω to an lcHs E weighted with a +Nachbin-family V that vanish at infinity in the weighted topology. For this purpose +we need the following definition. We call Vk locally bounded on Ω if +∀ K ⊂ Ω compact, j ∈ J, m ∈ N0, β ∈ Mm ∶ sup +x∈K +νj,m(β,x) < ∞. +4.2.22. Example. Let E be an lcHs, k ∈ N∞, Vk be a directed family of weights +which is locally bounded away from zero on an open set Ω ⊂ Rd. +a) CVk +0(Ω,E) ≅ CVk +0(Ω)εE if E is quasi-complete, Vk locally bounded and +CVk +0(Ω) barrelled. +b) CVk +0(Ω,E) ≅ CVk +0(Ω)εE if E has metric ccp, CVk +0(Ω) is barrelled and +νj,m(β,⋅) ∈ C(Ω) for all j ∈ J, m ∈ N0, β ∈ Nd +0, ∣β∣ ≤ min(m,k). +c) CVk +0(Ω,E) ≅ CVk +0(Ω)εE if E is locally complete and CVk +0(Ω) a Fréchet– +Schwartz space. +d) CVk +0,P (∂)(Ω,E) ≅ CVk +0,P (∂)(Ω)εE if E is quasi-complete, Vk loc. bounded +and CVk +0,P (∂)(Ω) barrelled. +e) CVk +0,P (∂)(Ω,E) ≅ CVk +0,P (∂)(Ω)εE if E has metric ccp, CVk +0,P (∂)(Ω) is bar- +relled and νj,m(β,⋅) ∈ C(Ω) for all j ∈ J, m ∈ N0, β ∈ Nd +0, ∣β∣ ≤ min(m,k). +f) CVk +0,P (∂)(Ω,E) ≅ CVk +0,P (∂)(Ω)εE if E is locally complete and CVk +0,P (∂)(Ω) +a Fréchet–Schwartz space. +Proof. The generator (T E +m,T K +m)m∈N0 for (CVk +0,E) and (CVk +0,P (∂),E) is given +by domT E +m ∶= Ck(Ω,E) and +T E +m∶Ck(Ω,E) → Eωm, f �→ [(β,x) ↦ (∂β)Ef(x)], +for all m ∈ N0 and the same with K instead of E. +Set X ∶= Ω, K ∶= {K ⊂ Ω ∣ K compact} and π∶⋃m∈N0 ωm → X, π(β,x) ∶= x. We +have +∣f∣Ω∖K,j,m,α = +sup +x∈ωm +π(x)∉K +pα(T E +m(f)(x))νj,m(x), +for f ∈ CVk +0(Ω,E), K ∈ K, j ∈ J and m ∈ N0, implying that (16) is satisfied. With +APπ,K(Ω,E) from Proposition 4.1.9 we note that +CVk +0(Ω,E) = CVk(Ω,E) ∩ APπ,K(Ω,E). +As in Proposition 3.1.10 it follows that the generator (T E +m,T K +m)m∈N0 fulfils (17) and +(18) where we use Proposition 3.1.11, the barrelledness of CVk +0(Ω) resp. CVk +0,P (∂)(Ω) + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +53 +and the assumption that Vk is locally bounded away from zero on Ω. Therefore the +generator is strong and consistent by virtue of Proposition 4.1.9. +a)+d) Let f ∈ CVk +0(Ω,E), K ∈ K, j ∈ J and m ∈ N0. We claim that the set +Nj,m(f) = {∂β)Ef(x)νj,m(β,x) ∣ x ∈ Ω, β ∈ Mm} +is precompact in E by Proposition A.1.6. Since f vanishes at infinity in the weighted +topology, condition (i) of Proposition A.1.6 is fulfilled. Hence we only need to show +that condition (ii) is satisfied as well, i.e. we have to show that +Nπ⊂K,j,m(f) = +⋃ +β∈Mm +(∂β)Efνj,m(β,⋅)(K) +is precompact in E. Thus we only have to prove that the sets (∂β)Efνj,m(β,⋅)(K) +are precompact since Nπ⊂K,j,m(f) is a finite union of these sets. +But this is a +consequence of the proof of [15, §1, 16. Lemma, p. 15] using the continuity of +(∂β)Ef and the boundedness of νj,m(β,K), which follows from Vk being locally +bounded. So we deduce statements a) and d) from Corollary 3.2.5 (iv), Proposition +A.1.6 and the quasi-completeness of E. +b)+e) The set Kβ ∶= acx((∂β)Efνj,m(β,⋅)(Ω)) is absolutely convex and com- +pact by Proposition A.1.3 (ii) for every f ∈ CVk +0(Ω,E), j ∈ J, m ∈ N0 and β ∈ Mm +as E has metric ccp and νj,m(β,⋅) ∈ C(Ω). We have +Nj,m(f) = {(∂β)Ef(x)νj,m(β,x) ∣ x ∈ Ω, β ∈ Mm} ⊂ acx( ⋃ +β∈Mm +Kβ) +and the set on the right-hand side is absolutely convex and compact by [89, 6.7.3 +Proposition, p. 113]. Now, statements b)+e) follow from Corollary 3.2.5 (iv). +c)+f) They follow from Corollary 3.2.5 (ii). +□ +The spaces CVk +0(Ω) are Fréchet spaces and thus barrelled if J is countable by +[107, Proposition 3.7, p. 240]. In [107, Theorem 5.2, p. 255] the question is answered +when they have the approximation property. The spaces CV∞ +0 (Ω) and CV∞ +P (∂),0(Ω) +are closed subspaces of CV∞(Ω) and CV∞ +P (∂)(Ω), respectively. For conditions that +they are Fréchet–Schwartz spaces see the remarks below Example 3.2.7. +We already saw different choices for K in Example 4.2.9 b) and Example 4.2.22. +For holomorphic functions on an open subset Ω of an infinite dimensional Banach +space X the family K of Ω-bounded sets, i.e. bounded sets K ⊂ Ω with positive +distance to X ∖ Ω, is used in [71, p. 2] and [93, p. 2]. This family is clearly closed +under taking finite unions, so Proposition 4.1.9 is applicable as well. +Now, we consider an example of weighted smooth functions where the corre- +sponding space of scalar-valued functions may not be barrelled. For an open set +Ω ⊂ Rd, an lcHs E and a linear partial differential operator P(∂)E which is hypoel- +liptic if E = K we define the space of bounded zero solutions +C∞ +P (∂),b(Ω,E) ∶= {f ∈ C∞ +P (∂)(Ω,E) ∣ ∀ α ∈ A ∶ ∥f∥∞,α ∶= sup +x∈Ω +pα(f(x)) < ∞} +where C∞ +P (∂)(Ω,E) is the kernel of P(∂)E in C∞(Ω,E). Further, we set C∞ +P (∂),b(Ω) ∶= +C∞ +P (∂),b(Ω,K). +Apart from the topology given by (∥ ⋅ ∥∞,α)α∈A there is another +weighted locally convex topology on C∞ +P (∂),b(Ω,E) which is of interest, namely, the +one induced by the seminorms +∣f∣ν,α ∶= sup +x∈Ω +pα(f(x))∣ν(x)∣, +f ∈ C∞ +P (∂),b(Ω,E), +for ν ∈ C0(Ω) and α ∈ A. We denote by (C∞ +P (∂),b(Ω,E),β) the space C∞ +P (∂),b(Ω,E) +equipped with the topology β induced by the seminorms (∣ ⋅ ∣ν,α)ν∈C0(Ω),α∈A. The +topology β is called the strict topology. It is a bit tricky to prove the ε-compatibility + +54 +4. CONSISTENCY +of (C∞ +P (∂),b(Ω),β) and (C∞ +P (∂),b(Ω,E),β) because (C∞ +P (∂),b(Ω),β) may not be bar- +relled. +4.2.23. Remark. Let Ω ⊂ Rd be open and P(∂)K a hypoelliptic linear partial +differential operator. Then (C∞ +P (∂),b(Ω),β) is non-barrelled if τc does not coincide +with the ∥⋅∥∞-topology by [46, Section I.1, 1.15 Proposition, p. 12], e.g. (C∞ +∂,b(D),β) +is non-barrelled. +Hence we cannot use Proposition 3.1.11 c) directly. +4.2.24. Proposition. Let Ω ⊂ Rd be open, P(∂)K a hypoelliptic linear partial +differential operator and E an lcHs. Then (C∞ +P (∂),b(Ω),β)εE ≅ (C∞ +P (∂),b(Ω,E),β) if +E has metric ccp. +Proof. We set AP(Ω,E) ∶= C∞ +P (∂),b(Ω,E) and observe that (idEΩ,idΩK) is the +generator of ((C∞ +P (∂),b(Ω),β),E). First, we prove that the generator is consistent. +Clearly, we only need to show that S(u) ∈ AP(Ω,E) for every u ∈ (C∞ +P (∂),b(Ω),β)εE. +Let u ∈ (C∞ +P (∂),b(Ω),β)εE. Next, we show that u ∈ CW∞ +P (∂)(Ω)εE with CW∞ +P (∂)(Ω) +from Example 3.1.9 b). For α ∈ A there are an absolutely convex, compact K ⊂ +(C∞ +P (∂),b(Ω),β) and C > 0 such that for all f ′ ∈ (C∞ +P (∂),b(Ω),β)′ it holds that +pα(u(f ′)) ≤ C sup +f∈K +∣f ′(f)∣. +(24) +We denote by τc the topology of compact convergence on C∞ +P (∂),b(Ω), i.e. the topol- +ogy of uniform convergence on compact subsets of Ω. From the compactness of K in +(C∞ +P (∂),b(Ω),β) it follows that K is ∥ ⋅ ∥∞-bounded and τc-compact by [45, Proposi- +tion 1 (viii), p. 586] since (C∞ +P (∂),b(Ω),β) carries the induced topology of (Cb(Ω),β) +and the strict topology β is the mixed topology γ(τc,∥ ⋅ ∥∞) by [45, Proposition 3, +p. 590]. Let f ′ ∈ (C∞ +P (∂)(Ω),τc)′. Then there are M ⊂ Ω compact and C0 > 0 such +that +∣f ′(f)∣ ≤ C0 sup +x∈M +∣f(x)∣ +for all f ∈ C∞ +P (∂)(Ω). Choosing a compactly supported cut-off function ν ∈ C∞ +c (Ω) +with ν = 1 near M, we obtain +∣f ′(f)∣ ≤ C0 sup +x∈Ω +∣f(x)∣∣ν(x)∣ = C0∣f∣ν +for all f ∈ C∞ +P (∂)(Ω). +Therefore f ′ ∈ (C∞ +P (∂)(Ω),β)′. +In combination with the +τc-compactness of K it follows from (24) that u ∈ (C∞ +P (∂)(Ω),τc)εE. Using that +(C∞ +P (∂)(Ω),τc) = CW∞ +P (∂)(Ω) as locally convex spaces by the hypoellipticity of +P(∂)K (see e.g. [70, p. 690]), we obtain that u ∈ CW∞ +P (∂)(Ω)εE. Due to Propo- +sition 3.1.11 c) this yields that S(u) ∈ C∞ +P (∂)(Ω,E). Furthermore, we note that +∥S(u)∥∞,α = sup +x∈Ω +pα(S(u)(x)) = sup +x∈Ω +pα(u(δx)) ≤ +(24) C sup +x∈Ω +sup +f∈K +∣δx(f)∣ += C sup +f∈K +∥f∥∞ < ∞ +as K is ∥ ⋅ ∥∞-bounded, implying that S(u) ∈ C∞ +P (∂),b(Ω,E) = AP(Ω,E). Hence the +generator (idEΩ,idΩK) is consistent. +It is easily seen that e′ ○ f ∈ C∞ +P (∂),b(Ω) = AP(Ω) for all e′ ∈ E′ and f ∈ +C∞ +P (∂),b(Ω,E) (see the proof of Proposition 3.1.10), which proves that the gener- +ator is strong as well. +Moreover, we define Nν(f) ∶= {f(x)∣ν(x)∣ ∣ x ∈ Ω} for +f ∈ (C∞ +P (∂),b(Ω,E),β) and ν ∈ C0(Ω). The set K ∶= acx(Nν(f)) is absolutely con- +vex and compact in E by Proposition A.1.3 (ii) because f∣ν∣ ∈ C0(Ω,E) and Ω +second-countable, yielding our statement by Corollary 3.2.5 (iv). +□ + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +55 +If Ω ⊂ C is an open, simply connected set, P(∂) = ∂ and E is complete, then +the preceding result is also a consequence of [17, 3.10 Satz, p. 146]. +Next, we consider the vector-valued Beurling–Björck space Sµ(Rd,E) which +generalises the Schwartz space and whose scalar-valued counterpart was studied by +Björck in [20], by Schmeisser and Triebel in [155] (see [20, Definition 1.8.1, p. 375], +[155, 1.2.1.2 Definition, p. 15]) whereas semigroups on its toplogical dual space +were treated by Alvarez et al. in [5]. Since Fourier transformation is involved in the +definition of Sµ(Rd,E), we start with the following statement. +4.2.25. Proposition. Let E be a locally complete lcHs over C, f ∈ S(Rd,E) +and x ∈ Rd. Then fe−i⟨x,⋅⟩ is Pettis-integrable on Rd where ⟨⋅,⋅⟩ is the usual scalar +product on Rd. +Proof. We choose m ∶= d + 1 and set ψ∶Rd → [0,∞), ψ(ζ) ∶= (1 + ∣ζ∣2)−m/2, as +well as g∶Rd → [0,∞), g(ζ) ∶= ψ(ζ)−1. Then ψ ∈ L1(Rd,λ) and ψg = 1. Moreover, +let x = (xi) ∈ Rd and set u∶Rd → E, u(ζ) ∶= f(ζ)e−i⟨x,ζ⟩g(ζ). We note that +(∂en)Eu(ζ) += (∂en)Ef(ζ)e−i⟨x,ζ⟩g(ζ) − ixnf(ζ)e−i⟨x,ζ⟩g(ζ) + mf(ζ)e−i⟨x,ζ⟩(1 + ∣ζ∣2)(m/2)−1ζn +for all ζ = (ζi) ∈ Rd and 1 ≤ n ≤ d, which implies +pα((∂en)Eu(ζ)) ≤ pα((∂en)Ef(ζ))g(ζ) + ∣xn∣pα(f(ζ))g(ζ) + mpα(f(ζ))g(ζ) +for all α ∈ A and hence +sup +ζ∈Rd +β∈Nd +0,∣β∣≤1 +pα((∂β)Eu(ζ)) ≤ (1 + ∣xn∣ + m)∣f∣S(Rd),m,α. +Therefore u = fe−i⟨x,⋅⟩g is (weakly) C1 +b , which yields u ∈ C[1] +b (Rd,E) by Proposition +A.1.5. Now, we choose h∶Rd → (0,∞), h(ζ) ∶= 1 + ∣ζ∣2. Then +sup +ζ∈Rd pα(u(ζ)h(ζ)) ≤ sup +ζ∈Rd pα(f(ζ))(1 + ∣ζ∣2)(m+2)/2 ≤ ∣f∣S(Rd),m+2,α < ∞ +for all α ∈ A, and for every ε > 0 there is r > 0 such that 1 ≤ εh(ζ) for all ζ ∉ Br(0) =∶ +K. +We deduce from Proposition A.2.7 (iii) that fe−i⟨x,⋅⟩ is Pettis-integrable on +Rd. +□ +Thus, for f ∈ S(Rd,E) with locally complete E the Fourier transformation +FE(f)∶Rd → E, FE(f)(x) ∶= (2π)−d/2 ∫ +Rd +f(ζ)e−i⟨x,ζ⟩dζ, +is defined. From the Pettis-integrability we get (e′ ○ FE)(f) = FC(e′ ○ f) for every +e′ ∈ E′. As FC(e′ ○ f) ∈ S(Rd) for every e′ ∈ E′ by [20, Proposition 1.8.2, p. 375], +we obtain from the weak-strong principle Corollary 5.2.21 (or [30, Theorem 9, p. +232] and [131, Mackey’s theorem 23.15, p. 268]) that FE(f) ∈ S(Rd,E). +For a locally complete lcHs E over C and a continuous function µ∶Rd → [0,∞) +such that +(γ) there are a ∈ R, b > 0 with µ(x) ≥ a + bln(1 + ∣x∣) for all x ∈ Rd, +we set +Sµ(Rd,E) ∶= {f ∈ C∞(Rd,E) ∣ ∀ m,j ∈ N0, α ∈ A ∶ ∣f∣j,m,α < ∞} +where ∣f∣m,j,α ∶= max(qm,j,α(f),qm,j,α(FE(f))) with +qm,j,α(f) ∶= +sup +x∈Rd +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x))ejµ(x). + +56 +4. CONSISTENCY +We note that from qm,j,α(f) < ∞ for all m,j ∈ N0, α ∈ A and condition (γ) it follows +that f ∈ S(Rd,E) and hence qm,j,α(FE(f)) is defined. Further, we set Sµ(Rd) ∶= +Sµ(Rd,C). We observe that Sµ(Rd,E) is a dom-space. Indeed, let ωm ∶= ̃ωm ∪ ̃ωm,1 +where ̃ωm ∶= Mm × Rd with Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ m} and ̃ωm,1 ∶= ̃ωm × {1} for all +m ∈ N0. Setting domT E +m ∶= S(Rd,E) and T E +m∶S(Rd,E) → Eωm by +T E +m(f)(β,x) ∶= (∂β)Ef(x) and T E +m(f)(β,x,1) ∶= ((∂β)E ○ FE)f(x), +(β,x) ∈ ̃ωm, +for every m ∈ N0 as well as AP(Rd,E) ∶= ERd, we have that Sµ(Rd,E) is a dom- +space with weights given by νj,m(β,x) ∶= νj,m(β,x,1) ∶= ejµ(x) for all (β,x) ∈ ̃ωm +and m,j ∈ N0. +The condition (γ) is introduced in [20, p. 363]. Choosing µ(x) ∶= ln(1 + ∣x∣), +x ∈ Rd, we get the Schwartz space Sµ(Rd,E) = S(Rd,E) back. +4.2.26. Example. Let E be a locally complete lcHs over C and µ∶Rd → [0,∞) +continuous such that condition (γ) is fulfilled. +(i) If E has metric ccp, or +(ii) if µ ∈ C1(Rd) and there are k ∈ N0, C > 0 such that ∣∂enµ(x)∣ ≤ Cekµ(x) +for all x ∈ Rd and 1 ≤ n ≤ d, +then Sµ(Rd,E) ≅ Sµ(Rd)εE. +Proof. First, we show that the generator (T E +m,T C +m)m∈N0 for (Sµ,E) is strong +and consistent. From +(∂β)C(e′ ○ f)(x) = e′((∂β)Ef(x)), +(β,x) ∈ ̃ωm, +where ̃ωm = {β ∈ Nd +0 ∣ ∣β∣ ≤ m} × Rd, we get in combination with the Pettis- +integrability by Proposition 4.2.25 that +((∂β)C ○ FC)(e′ ○ f)(x) = e′(((∂β)E ○ FE)f(x)), +(β,x) ∈ ̃ωm +(25) +for all e′ ∈ E′, f ∈ Sµ(Rd,E) and m ∈ N0, which means that the generator is +strong. For consistency we consider the case µ(x) = ln(1 + ∣x∣), x ∈ Rd, i.e. the +Schwartz space, first. Due to Corollary 3.2.10 the map S∶S(Rd)εE → S(Rd,E) is +an isomorphism and according to Theorem 3.2.4 its inverse is given by +Rt∶S(Rd,E) → S(Rd)εE, f ↦ J −1 ○ Rt +f. +Let u ∈ S(Rd)εE. Thanks to the proof of Corollary 3.2.10 we only need to show +that +u(δx ○ (∂β ○ FC)) = (∂β)EFE(S(u))(x), +x ∈ Rd. +We set f ∶= S(u) ∈ S(Rd,E) and from (25) we obtain +Rt +f(δx ○ ((∂β)C ○ FC))(e′) = (∂β)CFC(e′ ○ f)(x) = e′((∂β)EFE(f)(x)), +e′ ∈ E′, +for all x ∈ Rd and β ∈ Nd +0, which results in +u(δx ○ ((∂β)C ○ FC)) = S−1(f)(δx ○ ((∂β)C ○ FC)) = J −1(Rt +f(δx ○ ((∂β)C ○ FC))) += (∂β)EFE(f)(x) = (∂β)EFE(S(u))(x). +(26) +Thus (T E +m,T C +m)m∈N0 is a consistent generator for (S,E). +Let us turn to general µ. Let u ∈ Sµ(Rd)εE. We show that u ∈ S(Rd)εE. Then +it follows from the first part of the proof that (T E +m,T C +m)m∈N0 is a consistent generator +for (Sµ,E). For α ∈ A there are an absolutely convex compact set K ⊂ Sµ(Rd) and +C > 0 such that for all f ′ ∈ Sµ(Rd)′ it holds +pα(u(f ′)) ≤ C sup +f∈K +∣f ′(f)∣. +(27) + +4.2. FURTHER EXAMPLES OF ε-PRODUCTS +57 +The compactness of K in Sµ(Rd) and the estimate +sup +x∈Rd +β∈Nd +0,∣β∣≤m +∣(∂β)Cf(x)∣(1 + ∣x∣2)j/2 ≤ +sup +x∈Rd +β∈Nd +0,∣β∣≤m +∣(∂β)Cf(x)∣e(j/2)(2/b)(µ(x)−a) +≤ e−(aj)/b∣f∣j,m, +f ∈ Sµ(Rd), +for all j,m ∈ N0 by condition (γ) imply that the inclusion Sµ(Rd) ↪ S(Rd) is +continuous and thus that K is compact in S(Rd). Let f ′ ∈ S(Rd)′. Then there are +j,m ∈ N0 and C0 > 0 such that +∣f ′(f)∣ ≤ C0 +sup +x∈Rd +β∈Nd +0,∣β∣≤m +∣(∂β)Cf(x)∣(1 + ∣x∣2)j/2 ≤ C0e−(aj)/b∣f∣j,m +for all f ∈ Sµ(Rd). Hence f ′ ∈ Sµ(Rd)′ and from (27) we obtain that u ∈ S(Rd)εE +because K is absolutely convex and compact in S(Rd). +Condition (γ) implies that µ(x) → ∞ for ∣x∣ → ∞. Noting that for every j ∈ N +and ε > 0 there is r > 0 such that +ejµ(x) +e2jµ(x) = e−jµ(x) < ε +(28) +for all x ∉ Br(0), we deduce ∣f∣Rd∖Br(0),m,j,α ≤ ε∣f∣m,2j,α for every f ∈ Sµ(Rd,E), +m ∈ N0 and α ∈ A. +(i) Thus, if E has metric ccp, then the sets Kβ ∶= acx((∂β)Efejµ(Rd)) and +Kβ,1 ∶= acx((∂β)EFE(f)ejµ(Rd)) are absolutely convex and compact by Propo- +sition A.1.3 (ii) for every f ∈ Sµ(Rd,E), j,m ∈ N0 and β ∈ Mm as (∂β)Efejµ ∈ +C0(Rd,E) and (∂β)EFE(f)ejµ ∈ C0(Rd,E). +(ii) We set g0∶Rd → E, g0(x) ∶= (∂β)Ef(x)ejµ(x), and g1∶Rd → E, g1(x) ∶= +(∂β)EFE(f)(x)ejµ(x), for j,m ∈ N0 and β ∈ Mm. We observe that +(∂en)Eg0(x) = (∂β+en)Ef(x)ejµ(x) + j(∂β)Ef(x)ejµ(x)∂enµ(x) +and +(∂en)Eg1(x) = (∂β+en)EFE(f)(x)ejµ(x) + j(∂β)EFE(f)(x)ejµ(x)∂enµ(x) +for all x ∈ Rd and 1 ≤ n ≤ d. As in Example 3.2.9 it follows from condition (ii) that +there are k ∈ N0, C > 0 such that +sup +x∈Rd +γ∈Nd +0,∣γ∣≤1 +pα((∂γ)Egi(x)) ≤ ∣f∣m+1,j,α + Cj∣f∣m,j+k,α +for all α ∈ A and i = 0,1. Thus g0 and g1 are (weakly) C1 +b . We set h ∶= ejµ and note +that +sup +x∈Rd pα(gi(x)h(x)) ≤ ∣f∣m,2j,α < ∞ +for all α ∈ A and i = 0,1. This yields that Kβ = acx(g0(Rd)) and Kβ,1 = acx(g1(Rd)) +are absolutely convex and compact by Proposition A.1.4 with (28) and Proposition +A.1.5. +Then we have in both cases +Nj,m(f) = ({(∂β)Ef(x)ejµ(x) ∣ x ∈ Rd, β ∈ Mm} +∪ {(∂β)EFE(f)(x)ejµ(x) ∣ x ∈ Rd, β ∈ Mm}) +⊂acx( ⋃ +β∈Mm +(Kβ ∪ Kβ,1)) + +58 +4. CONSISTENCY +and the set on the right-hand side is absolutely convex and compact by [89, 6.7.3 +Proposition, p. 113], which implies that Sµ(Rd,E) ≅ Sµ(Rd)εE by Corollary 3.2.5 +(iv). +□ +We come back to these spaces in Theorem 5.1.5. +Another example that is +related to Fourier transformation is the space of vector-valued smooth functions +that are 2π-periodic in each variable. We equip the space C∞(Rd,E) for an lcHs E +with the system of seminorms generated by +∣f∣K,m,α ∶= +sup +x∈Rd +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x))χK(x) = +sup +x∈K +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x)), +f ∈ C∞(Rd,E), +for K ⊂ Rd compact, m ∈ N0 and α ∈ A, i.e. we consider CW∞(Rd,E). +By +C∞ +2π(Rd,E) we denote the topological subspace of C∞(Rd,E) consisting of the func- +tions which are 2π-periodic in each variable. Further, we set C∞ +2π(Rd) ∶= C∞ +2π(Rd,K). +4.2.27. Example. If E is a locally complete lcHs, then C∞ +2π(Rd,E) ≅ C∞ +2π(Rd)εE. +Proof. First, we note that for each x ∈ Rd and 1 ≤ n ≤ d we have δx = δx+2πen +in C∞ +2π(Rd)′ and thus +SC∞ +2π(Rd)(u)(x) − SC∞ +2π(Rd)(u)(x + 2πen) = u(δx − δx+2πen) = 0, +u ∈ C∞ +2π(Rd)εE, +implying that SC∞ +2π(Rd)(u) is 2π-periodic in each variable. In addition, we observe +that e′ ○ f is 2π-periodic in each variable for all e′ ∈ E′ and f ∈ C∞ +2π(Rd,E). Now, +we obtain as in Example 3.2.8 a) for k = ∞ that SC∞ +2π(Rd)∶C∞ +2π(Rd)εE → C∞ +2π(Rd,E) +is an isomorphism. +□ +We return to C∞ +2π(Rd,E) in Theorem 5.1.4 and Theorem 5.6.14. Now, we direct +our attention to spaces of continuously partially differentiable functions on an open +bounded set such that all derivatives can be continuously extended to the boundary. +Let E be an lcHs, k ∈ N∞ and Ω ⊂ Rd open and bounded. The space Ck(Ω,E) is +given by +Ck(Ω,E) ∶= {f ∈ Ck(Ω,E) ∣ (∂β)Ef cont. extendable on Ω for all β ∈ Nd +0, ∣β∣ ≤ k} +and equipped with the system of seminorms given by +∣f∣α ∶= +sup +x∈Ω +β∈Nd +0,∣β∣≤k +pα((∂β)Ef(x)), +f ∈ Ck(Ω,E), +for α ∈ A if k < ∞, and by +∣f∣m,α ∶= +sup +x∈Ω +β∈Nd +0,∣β∣≤m +pα((∂β)Ef(x)), +f ∈ C∞(Ω,E), +for m ∈ N0 and α ∈ A if k = ∞. Further, we set Ck(Ω) ∶= Ck(Ω,K). +4.2.28. Example. Let E be an lcHs, k ∈ N∞ and Ω ⊂ Rd open and bounded. +(i) If E has metric ccp, or +(ii) if E is locally complete, k = ∞ and there exists C > 0 such that for each +x,y ∈ Ω there is a continuous path from x to y in Ω whose length is +bounded by C∣x − y∣, +then Ck(Ω,E) ≅ Ck(Ω)εE. +Proof. The generator coincides with the one of Example 4.2.22. +Due to +Proposition 3.1.11 we have S(u) ∈ Ck(Ω,E) and +(∂β)ES(u)(x) = u(δx ○ (∂β)K), +β ∈ Nd +0, ∣β∣ ≤ k, x ∈ Ω, + +4.3. RIESZ–MARKOV–KAKUTANI REPRESENTATION THEOREMS +59 +for all u ∈ Ck(Ω)εE since Ck(Ω) is a Banach space if k < ∞, and a Fréchet space +if k = ∞, in particular, both are barrelled. As a consequence of Proposition 4.1.7 +and Lemma 4.1.8 with T = (∂β)K for β ∈ Nd +0, ∣β∣ ≤ k, we obtain that (∂β)ES(u) ∈ +Cext(Ω,E) for all u ∈ Ck(Ω)εE. Thus the generator is consistent. It is easy to check +that it is strong, too. This yields (ii) by Corollary 3.2.5 (ii) since C∞(Ω) is a nuclear +Fréchet space by [131, Examples 28.9 (5), p. 350] under the conditions on Ω. +Let us turn to part (i). Let f ∈ Ck(Ω,E), J ∶= {1}, m ∈ N0 and set Mm ∶= {β ∈ +Nd +0 ∣ ∣β∣ ≤ k} if k < ∞, and Mm ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ m} if k = ∞. We denote by fβ the +continuous extension of (∂β)Ef on the compact metrisable set Ω. The set +N1,m(f) = {(∂β)Ef(x) ∣ x ∈ Ω, β ∈ Mm} ⊂ +⋃ +β∈Mm +fβ(Ω) +is relatively compact and metrisable since it is a subset of a finite union of the +compact metrisable sets fβ(Ω) as in Example 3.2.8. Due to Corollary 3.2.5 (iv) we +obtain our statement (i) as E has metric ccp. +□ +We close this section by an examination of the topological subspace +E0(E) ∶= {f ∈ C∞([0,1],E) ∣ ∀ k ∈ N0 ∶ (∂k)Ef(1) = 0} +where (∂k)Ef(1) ∶= limx→1+(∂k)Ef(x). Further, we set E0 ∶= E0(K). +4.2.29. Example. Let E be a locally complete lcHs. Then E0εE ≅ E0(E). +Proof. We note that Ω ∶= (0,1) satisfies the condition on Ω in Example 4.2.28 +(ii) with C ∶= 1 and thus C∞([0,1]) and its closed subspace E0 are nuclear Fréchet +spaces. The generator coincides with the one of Example 4.2.28. From the proof of +Example 4.2.28 we know that +lim +x→1+(∂k)ES(u)(x) = u(δ1 ○ (∂k)K) = u(0) = 0, +k ∈ N0, +for all u ∈ E0εE. In combination with Example 4.2.28 this yields the consistency +of the generator. Again, its strength is easy to check. Therefore our statement is +valid by Corollary 3.2.5 (ii). +□ +4.3. Riesz–Markov–Kakutani representation theorems +In this subsection we generalise the concept of strength and consistency such +that it is not strictly bounded to dom-spaces and their generators anymore. This +allows us to answer the question: Given T K +m ∈ F(Ω)′ is there T E +m ∈ L(F(Ω,E),E) +such that (T E +m,T K +m) is strong and consistent? Furthermore, we will see that the +operators T E +m are usually the ones that can be obtained from integral representations +of T K +m, i.e. we transfer Riesz–Markov–Kakutani theorems from the scalar-valued to +the vector-valued case. +We recall that the Riesz–Markov–Kakutani theorem for +compact topological Hausdorff spaces Ω says that for every T R ∈ Cb(Ω)′ there is a +unique regular R-valued Borel measure µ on Ω such that +T R(f) = ∫ +Ω +f(x)dµ(x), +f ∈ Cb(Ω), +(29) +which was proved by Riesz [146, p. 976] in the case Ω ∶= [0,1] and by Kakutani [95, +Theorem 9, p. 1009] for general compact Hausdorff Ω (see Saks [152, Eq. (1.1), 6., p. +408, 411] for compact metric Ω). Markov treated the case where Ω is a normal (not +necessarily Hausdorff) topological space and the T R are positive linear functionals +on Cb(Ω) such that T R(1) = 1 [127, Definition 2, p. 167]. In this case, for every +such T R there is a unique exterior density µ on Ω in the sense of [127, Definition +3, p. 167] such that (29) holds by [127, Theorem 22, p. 184] and the right-hand +side is read in the sense of [127, Eq. (71), (72), (80), p. 180–181] (see also [60, + +60 +4. CONSISTENCY +IV.6.2 Theorem, p. 262] for a more familiar version with regular (finitely) additive +bounded Borel measures µ). +4.3.1. Definition (strong, consistent). Let E be an lcHs and Ω a non-empty +set. +Let F(Ω) ⊂ KΩ and F(Ω,E) ⊂ EΩ be lcHs such that δx ∈ F(Ω)′ for all +x ∈ Ω. +Let (ωm)m∈M be a family of non-empty sets, T K +m∶domT K +m → Kωm and +T E +m∶domT E +m → Eωm be linear with F(Ω) ⊂ domT K +m ⊂ KΩ and F(Ω,E) ⊂ domT E +m ⊂ +EΩ for all m ∈ M. +a) We call (T E +m,T K +m)m∈M a consistent family for (F(Ω),E), in short (F,E), +if we have for every u ∈ F(Ω)εE, m ∈ M and x ∈ ωm that +(i) S(u) ∈ F(Ω,E) and T K +m,x ∶= δx ○ T K +m ∈ F(Ω)′, +(ii) T E +mS(u)(x) = u(T K +m,x). +b) We call (T E +m,T K +m)m∈M a strong family for (F(Ω),E), in short (F,E), if +we have for every e′ ∈ E′, f ∈ F(Ω,E), m ∈ M and x ∈ ωm that +(i) e′ ○ f ∈ F(Ω), +(ii) T K +m(e′ ○ f)(x) = (e′ ○ T E +m(f))(x). +Note that ωm need not be a subset of Ω. As a convention we omit the index m +of the set ωm, the operators T E +m and T K +m if M is a singleton. The following remark +shows that the preceding definition of a consistent resp. strong family coincides with +the usual definition in the case of generators of dom-spaces (see Definition 3.1.7). +4.3.2. Remark. Let (T E +m,T K +m)m∈M be a generator for (FV,E). We note that +the condition T K +m,x ∈ FV(Ω)′ for all m ∈ M and x ∈ ωm in a)(i) of Definition +4.3.1 is always satisfied for generators by Remark 3.1.6 b). Moreover, if S(u) ∈ +AP(Ω,E) ∩ domT E +m for u ∈ FV(Ω)εE and all m ∈ M and a)(ii) of Definition 4.3.1 +is fulfilled, then S(u) ∈ FV(Ω,E) by Lemma 3.1.8, implying that a)(i) is satisfied. +Further, if f ∈ FV(Ω,E) and e′ ○f ∈ AP(Ω)∩domT K +m for all e′ ∈ E′ and m ∈ M and +b)(ii) of Definition 4.3.1 is fulfilled, then e′ ○ f ∈ FV(Ω) by Lemma 3.2.1, implying +that b)(i) is satisfied. +The next proposition is the key result in transferring Riesz–Markov–Kakutani +theorems from the scalar-valued to the vector-valued case. To state this proposition +we need that our map S∶F(Ω)εE → F(Ω,E) is an isomorphism and that its inverse +is given as in Theorem 3.2.4, i.e. that +Rt∶F(Ω,E) → F(Ω)εE, f ↦ J −1 ○ Rt +f, +is the inverse of S where Rt +f(f ′)(e′) = f ′(e′ ○ f), for f ′ ∈ F(Ω)′ and e′ ∈ E′, and +J ∶E → E′⋆ is the canonical injection in the algebraic dual E′⋆ of E′. +4.3.3. Proposition. Let E be an lcHs, (Ω,Σ,µ) a measure space and F(Ω) +and F(Ω,E) ε-compatible with inverse Rt of S and (T E +0 ,T K +0 ) a strong family for +(F,E) with ω0 ∶= Ω. +If T K +0 (f) is integrable for every f ∈ F(Ω) and T E +0 (f) is +Pettis-integrable on Ω for every f ∈ F(Ω,E) and +T K∶F(Ω) → K, T K(f) ∶= ∫ +Ω +T K +0 (f)(x)dµ(x), +is continuous, then +u(T K) = ∫ +Ω +T E +0 S(u)(x)dµ(x), +u ∈ F(Ω)εE. +Proof. Let u ∈ F(Ω)εE and set f ∶= S(u) ∈ F(Ω,E). We have +Rt +f(T K)(e′) = T K(e′○f) = ∫ +Ω +T K +0 (e′○f)(x)dµ(x) = ⟨e′,∫ +Ω +T E +0 f(x)dµ(x)⟩, +e′ ∈ E′, + +4.3. RIESZ–MARKOV–KAKUTANI REPRESENTATION THEOREMS +61 +by the strength of (T E +0 ,T K +0 ) and the Pettis-integrability of T E +0 (f), which yields +u(T K) = S−1(f)(T K) = J −1(Rt +f(T K)) = ∫ +Ω +T E +0 f(x)dµ(x) = ∫ +Ω +T E +0 S(u)(x)dµ(x) +due to Rt being the inverse of S. +□ +4.3.4. Proposition. Let E be an lcHs, (Ω,Σ,µ) a measure space, (T E +0 ,T K +0 ) a +strong family for (F,E) with ω0 ∶= Ω such that T E +0 (f) is Pettis-integrable on Ω for +every f ∈ F(Ω,E), and (T E,T K) a consistent family for (F,E) such that +T E(f) = ∫ +Ω +T E +0 f(x)dµ(x), +f ∈ F(Ω,E). +Then (T E,T K) is a strong family for (F,E), T K +0 (f) is integrable for every f ∈ F(Ω) +and +T K = ∫ +Ω +T K +0 (f)(x)dµ(x), +f ∈ F(Ω). +Proof. We set f ⋅ e∶Ω → E, (f ⋅ e)(x) ∶= f(x)e, for e ∈ E and f ∈ F(Ω). Since +(T E,T K) is a consistent family for (F,E), we get f ⋅e = S(Θ(e⊗f)) ∈ F(Ω,E) and +T E(f ⋅ e) = Θ(e ⊗ f)(T K) = T K(f)e +(30) +with the map Θ from (3). From the strength of (T E +0 ,T K +0 ) we deduce that +e′ ○ T E +0 (f ⋅ e) = T K +0 (e′ ○ (f ⋅ e)) = T K +0 (e′(e)f) = e′(e)T K +0 (f) +and from the Pettis-integrability of T E +0 (f ⋅ e) that +T K(f)e′(e) = +(30) ⟨e′,T E(f ⋅ e)⟩ = ∫ +Ω +e′(e)T K +0 (f)(x)dµ(x) +for all e′ ∈ E′. This implies that e′(e)T K +0 (f) is integrable for all e′ ∈ E′. Further, +since E is non-trivial by our assumptions in Chapter 2, there is some e0 ∈ E, e0 ≠ 0. +By the Hahn–Banach theorem there is some e′ +0 ∈ E′ with e′ +0(e0) ≠ 0, which yields +that T K +0 (f) is integrable and +T K(f) = ∫ +Ω +T K +0 (f)(x)dµ(x). +Furthermore, we conclude in combination with the strength of (T E +0 ,T K +0 ) and the +Pettis-integrability of T E +0 (f) for all f ∈ F(Ω,E) that +⟨e′,T E(f)⟩ = ∫ +Ω +T K +0 (e′ ○ f)(x)dµ(x) = T K(e′ ○ f) +for all f ∈ F(Ω,E) and e′ ∈ E′, which means that (T E,T K) is a strong family for +(F,E). +□ +Let us apply the preceding propositions to the space D([0,1],E) of E-valued +càdlàg functions on [0,1]. For f ∈ D([0,1],E) we set f(x−) ∶= limw→x− f(w) if +x ∈ (0,1], and f(0−) ∶= 0. +4.3.5. Proposition. Let E be a quasi-complete lcHs. Then for every T K ∈ +D([0,1])′ there is T E ∈ L(D([0,1],E),E) such that (T E,T K) is a consistent family +for (D,E) and there are a unique regular K-valued Borel measure µ on [0,1] and +a unique ϕ ∈ ℓ1([0,1],K) such that +T E(f) = ∫ +[0,1] +f(x)dµ(x) + +∑ +x∈[0,1] +(f(x) − f(x−))ϕ(x), +f ∈ D([0,1],E). +(31) + +62 +4. CONSISTENCY +On the other hand, if (T E,T K) is a consistent family, there is a unique regular K- +valued Borel measure µ on [0,1] such that (31) holds and T E ∈ L(D([0,1],E),E). +Proof. Due to the representation theorem [139, Theorem 1, p. 383] there are +a unique regular K-valued Borel measure µ on [0,1] and a unique ϕ ∈ ℓ1([0,1],K) +such that +T K(f) = ∫ +[0,1] +f(x)dµ(x) + +∑ +x∈[0,1] +(f(x) − f(x−))ϕ(x), +f ∈ D([0,1],K). +(32) +By Example 4.2.5 S∶D([0,1])εE → D([0,1],E) is an isomorphism with inverse +Rt∶f ↦ J ○ Rt +f. +The next part is the analogon of Proposition 4.3.3 for ∑x∈[0,1]. We set +T K +1 ∶D([0,1]) → K, T K +1 (f) ∶= +∑ +x∈[0,1] +(f(x) − f(x−))ϕ(x), +and note that T K +1 ∈ D([0,1])′. Let u ∈ D([0,1])εE and set f ∶= S(u) ∈ D([0,1],E). +We have +Rt +f(T K +1 )(e′) = T K +1 (e′ ○ f) = +∑ +x∈[0,1] +((e′ ○ f)(x) − (e′ ○ f)(x−))ϕ(x) += ⟨e′, +∑ +x∈[0,1] +(f(x) − f(x−))ϕ(x)⟩, +e′ ∈ E′, +due to the Pettis-summability of x ↦ (f(x) − f(x−))ϕ(x) on [0,1] by Proposition +A.2.6, which yields +u(T K +1 ) = S−1(f)(T K +1 ) = J −1(Rt +f(T K +1 )) = +∑ +x∈[0,1] +(f(x) − f(x−))ϕ(x) += +∑ +x∈[0,1] +(S(u)(x) − S(u)(x−))ϕ(x). +(33) +We note that every f ∈ D([0,1],E) is Pettis-integrable and that x ↦ (f(x) �� +f(x−))ϕ(x) is Pettis-summable on [0,1] by Proposition A.2.6. Further, +pα( ∫ +[0,1] +f(x)dµ(x)+ ∑ +x∈[0,1] +(f(x)−f(x−))ϕ(x)) ≤ (∣µ∣([0,1])+2∥ϕ∥ℓ1) sup +x∈[0,1] +pα(f(x)) +for all f ∈ D([0,1],E) and α ∈ A. The rest follows from Proposition 4.3.3 with +(T E +0 ,T K +0 ) ∶= (idE[0,1],idK[0,1]) combined with (33). For the uniqueness of µ in (31) +use that the µ in (32) is unique and Proposition 4.3.4 (and for the uniqueness of ϕ +use an analogon of Proposition 4.3.4 for (T E +1 ,T K +1 )). +□ +Let us turn to continuous functions that vanish at infinity. +4.3.6. Proposition. Let Ω be a locally compact [second countable] topological +Hausdorff space and E an lcHs with [metric] ccp. Then for every T K ∈ C0(Ω)′ there +is T E ∈ L(C0(Ω,E),E) such that (T E,T K) is a consistent family for (C0,E) and +there is a unique regular K-valued Borel measure µ on Ω such that +T E(f) = ∫ +Ω +f(x)dµ(x), +f ∈ C0(Ω,E). +(34) +On the other hand, if (T E,T K) is a consistent family, then there is a unique regular +K-valued Borel measure µ on Ω such that (34) holds and T E ∈ L(C0(Ω,E),E). + +4.3. RIESZ–MARKOV–KAKUTANI REPRESENTATION THEOREMS +63 +Proof. Due to the Riesz–Markov–Kakutani representation theorem (see [149, +6.19 Theorem, p. 130]) there is a unique regular K-valued Borel measure µ on Ω +such that +T K(f) = ∫ +Ω +f(x)dµ(x), +f ∈ C0(Ω). +(35) +By Example 4.2.11 S∶C0(Ω)εE → C0(Ω,E) is an isomorphism with inverse Rt∶f ↦ +J ○ Rt +f. We note that every f ∈ C0(Ω,E) is Pettis-integrable by Proposition A.2.7 +(i) resp. (ii) with ψ ∶= g ∶= 1 since +∫ +Ω +∣ψ(x)∣d∣µ∣(x) = ∣µ∣(Ω) < ∞ +and +pα(∫ +Ω +f(x)dµ(x)) ≤ ∣µ∣(Ω)sup +x∈Ω +pα(f(x)), +f ∈ C0(Ω,E), +for all α ∈ A. The rest follows from Proposition 4.3.3 with (T E +0 ,T K +0 ) ∶= (idEΩ,idKΩ). +For the uniqueness of µ in (34) use that the µ in (35) is unique and Proposition +4.3.4. +□ +Next, we consider the space of bounded continuous E-valued functions on a +locally compact topological Hausdorff space Ω, i.e. +Cb(Ω,E) = {f ∈ C(Ω,E) ∣ ∀ α ∈ A ∶ sup +x∈Ω +pα(f(x)) < ∞}, +but equipped with the strict topology β (see Remark 4.2.23) which is induced by +the seminorms +∣f∣ν,α ∶= sup +x∈Ω +pα(f(x))∣ν(x)∣, +f ∈ Cb(Ω,E), +for ν ∈ C0(Ω) and α ∈ A. +4.3.7. Proposition. Let Ω be a locally compact [second countable] topological +Hausdorff space and E an lcHs with [metric] ccp. Then for every T K ∈ (Cb(Ω),β)′ +there is T E ∈ L((Cb(Ω,E),β),E) such that (T E,T K) is a consistent family for +((Cb(Ω),β),E) and there is a unique regular K-valued Borel measure µ on Ω such +that +T E(f) = ∫ +Ω +f(x)dµ(x), +f ∈ Cb(Ω,E). +(36) +On the other hand, if (T E,T K) is a consistent family, then there is a unique regular +K-valued Borel measure µ on Ω such that (36) holds and T E ∈ L((Cb(Ω,E),β),E). +Proof. Due to the Riesz–Markov–Kakutani representation theorem [89, 7.6.3 +Theorem, p. 141] for the strict topology there is a unique regular K-valued Borel +measure µ on Ω such that +T K(f) = ∫ +Ω +f(x)dµ(x), +f ∈ Cb(Ω). +Since T K is continuous, there are ν ∈ C0(Ω) and C > 0 such that +∣∫ +Ω +⟨e′,f(x)⟩dµ(x)∣ = ∣T K(e′ ○ f)∣ ≤ C sup +x∈Ω +∣(e′ ○ f)(x)ν(x)∣ ≤ C sup +x∈K +∣e′(x)∣, +e′ ∈ E′, +for f ∈ Cb(Ω,E) with K ∶= acx(fν(Ω)). As K is absolutely convex and compact by +Proposition A.1.3, f is Pettis-integrable on Ω w.r.t. µ by the Mackey–Arens theo- +rem. The remaining parts of the proof follow from Example 4.2.11 and Proposition +4.3.3 as in Proposition 4.3.6. +□ + +64 +4. CONSISTENCY +4.3.8. Proposition. Let Ω be a locally compact [second countable] topological +Hausdorff space and E an lcHs with [metric] ccp. Then for every T K ∈ CW(Ω)′ there +is T E ∈ L(CW(Ω,E),E) such that (T E,T K) is a consistent family for (CW,E) and +there is a unique regular K-valued Borel measure µ on Ω with compact support such +that +T E(f) = ∫ +Ω +f(x)dµ(x), +f ∈ C(Ω,E). +(37) +On the other hand, if (T E,T K) is a consistent family, then there is a unique regular +K-valued Borel measure µ on Ω with compact support such that (37) holds and +T E ∈ L(CW(Ω,E),E). +Proof. By the Riesz–Markov–Kakutani representation theorem given in the +remark after [35, Chap. 4, §4.8, Proposition 14, p. INT IV.48] for the topology of +compact convergence there is a unique regular K-valued Borel measure µ on Ω with +compact support such that +T K(f) = ∫ +Ω +f(x)dµ(x), +f ∈ C(Ω). +Since T K is continuous, there are a compact set M ⊂ Ω and C > 0 such that +∣∫ +Ω +⟨e′,f(x)⟩dµ(x)∣ = ∣T K(e′ ○ f)∣ ≤ C sup +x∈M +∣(e′ ○ f)(x)∣ ≤ C sup +x∈K +∣e′(x)∣, +e′ ∈ E′, +for f ∈ C(Ω,E) with the absolutely convex and compact set K ∶= acx(f(M)), +implying that f is Pettis-integrable on Ω w.r.t. µ by the Mackey–Arens theorem. +The rest of the proof is identical to the one of Proposition 4.3.7. +□ +4.3.9. Proposition. Let Ω ⊂ Rd be open and E a locally complete lcHs. Then +for every T K ∈ CW∞(Ω)′ there is T E ∈ L(CW∞(Ω,E),E) such that (T E,T K) is +a consistent family for (CW∞,E). Given any open neighbourhood U ⊂ Ω of the +compact distributional support suppT K of T K there are m ∈ N0 and a family of +K-valued Radon measures (µβ)β∈Nd +0,∣β∣≤m on Ω such that suppµβ ⊂ U for all β ∈ Nd +0, +∣β∣ ≤ m, and +T E(f) = ∑ +∣β∣≤m +∫ +Ω +(∂β)Ef(x)dµβ(x), +f ∈ C∞(Ω,E). +(38) +On the other hand, if (T E,T K) is a consistent family, then given any open neigh- +bourhood U ⊂ Ω of the compact distributional support suppT K of T K there are +m ∈ N0 and a family of K-valued Radon measures (µβ)β∈Nd +0,∣β∣≤m on Ω such that +(38) holds, suppµβ ⊂ U for all β ∈ Nd +0, ∣β∣ ≤ m and T E ∈ L(CW∞(Ω,E),E). +Proof. T K ∈ CW∞(Ω)′ is a distribution with compact support and thus has +finite order by [171, Corollary, p. 259]. Denote by m ∈ N0 the order of T K. Given +any open neighbourhood U ⊂ Ω of suppT K there is a family of K-valued Radon +measures (µβ)β∈Nd +0,∣β∣≤m on Ω such that +T K(f) = ∑ +∣β∣≤m +∫ +Ω +(∂β)Kf(x)dµβ(x), +f ∈ C∞(Ω), +and suppµβ ⊂ U for all β ∈ Nd +0, ∣β∣ ≤ m by [171, Theorem 24.4, p. 259]. Since the +support Kβ ∶= suppµβ of µβ is compact +∫ +Ω +(∂β)Kf(x)dµβ(x) = ∫ +Kβ +(∂β)Kf(x)dµβ(x), +f ∈ C∞(Ω), + +4.3. RIESZ–MARKOV–KAKUTANI REPRESENTATION THEOREMS +65 +and (∂β)Ef ∈ C1(Ω,E) for f ∈ C∞(Ω,E), it follows from Lemma A.2.2 and Remark +A.2.4 that (∂β)Ef is Pettis-integrable on Ω w.r.t. µβ for all β and that +pα(∫ +Ω +(∂β)Ef(x)dµβ(x)) ≤ ∣µβ∣(Kβ) sup +x∈Kβ +pα((∂β)Ef(x)), +α ∈ A. +By Example 3.2.8 a) the map S∶CW∞(Ω)εE → CW∞(Ω,E) is an isomorphism with +inverse Rt∶f ↦ J ○ Rt +f. The remaining parts of the proof follow from Proposition +4.3.3 with (T E +0 ,T K +0 ) ∶= ((∂β)E,(∂β)K). +□ +4.3.10. Proposition. Let E be a locally complete lcHs. Then for every T K ∈ +S(Rd)′ there is T E ∈ L(S(Rd,E),E) such that (T E,T K) is a consistent family for +(S,E) and there are m ∈ N0 and a family of continuous functions (gβ)β∈Nd +0,∣β∣≤m on +Rd growing at infinity slower than some polynomial such that +T E(f) = ∑ +∣β∣≤m +∫ +Rd +gβ(x)(∂β)Ef(x)dx, +f ∈ S(Rd,E). +(39) +On the other hand, if (T E,T K) is a consistent family, then there are m ∈ N0 and a +family of continuous functions (gβ)β∈Nd +0,∣β∣≤m on Rd growing at infinity slower than +some polynomial such that (39) holds and T E ∈ L(S(Rd,E),E). +Proof. Let T K ∈ S(Rd)′. Then there are m ∈ N0 and a family of continuous +functions (gβ)β∈Nd +0,∣β∣≤m on Rd growing at infinity slower than some polynomial such +that +T K(f) = ∑ +∣β∣≤m +∫ +Rd +gβ(x)(∂β)Kf(x)dx, +f ∈ S(Rd), +by [171, Theorem 25.4, p. 272]. +Here, gβ growing at infinity slower than some +polynomial means that there are k ∈ N0 and C > 0 such that ∣gβ(x)∣ ≤ C(1+∣x∣2)k/2 +for all x ∈ Rd. Since the family (gβ) is finite, we can take one k and one C for all +β. Due to the proof of Example 3.2.9 and Corollary 3.2.10 we know that Kβ ∶= +acx(((∂β)Ef)(1 + ∣ ⋅ ∣2)k/2(Rd)) is absolutely convex and compact for f ∈ S(Rd,E). +The estimate +∣∫ +Rd +⟨e′,gβ(x)(∂β)Ef(x)⟩dx∣ ≤ C sup +x∈Rd ∣e′((∂β)Ef(x))∣(1 + ∣x∣2)k/2 = C sup +x∈Kβ +∣e′(x)∣ +for all e′ ∈ E′ and f ∈ S(Rd,E) yields that gβ(∂β)Ef is Pettis-integrable on Rd +w.r.t. the Lebesgue measure by the Mackey–Arens theorem. Further, it implies +that +pα(∫ +Rd +gβ(x)(∂β)Ef(x)dx) ≤ C sup +x∈Rd pα((∂β)Ef(x))(1 + ∣x∣2)k/2, +α ∈ A, +as in Lemma A.2.2. By Corollary 3.2.10 the map S∶S(Rd)εE → S(Rd,E) is an +isomorphism with inverse Rt∶f ↦ J ○ Rt +f. The remaining parts of the proof follow +from Proposition 4.3.3 with (T E +0 ,T K +0 ) ∶= (gβ(∂β)E,gβ(∂β)K). +□ +4.3.11. Remark. +a) Let Ω ⊂ Rd be open and E a locally complete lcHs. +Then Proposition 4.3.9 is still valid with CW∞ replaced by CW∞ +P (∂) due +to the Hahn–Banach theorem and Example 3.2.8 b). If P(∂)K is a hy- +poelliptic linear partial differential operator, then one can represent T E +as in (37) due to Proposition 4.2.17 but the measure µ need not be unique +anymore. + +66 +4. CONSISTENCY +b) Let Ω ⊂ Rd be open and E an lcHs with metric ccp. Then Proposition +4.3.7 is still valid with Cb replaced by C∞ +P (∂),b for a hypoelliptic linear +partial differential operator P(∂)K due to the Hahn–Banach theorem and +Proposition 4.2.24 but the measure µ need not be unique anymore. +c) All families (T E,T K) considered in this section are strong which is a con- +sequence of Proposition 4.3.4 (and of Pettis-summability in Proposition +4.3.5). + +CHAPTER 5 +Applications +5.1. Lifting the properties of maps from the scalar-valued case +In this section we briefly show how to use the ε-compatibility of spaces F(Ω) +and F(Ω,E) to lift properties like injectivity, surjectivity, bijectivity and continu- +ity from a map T K to a map T E if (T E,T K) forms a consistent family. Especially, +we pay attention to surjectivity whose transfer to the vector-valued case is ac- +complished by Grothendieck’s classical theory of tensor products of Fréchet spaces +[83] and by the splitting theory of Vogt for Fréchet spaces [173] and of Bonet and +Domański for PLS-spaces [54]. In order to apply splitting theory, we recall the +definitions of the topological invariants (Ω), (DN) and (PA). +Let us recall that a Fréchet space F with an increasing fundamental system of +seminorms (∣∣∣⋅∣∣∣k)k∈N satisfies (Ω) if +∀ p ∈ N ∃ q ∈ N ∀ k ∈ N ∃ n ∈ N, C > 0 ∀ r > 0 ∶ Uq ⊂ CrnUk + 1 +r Up +where Uk ∶= {x ∈ F ∣ ∣∣∣x∣∣∣k ≤ 1} (see [131, Chap. 29, Definition, p. 367]). +We recall that a Fréchet space (F,(∣∣∣⋅∣∣∣k)k∈N) satisfies (DN) by [131, Chap. 29, +Definition, p. 359] if +∃ p ∈ N ∀ k ∈ N ∃ n ∈ N, C > 0 ∀ x ∈ F ∶ ∣∣∣x∣∣∣2 +k ≤ C∣∣∣x∣∣∣p∣∣∣x∣∣∣n. +A PLS-space is a projective limit X = lim +←� +N∈N +XN, where the XN given by inductive +limits XN = lim +�→ +n∈N +(XN,n,∣∣∣⋅∣∣∣N,n) are DFS-spaces (which are also called LS-spaces), +and it satisfies (PA) if +∀ N ∃ M ∀ K ∃ n ∀ m ∀ η > 0 ∃ k,C,r0 > 0 ∀ r > r0 ∀ x′ ∈ X′ +N ∶ +∣∣∣x′ ○ iM +N ∣∣∣ +∗ +M,m ≤ C(rη∣∣∣x′ ○ iK +N∣∣∣ +∗ +K,k + 1 +r ∣∣∣x′∣∣∣ +∗ +N,n) +where ∣∣∣⋅∣∣∣∗ denotes the dual norm of ∣∣∣⋅∣∣∣ and iM +N , iK +N the linking maps (see [26, +Section 4, Eq. (24), p. 577]). +Examples of Fréchet spaces with (DN) are the spaces of rapidly decreasing +sequences s(Nd), s(Nd +0) and s(Zd), the space C∞([a,b]) of all C∞-smooth functions +on (a,b) such that all derivatives can be continuously extended to the boundary +and the space of smooth functions C∞ +2π(Rd) that are 2π-periodic in each variable. +Examples of ultrabornological PLS-space with (PA) are Fréchet–Schwartz spaces, +the space of tempered distributions S(Rd)′ +b, the space of distributions D(Ω)′ +b and +ultradistributions of Beurling type D(ω)(Ω)′ +b on an open set Ω ⊂ Rd. These and +many more examples may be found in [26], [54, Corollary 4.8, p. 1116] and [112, +Example 3, p. 7]. +5.1.1. Proposition. +a) Let Y be a Fréchet space and X a semi-reflexive +lcHs. Then Lb(X′ +b,Y ′ +b ) ≅ Lb(Y,(X′ +b)′ +b) via taking adjoints. +b) Let E be an lcHs and X a Montel space. Then Lb(X′ +b,E) ≅ XεE where +the isomorphism is the identity map. +67 + +68 +5. APPLICATIONS +Proof. a) We consider the map +t(⋅)∶Lb(X′ +b,Y ′ +b ) → Lb(Y,(X′ +b)′ +b), u ↦ tu, +defined by tu(y)(x′) ∶= u(x′)(y) for y ∈ Y and x′ ∈ X′. First, we prove that t(⋅) is +well-defined. Let u ∈ L(X′ +b,Y ′ +b ) and y ∈ Y . Since u ∈ L(X′ +b,Y ′ +b ) and {y} is bounded +in Y , there are a bounded set B ⊂ X and C > 0 such that +∣tu(y)(x′)∣ = ∣u(x′)(y)∣ ≤ C sup +x∈B +∣x′(x)∣ +for all x′ ∈ X′, implying tu(y) ∈ (X′ +b)′. +Let us denote by (∥⋅∥Y,n)n∈N the (directed) system of seminorms generating the +metrisable locally convex topology of Y . The canonical embedding J∶Y → (Y ′ +b )′ +b is +an isomorphism between Y and J(Y ) by [131, Corollary 25.10, p. 298] because Y +is a Fréchet space. For a bounded set M ⊂ X′ +b we note that +sup +x′∈M +∣tu(y)(x′)∣ = sup +x′∈M +∣u(x′)(y)∣ = sup +x′∈M +∣⟨J(y),u(x′)⟩∣. +The next step is to prove that u(M) is bounded in Y ′ +b . Let N ⊂ Y be bounded. +Since u ∈ L(X′ +b,Y ′ +b ), there are again a bounded set B ⊂ X and a constant C > 0 +such that +sup +x′∈M +sup +y∈N +∣u(x′)(y)∣ ≤ C sup +x′∈M +sup +x∈B +∣x′(x)∣ < ∞ +where the last estimate follows from the boundedness of M ⊂ X′ +b. Hence u(M) is +bounded in Y ′ +b . By the remark about the canonical embedding there are n ∈ N and +C0 > 0 such that +sup +x′∈M +∣tu(y)(x′)∣ = +sup +y′∈u(M) +∣⟨J(y),y′⟩∣ ≤ C0∥y∥Y,n, +so tu ∈ L(Y,(X′ +b)′ +b) and the map t(⋅) is well-defined. +Let us turn to injectivity. Let u,v ∈ L(X′ +b,Y ′ +b ) with tu = tv. This is equivalent +to +u(x′)(y) = tu(y)(x′) = tv(y)(x′) = v(x′)(y) +for all y ∈ Y and x′ ∈ X′. This implies u(x′) = v(x′) for all x′ ∈ X′, hence u = v. +Next, we turn to surjectivity. We consider the map +t(⋅)∶Lb(Y,(X′ +b)′ +b) → Lb(X′ +b,Y ′ +b ), u ↦ tu, +defined by tu(x′)(y) ∶= u(y)(x′) for x′ ∈ X′ and y ∈ Y . We show that this map is +well-defined. Let u ∈ Lb(Y,(X′ +b)′ +b) and x′ ∈ X′. Since u ∈ Lb(Y,(X′ +b)′ +b) and {x′} is +bounded in X′, there are n ∈ N and C > 0 such that +∣tu(x′)(y)∣ = ∣u(y)(x′)∣ ≤ C∥y∥Y,n +for all y ∈ Y , yielding tu(x′) ∈ Y ′. Let B ⊂ Y be bounded. The semi-reflexivity +of X implies that for every u(y), y ∈ B, there is a unique xu(y) ∈ X such that +u(y)(x′) = x′(xu(y)) for all x′ ∈ X′. Then we get +sup +y∈B +∣tu(x′)(y)∣ = sup +y∈B +∣u(y)(x′)∣ = sup +y∈B +∣x′(xu(y))∣. +We claim that D ∶= {xu(y) ∣ y ∈ B} is a bounded set in X. Let N ⊂ X′ be finite. +Then the set M ∶= {tu(x′) ∣ x′ ∈ N} ⊂ Y ′ is finite. We have +sup +y∈B +sup +x′∈N +∣x′(xu(y))∣ = sup +y∈B +sup +x′∈N +∣tu(x′)(y)∣ = sup +y∈B +sup +y′∈M +∣y′(y)∣ < ∞ +where the last estimate follows from the fact that the bounded set B is weakly +bounded. Thus D is weakly bounded and by [131, Mackey’s theorem 23.15, p. 268] +bounded in X. Therefore it follows from +sup +y∈B +∣tu(x′)(y)∣ = sup +y∈B +∣x′(xu(y))∣ = sup +x∈D +∣x′(x)∣ + +5.1. LIFTING THE PROPERTIES OF MAPS FROM THE SCALAR-VALUED CASE +69 +for all x′ ∈ X′ that tu ∈ L(X′ +b,Y ′ +b ), which means that t(⋅) is well-defined. +Let +u ∈ L(Y,(X′ +b)′ +b). Then we have tu ∈ Lb(X′ +b,Y ′ +b ). In addition, for all y ∈ Y and all +x′ ∈ X′ +t(tu)(y)(x′) = tu(x′)(y) = u(y)(x′) +is valid and so t(tu)(y) = u(y) for all y ∈ Y , proving the surjectivity. +The last step is to prove the continuity of t(⋅) and its inverse. Let M ⊂ Y and +B ⊂ X′ +b be bounded sets. Then +sup +y∈M +sup +x′∈B +∣tu(y)(x′)∣ = sup +y∈M +sup +x′∈B +∣u(x′)(y)∣ = sup +x′∈B +sup +y∈M +∣u(x′)(y)∣ += sup +x′∈B +sup +y∈M +∣t(tu)(x′)(y)∣ +holds for all u ∈ L(X′ +b,Y ′ +b ). Therefore, t(⋅) and its inverse are continuous. +b) Let T ∈ L(X′ +b,E). For α ∈ A there are a bounded set B ⊂ X and C > 0 such +that +pα(T(x′)) ≤ C sup +x∈B +∣x′(x)∣ ≤ C +sup +x∈acx(B) +∣x′(x)∣ +for every x′ ∈ X′ where acx(B) is the closure of the absolutely convex hull of B. +The set acx(B) is absolutely convex and compact by [89, 6.2.1 Proposition, p. 103] +and [89, 6.7.1 Proposition, p. 112] since B is bounded in the Montel space X. Hence +we gain T ∈ L(X′ +κ,E). +Let M ⊂ X′ be equicontinuous. Due to [89, 8.5.1 Theorem (a), p. 156] M is +bounded in X′ +b. Therefore, +id∶Lb(X′ +b,E) → Le(X′ +κ,E) = XεE +is continuous. +Let T ∈ L(X′ +κ,E). For α ∈ A there are an absolutely convex compact set B ⊂ X +and C > 0 such that +pα(T(x′)) ≤ C sup +x∈B +∣x′(x)∣ +for every x′ ∈ X′. Since the compact set B is bounded, we get T ∈ L(X′ +b,E). +Let M be a bounded set in X′ +b. Then M is equicontinuous by virtue of [171, +Theorem 33.2, p. 349], as X, being a Montel space, is barrelled. Thus +id∶Le(X′ +κ,E) → Lb(X′ +b,E) +is continuous. +□ +For part e) of the next theorem we need that our map SF2(Ω2)∶F2(Ω2)εE → +F2(Ω2,E) is an isomorphism and that its inverse is given as in Theorem 3.2.4, i.e. +that +Rt∶F2(Ω2,E) → F2(Ω2)εE, f ↦ J −1 ○ Rt +f, +is the inverse of SF2(Ω2) where Rt +f(f ′)(e′) = f ′(e′ ○ f) for f ′ ∈ F2(Ω2)′ and e′ ∈ E′, +and J ∶E → E′⋆ is the canonical injection in the algebraic dual E′⋆ of E′. +5.1.2. Theorem. Let E be an lcHs, F1(Ω1) and F1(Ω1,E) as well as F2(Ω2) +and F2(Ω2,E) be ε-into-compatible. Let (T E,T K) be a consistent family for (F1,E) +such that T K∶F1(Ω1) → F2(Ω2) is continuous and T E∶F1(Ω1,E) → F2(Ω2,E). +Then the following holds: +a) T E ○ SF1(Ω1) = SF2(Ω2) ○ (T KεidE). +b) If SF1(Ω1) is surjective and T K is injective, then T E is injective, continu- +ous and +T E = SF2(Ω2) ○ (T KεidE) ○ S−1 +F1(Ω1). + +70 +5. APPLICATIONS +If in addition SF2(Ω2) is surjective and T K an isomorphism, then T E is +an isomorphism with inverse +(T E)−1 = SF1(Ω1) ○ ((T K)−1εidE) ○ S−1 +F2(Ω2). +c) If SF2(Ω2) and T KεidE are surjective, then T E is surjective. +d) If SF2(Ω2) and T K are surjective, F1(Ω1), F2(Ω2) and E are Fréchet +spaces and +(i) F1(Ω1) and F2(Ω2) are nuclear, or +(ii) E is nuclear, +then T E is surjective. +e) If SF2(Ω2) is surjective with inverse Rt, T K is surjective, F1(Ω1) and +F2(Ω2) are Fréchet spaces, kerT K is nuclear and has (Ω), and +(i) F1(Ω1) and F2(Ω2) are Montel spaces, E = F ′ +b where F is a Fréchet +space satisfying (DN), or +(ii) F1(Ω1) and F2(Ω2) are Schwartz spaces, E is an ultrabornological +PLS-space satisfying (PA), +then T E is surjective. +Proof. a) Let u ∈ F1(Ω1)εE. Then +(T E ○ SF1(Ω1))(u)(x) = u(δx ○ T K) = (u ○ (T K)t)(δx) = (T KεidE)(u)(δx) += SF2(Ω2)((T KεidE)(u))(x), +x ∈ Ω2, +as (T E,T K) is consistent for (F1,E), which proves part a). +b) If SF1(Ω1) is surjective, then SF1(Ω1) is an isomorphism, because it is an +isomorphism into, and we have +T E = SF2(Ω2) ○ (T KεidE) ��� S−1 +F1(Ω1) +by part a). If T K is injective, then T KεidE is also injective by [159, Chap. I, §1, +Proposition 1, p. 20] and thus T E by the formula above as well since SF1(Ω1) is +an isomorphism and SF2(Ω2) an isomorphism into. If SF2(Ω2) is surjective and T K +an isomorphism, then SF2(Ω2) and T KεidE are isomorphisms, the latter by [159, +Chap. I, §1, Proposition 1, p. 20] and its inverse is (T K)−1εidE. The rest of part +b) follows from the formula for T E above. +c) Let f ∈ F2(Ω2,E). +Then there is g ∈ F1(Ω1)εE such that (SF2(Ω2) ○ +(T KεidE))(g) = f. Hence we obtain h ∶= SF1(Ω1)(g) ∈ F1(Ω1,E) and T E(h) = f by +part a). +d) For n = 1,2 the continuous linear injection (see (3)) +Θn∶Fn(Ωn) ⊗π E → Fn(Ωn)εE, +k +∑ +j=1 +fj ⊗ ej �→ [y ↦ +k +∑ +j=1 +y(fj)ej], +from the tensor product Fn(Ωn) ⊗π E with the projective topology extends to a con- +tinuous linear map ̂Θn∶Fn(Ωn)̂⊗πE → Fn(Ωn)εE on the completion Fn(Ωn)̂⊗πE +of Fn(Ωn) ⊗π E. +The map ̂Θn is also a topological isomorphism since Fn(Ωn) +is nuclear for n = 1,2 in case (i) resp. E is nuclear in case (ii). +Furthermore, +T K ⊗π idE∶F1(Ω1)⊗π E → F2(Ω2)⊗π E is defined by the relation Θ2 ○(T K ⊗π idE) = +(T KεidE)○Θ1. We denote by T K ̂⊗π idE the continuous linear extension of T K⊗πidE +to the completion F1(Ω1)̂⊗πE. Moreover, Fn(Ωn) for n = 1,2 and E are Fréchet +spaces, T K and idE are linear, continuous and surjective, so T K ̂⊗π idE is surjective +by [94, 10.24 Satz, p. 255]. We observe that +T KεidE = ̂Θ2 ○ (T K ̂⊗π idE) ○ ̂Θ −1 +1 + +5.1. LIFTING THE PROPERTIES OF MAPS FROM THE SCALAR-VALUED CASE +71 +and deduce that T KεidE is surjective. Now, we apply part c), which proves part +d). +e) Throughout this proof we use the notation X′′ ∶= (X′ +b)′ +b for a locally convex +Hausdorff space X and T ∶= T K. The space F1(Ω1) is a Fréchet space and so its +closed subspace kerT as well. Further, Fn(Ωn) is a Montel space for n = 1,2 and +kerT nuclear, thus they are reflexive. The sequence +0 → kerT +i→ F1(Ω1) +T→ F2(Ω2) → 0, +(40) +where i means the inclusion, is a topologically exact sequence of Fréchet spaces +because T is surjective by assumption. Let us denote by J0∶kerT → (kerT)′′ and +Jn∶Fn(Ωn) → Fn(Ωn)′′ for n = 1,2 the canonical embeddings which are topological +isomorphisms since kerT and Fn(Ωn) are reflexive for n = 1,2. Then the exactness +of (40) implies that +0 → (kerT)′′ i0→ F1(Ω1)′′ T1→ F2(Ω2)′′ → 0, +(41) +where i0 ∶= J0 ○ i ○ J−1 +0 +and T1 ∶= J2 ○ T ○ J−1 +1 , is an exact topological sequence. This +exact sequence is topological because the (strong) bidual of a Fréchet space is again +a Fréchet space by [131, Corollary 25.10, p. 298]. +(i) Let E = F ′ +b where F is a Fréchet space with (DN). Then Ext1(F,(kerT)′′) = +0 by [174, 5.1 Theorem, p. 186] since kerT is nuclear and satisfies (Ω) and therefore +(kerT)′′ as well. Combined with the exactness of (41) this implies that the sequence +0 → L(F,(kerT)′′) +i∗ +0→ L(F,F1(Ω1)′′) +T ∗ +1→ L(F,F2(Ω2)′′) → 0 +is exact by [137, Proposition 2.1, p. 13–14] where i∗ +0(B) ∶= i0○B and T ∗ +1 (D) ∶= T1○D +for B ∈ L(F,(kerT)′′) and D ∈ L(F,F1(Ω1)′′). In particular, we obtain that +T ∗ +1 ∶L(F,F1(Ω1)′′) → L(F,F2(Ω2)′′) +(42) +is surjective. Via E = F ′ +b and Proposition 5.1.1 (X = Fn(Ωn) and Y = F) we have +the isomorphisms into +ψn ∶= SFn(Ωn) ○ t(⋅)∶L(F,Fn(Ωn)′′) → Fn(Ωn,E), +ψn(u) = (SFn(Ωn) ○ t(⋅))(u) = [x ↦ tu(δx)], +for n = 1,2 and the inverse +ψ−1 +2 (f) = (S ○ t(⋅))−1(f) = (t(⋅) ○ S−1 +F2(Ω2))(f) = t(J −1 ○ Rt +f), +f ∈ F2(Ω2,E). +Let g ∈ F2(Ω2,E). Then ψ−1 +2 (g) ∈ L(F,F2(Ω2)′′) and by the surjectivity of (42) +there is u ∈ L(F,F1(Ω1)′′) such that T ∗ +1 u = ψ−1 +2 (g). So we get ψ1(u) ∈ F1(Ω1,E). +Next, we show that T Eψ1(u) = g is valid. Let y ∈ F and x ∈ Ω2. Then +T E(ψ1(u))(x) = tu(δx ○ T) +by consistency and +T E(ψ1(u))(x)(y) = tu(δx ○ T)(y) = u(y)(δx ○ T) = ⟨δx ○ T,J−1 +1 (u(y))⟩ += ⟨δx,TJ−1 +1 (u(y))⟩ = ⟨[J2 ○ T ○ J−1 +1 ](u(y)),δx⟩ = ⟨(T1 ○ u)(y),δx⟩ += ⟨(T ∗ +1 u)(y),δx⟩ = ψ−1 +2 (g)(y)(δx) = t(J −1 ○ Rt +g)(y)(δx) += (J −1 ○ Rt +g)(δx)(y) = J −1(J (g(x))(y) = g(x)(y). +Thus T E(ψ1(u))(x) = g(x) for every x ∈ Ω2, which proves the surjectivity. +(ii) Let E be an ultrabornological PLS-space satisfying (PA). Since the nuclear +Fréchet space kerT is also a Schwartz space, its strong dual (kerT)′ +b is a DFS- +space. +By [26, Theorem 4.1, p. 577] we obtain Ext1 +P LS((kerT)′ +b,E) = 0 as the +bidual (kerT)′′ satisfies (Ω), E is a PLS-space satisfying (PA) and condition (c) +in the theorem is fulfilled because (kerT)′ +b is the strong dual of a nuclear Fréchet + +72 +5. APPLICATIONS +space. Moreover, we have Proj1 E = 0 due to [180, Corollary 3.3.10, p. 46] because +E is an ultrabornological PLS-space. +Then the exactness of the sequence (41), +[26, Theorem 3.4, p. 567] and [26, Lemma 3.3, p. 567] (in the lemma the same +condition (c) as in [26, Theorem 4.1, p. 577] is fulfilled and we choose H = (kerT)′′, +F = F1(Ω1)′′ and G = F2(Ω2)′′), imply that the sequence +0 → L(E′ +b,(kerT)′′) +i∗ +0→ L(E′ +b,F1(Ω1)′′) +T ∗ +1→ L(E′ +b,F2(Ω2)′′) → 0 +is exact. The maps i∗ +0 and T ∗ +1 are defined as in part (i). Especially, we get that +T ∗ +1 ∶L(E′ +b,F1(Ω1)′′) → L(E′ +b,F2(Ω2)′′) +(43) +is surjective. +By [54, Remark 4.4, p. 1114] we have Lb(Fn(Ωn)′ +b,E′′) ≅ Lb(E′ +b,Fn(Ωn)′′) for +n = 1,2 via taking adjoints since Fn(Ωn), being a Fréchet–Schwartz space, is a PLS- +space and hence its strong dual an LFS-space, which is regular by [180, Corollary +6.7, 10. ⇔ 11., p. 114], and E is an ultrabornological PLS-space, in particular, +reflexive by [53, Theorem 3.2, p. 58]. In addition, the map +P∶Lb(Fn(Ωn)′ +b,E′′) → Lb(Fn(Ωn)′ +b,E), +defined by P(u)(y) ∶= J −1(u(y)) for u ∈ L(Fn(Ωn)′ +b,E′′) and y ∈ Fn(Ωn)′, is an +isomorphism because E is reflexive. Due to Proposition 5.1.1 b) with X = Fn(Ωn) +we obtain the isomorphisms into +ψn ∶= S ○ J −1 ○ t(⋅)∶Lb(E′ +b,Fn(Ωn)′′) → Fn(Ωn,E), +ψn(u) = [SFn(Ωn) ○ J −1 ○ t(⋅)](u) = [x ↦ J −1(tu(δx))], +for n = 1,2 and the inverse given by +ψ−1 +2 (f) = (SF2(Ω2) ○ J −1 ○ t(⋅))−1(f) = [t(⋅) ○ J ○ S−1 +F2(Ω2)](f) = t(J ○ J −1 ○ Rt +f) += t(Rt +f) +for f ∈ F2(Ω2,E). +Let g ∈ F2(Ω2,E). +Then ψ−1 +2 (g) ∈ L(E′ +b,F2(Ω2)′′) and by the surjectivity +of (43) there exists u ∈ L(E′ +b,F1(Ω1)′′) such that T ∗ +1 u = ψ−1 +2 (g). +So we have +ψ1(u) ∈ F1(Ω1,E). The last step is to show that T Eψ1(u) = g. As in part (i) we +gain for every x ∈ Ω2 +T E(ψ1(u))(x) = J −1(tu(δx ○ T)) +by consistency and for every y ∈ E′ +tu(δx ○ T)(y) = u(y)(δx ○ T) = (T ∗ +1 u)(y)(δx) = ψ−1 +2 (g)(y)(δx) = t(Rt +g)(y)(δx) += δx(y ○ g) = y(g(x)) = J (g(x))(y). +Thus we have tu(δx ○ T) = J (g(x)) and therefore T E(ψ1(u))(x) = g(x) for all +x ∈ Ω2. +□ +Theorem 5.1.2 d) and e) are generalisations of [116, Corollary 4.3, p. 2689] and +[112, Theorem 5, p. 7–8] where T C is the Cauchy–Riemann operator ∂ on certain +weighted spaces CV∞(Ω) of smooth functions. Our next result is the well-known +application of tensor product theory and splitting theory to linear partial differential +operators we already mentioned in the introduction. +5.1.3. Corollary. Let E be a locally complete lcHs, Ω1 ⊂ Rd open and P(∂)K +be a linear partial differential operator with C∞-smooth coefficients. Then the fol- +lowing holds: +a) P(∂)E = SC∞(Ω1) ○ (P(∂)KεidE) ○ S−1 +C∞(Ω1). +b) If K = C, P(D) ∶= P(D)C ∶= P(−i∂)C has constant coefficients and is + +5.1. LIFTING THE PROPERTIES OF MAPS FROM THE SCALAR-VALUED CASE +73 +(i) elliptic, or +(ii) hypoelliptic and Ω1 convex, +and +(iii) E is a Fréchet space, or +(iv) E = F ′ +b where F is a Fréchet space satisfying (DN), d ≥ 2, or +(v) E is an ultrabornological PLS-space satisfying (PA), d ≥ 2, +then P(D)E∶C∞(Ω1,E) → C∞(Ω1,E) is surjective. +Proof. Part a) follows from Theorem 5.1.2 a), Example 3.2.8 a) and the con- +sistency of (P(∂)E,P(∂)K) because (7) holds for u ∈ CW∞(Ω1)εE as well. +Let us turn to part b). The inverse of SC∞(Ω1) is given by Rt by Example 3.2.8 +a). The map P(D) = P(D)C∶C∞(Ω1) → C∞(Ω1) is surjective by [86, Corollary +10.6.8, p. 43] and [86, Theorem 10.6.2, p. 41] in case (ii) resp. by [86, Corollary +10.8.2, p. 51] in case (i). The space CW∞(Ω1), i.e. C∞(Ω1) with its usual topology +(see Example 3.1.9 b)), is a nuclear Fréchet space and thus its closed subspace +kerP(D)C as well. In case (i) kerP(D)C has (Ω) due to [173, Proposition 2.5 (b), +p. 173] and in case (ii) due to [140, 4.5 Corollary (a), p. 202]. Hence the surjectivity +of P(D)E follows from Theorem 5.1.2 d)+e). +□ +Recently, it was shown in [47, Theorem 4.2, p. 13] that in the case that +Ω1 is convex, kerP(D)C has property (Ω) for any P(D) with constant coeffi- +cients (so without the assumption of P(D) being hypoelliptic). +Hence we may +replace the assumption of P(D) being hypoelliptic in (ii) by the assumption that +P(D)C∶C∞(Ω1) → C∞(Ω1) is surjective. Even more recently, a necessary and suf- +ficient condition for the surjectivity of P(D)C and kerP(D)C having property (Ω) +for P(D) with constant coefficients and general open Ω1 ⊂ Rd was derived in [48, +Theorem 1.1. (a), p. 3] using shifted fundamental solutions. +Even though Corollary 5.1.3 b) is known, it is often proved without using tensor +products or splitting theory (see e.g. [94, Theorem 10.10, p. 240]) or it is phrased +as the surjectivity of P(D)Ĉ⊗π idE (see e.g. [171, Eq. (52.4), p. 541]) and the proof +of the relation +P(D)E = SC∞(Ω1) ○ (̂Θ1 ○ (P(D)C ̂⊗π idE) ○ ̂Θ −1 +1 ) ○ S−1 +C∞(Ω1) +for Fréchet spaces E is omitted (see e.g. [171, p. 545–546]), or only the surjectivity +of T ∗ +1 = P(D)∗ +1 in part e) of Theorem 5.1.2 is actually shown and it is only stated +but not proved that this implies the surjectivity of P(D)E (see e.g. the statement +of surjectivity of P(D)E in [173, p. 168] for elliptic P(D) and E = F ′ +b for a Fréchet +space F with (DN) and that it is ‘only’ shown that P(D)∗ +1 is surjective by [173, +Proposition 2.5 (b), p. 173] and [173, Theorem 2.4 (b), p. 173] where the symbol +P(D)∗ is used instead of P(D)∗ +1 in [173, p. 172] since the isomorphism J1 = J2 is +omitted). So, apart from being the probably most classical application of tensor +products or splitting theory, that is the reason why we still included Corollary 5.1.3. +Let us give another application of Theorem 5.1.2 d) and e), namely, a vector- +valued Borel–Ritt theorem. +5.1.4. Theorem. Let E be an lcHs and (xn)n∈N0 a sequence in E. If +(i) E is a Fréchet space, or +(ii) E = F ′ +b where F is a Fréchet space satisfying (DN), or +(iii) E is an ultrabornological PLS-space satisfying (PA), +then there is f ∈ C∞ +2π(R,E) such that (∂n)Ef(0) = xn for all n ∈ N0. +Proof. By the Borel–Ritt theorem [94, Satz 9.12, p. 206] the map +T K∶C∞ +2π(R) → KN0, T K(f) ∶= ((∂n)Kf(0))n∈N0, + +74 +5. APPLICATIONS +is surjective and obviously linear and continuous as well. Now, we define the map +T E∶C∞ +2π(R,E) → EN0 by replacing K by E in the definition of T K. Due to Example +4.2.1 KN0 and EN0 are ε-compatible and the inverse of SKN0 is given by Rt. In +addition, C∞ +2π(R) and C∞ +2π(R,E) are ε-compatible by Example 4.2.27 as in all three +cases E is complete. We observe that (T E,T K) is consistent by Proposition 3.1.11 +c). +The spaces KN0 and C∞ +2π(R) are nuclear Fréchet spaces. +The first by [171, +Theorem 51.1, p. 526] and the second because it is a subspace of the nuclear space +C∞(R) by [131, Examples 28.9 (1), p. 349–350] and [131, Proposition 28.6, p. 347]. +Hence in case (i) our statement follows from Theorem 5.1.2 d). Moreover, kerT K +is nuclear since C∞ +2π(R) is nuclear. By the proof of [131, Lemma 31.3, p. 392–393] +kerT K is isomorphic to s(N0). The space s(N0) has (Ω) by [131, Lemma 29.11 +(3), p. 368] and thus kerT K as well because (Ω) is a linear topological invariant +by [131, Lemma 29.11 (1), p. 368]. Therefore our statement in case (ii) and (iii) +follows from Theorem 5.1.2 e). +□ +We close this section with an application of Theorem 5.1.2 b) to the Fourier +transformation on the Beurling–Björck spaces Sµ(Rd,E) from Example 4.2.26. +5.1.5. Theorem. Let E be a locally complete lcHs over C and µ∶Rd → [0,∞) +continuous such that µ(x) = µ(−x) for all x ∈ Rd and condition (γ) is fulfilled. +(i) If E has metric ccp, or +(ii) if µ ∈ C1(Rd) and there are k ∈ N0, C > 0 such that ∣∂enµ(x)∣ ≤ Cekµ(x) +for all x ∈ Rd and 1 ≤ n ≤ d, +then FE∶Sµ(Rd,E) → Sµ(Rd,E) is an isomorphism with FE = S ○ (FCεidE) ○ S−1. +Proof. Due to Example 4.2.26 Sµ(Rd) and Sµ(Rd,E) are ε-compatible. The +Fourier transformation FC∶Sµ(Rd) → Sµ(Rd) is a well-defined isomorphism by the +definition of Sµ(Rd) and since (FC ○ FC)(f)(x) = f(−x) for all f ∈ Sµ(Rd) as well +as µ(x) = µ(−x) for all x ∈ Rd. Due to (26) with β = 0 we have that (FE,FC) is a +consistent family for (Sµ,E) and thus it follows from Theorem 5.1.2 b) that FE is +an isomorphism and FE = S ○ (FCεidE) ○ S−1, which completes the proof. +□ +5.2. Extension of vector-valued functions +We study the problem of extending vector-valued functions via the existence +of weak extensions in this section. The precise description of this problem reads +as follows. Let E be a locally convex Hausdorff space over the field K of real or +complex numbers and F(Ω) ∶= F(Ω,K) a locally convex Hausdorff space of K- +valued functions on a set Ω. Suppose that the point evaluations δx belong to the +dual F(Ω)′ for every x ∈ Ω and that there is a locally convex Hausdorff space +F(Ω,E) of E-valued functions on Ω such that the map +S∶F(Ω)εE → F(Ω,E), u �→ [x ↦ u(δx)], +(44) +is an isomorphism into, i.e. F(Ω) and F(Ω,E) are ε-into-compatible. Thus F(Ω)εE +is a linearisation of a subspace of F(Ω,E). Linearisations that are based on the +Dixmier–Ng theorem were used by Bonet, Domański and Lindström in [28, Lemma +10, p. 243] resp. Laitila and Tylli in [121, Lemma 5.2, p. 14] to describe the space +of weakly holomorphic resp. harmonic functions on the unit disc Ω = D ⊂ C with +values in a (complex) Banach space E (see also [118]). +5.2.1. Question. Let Λ be a subset of Ω and G a linear subspace of E′. Let +f∶Λ → E be such that for every e′ ∈ G, the function e′ ○ f∶Λ → K has an extension +in F(Ω). When is there an extension F ∈ F(Ω,E) of f, i.e. F∣Λ = f ? + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +75 +An affirmative answer for Λ = Ω and G = E′ is called a weak-strong principle. +For weighted continuous functions on a completely regular Hausdorff space Ω with +values in a semi-Montel or Schwartz space E a weak-strong principle is given by +Bierstedt in [17, 2.10 Lemma, p. 140]. +Weak-strong principles for holomorphic +functions on open subsets Ω ⊂ C were shown by Dunford in [59, Theorem 76, p. +354] for Banach spaces E and by Grothendieck in [82, Théorème 1, p. 37–38] for +quasi-complete E. For a wider class of function spaces weak-strong principles are +due to Grothendieck, mainly, in the case that F(Ω) is nuclear and E complete (see +[83, Chap. II, §3, n○3, Théorème 13, p. 80]), which covers the case that F(Ω) is the +space C∞(Ω) of smooth functions on an open set Ω ⊂ Rd (with its usual topology). +Gramsch [77] analysed the weak-strong principles of Grothendieck and realised +that they can be used to extend functions if Λ is a set of uniqueness, i.e. from +f ∈ F(Ω) and f(x) = 0 for all x ∈ Λ follows that f = 0, and F(Ω) a semi-Montel +space, E complete and G = E′ (see [77, 0.1, p. 217]). +An extension result for +holomorphic functions where G = E′ and E is sequentially complete was shown by +Bogdanowicz in [25, Corollary 3, p. 665]. +Grosse-Erdmann proved for holomorphic functions on Λ = Ω in [79, 5.2 The- +orem, p. 35] that it is sufficient to test locally bounded functions f with values +in a locally complete space E with functionals from a weak⋆-dense subspace G of +E′. Arendt and Nikolski [7, 8] shortened his proof in the case that E is a Fréchet +space (see [7, Theorem 3.1, p. 787] and [7, Remark 3.3, p. 787]). Arendt gave an +affirmative answer in [6, Theorem 5.4, p. 74] for harmonic functions on an open +subset Λ = Ω ⊂ Rd where the range space E is a Banach space and G a weak⋆-dense +subspace of E′. +In [77] Gramsch also derived extension results for a large class of Fréchet– +Montel spaces F(Ω) in the case that Λ is a special set of uniqueness, E sequentially +complete and G strongly dense in E′ (see [77, 3.3 Satz, p. 228–229]). He applied +it to the space of holomorphic functions and Grosse-Erdmann [81] expanded this +result to the case of E being Br-complete and G only a weak⋆-dense subspace of +E′ (see [81, Theorem 2, p. 401] and [81, Remark 2 (a), p. 406]). In a series of +papers [30, 69, 70, 92, 93] these results were generalised and improved by Bonet, +Frerick, Jordá and Wengenroth who used (44) to obtain extensions for vector-valued +functions via extensions of linear operators. In [92, 93] this was done by Jordá for +holomorphic functions on a domain (i.e. open and connected) Ω ⊂ C and weighted +holomorphic functions on a domain Ω in a Banach space. In [30] this was done +by Bonet, Frerick and Jordá for closed subsheaves F(Ω) of the sheaf of smooth +functions C∞(Ω) on a domain Ω ⊂ Rd. Their results implied some consequences +on the work of Bierstedt and Holtmanns [18] as well. Further, in [69] this was +done by Frerick and Jordá for closed subsheaves F(Ω) of smooth functions on a +domain Ω ⊂ Rd which are closed in the sheaf C(Ω) of continuous functions and in +[70] by the first two authors and Wengenroth in the case that F(Ω) is the space of +bounded functions in the kernel of a hypoelliptic linear partial differential operator, +in particular, the spaces of bounded holomorphic or harmonic functions. +In this section we present a unified approach to the extension problem for a large +class of function spaces. The spaces we treat are usually of the kind that F(Ω) +belongs to the class of semi-Montel spaces, Fréchet–Schwartz spaces or Banach +spaces. +Even quite general weighted spaces F(Ω) are treated, at least, if E is +a semi-Montel space. +Our approach is based on three ideas. +First, it is based +on the representation of (a subspace of) F(Ω,E) as a space of continuous linear +operators via the map S from (44). We note that almost all our examples of such +spaces F(Ω,E) are actually of the form of a general weighted space FV(Ω,E) from +Definition 3.1.4. Second, it is based on the idea to consider a set of uniqueness Λ + +76 +5. APPLICATIONS +not necessarily as a subset of Ω but rather as a set of functionals acting on F(Ω). +In the definition of a set of uniqueness given above one may identify Λ with the set +of functionals {δx ∣ x ∈ Λ} and this shift of perspective allows us to consider certain +sets of functionals of the form T K +m,x as sets of uniqueness for F(Ω) (see Definition +5.2.2). Third, the generalised concept of consistency and strength of a family of +operators (T E +m,T K +m)m∈M acting on (F(Ω,E),F(Ω)) from Definition 4.3.1 enables +us to generalise Question 5.2.1 and affirmatively answer this generalised question. +These three ideas are used to extend the mentioned results and we always have +to balance the sets Λ from which we extend our functions and the subspaces G ⊂ E′ +with which we test. The case of ‘thin’ sets Λ and ‘thick’ subspaces G is handled in +Section 5.2.1, the converse case of ‘thick’ sets Λ and ‘thin’ subspaces G in Section +5.2.2. +5.2.1. Extension from thin sets. Using the functionals T K +m,x, we extend the +definition of a set of uniqueness and a space of restrictions given in [30, Definition +4, 5, p. 230]. This prepares the ground for a generalisation of Question 5.2.1 using +a strong, consistent family (T E +m,T K +m)m∈M. +5.2.2. Definition (set of uniqueness). Let Ω be a non-empty set, F(Ω) ⊂ KΩ +an lcHs, (ωm)m∈M be a family of non-empty sets and T K +m∶F(Ω) → Kωm be linear for +all m ∈ M. Then U ⊂ ⋃m∈M({m}×ωm) is called a set of uniqueness for (T K +m,F)m∈M +if +(i) ∀ (m,x) ∈ U ∶ T K +m,x ∈ F(Ω)′, +(ii) ∀ f ∈ F(Ω) ∶ [∀(m,x) ∈ U ∶ T K +m(f)(x) = 0] ⇒ f = 0. +We omit the index m in ωm and T K +m if M is a singleton and consider U as a subset +of ω. +If U is a set of uniqueness for (T K +m,F)m∈M, then span{T K +m,x ∣ (m,x) ∈ U} is +dense in F(Ω)′ +σ (and F(Ω)′ +κ) by the bipolar theorem. +5.2.3. Remark. Let Ω be a non-empty set and F(Ω) ⊂ KΩ an lcHs. +a) A simple set of uniqueness for (idKΩ,F) is given by U ∶= Ω if δx ∈ F(Ω)′ +for all x ∈ Ω. +b) If F(Ω) has a Schauder basis (fn)n∈N with associated sequence of coef- +ficient functionals T K ∶= (T K +n )n∈N, then U ∶= N is a set of uniqueness for +(T K,F). +An example for b) is the space of holomorphic functions on an open disc +Dr(z0) ⊂ C with radius 0 < r ≤ ∞ and center z0 ∈ C. +If we equip this space +with the topology of compact convergence, then it has the shifted monomials +((⋅ − z0)n)n∈N0 as a Schauder basis with the point evaluations (δz0 ○ ∂n +C)n∈N0 given +by (δz0 ○ ∂n +C)(f) ∶= f (n)(z0) as associated sequence of coefficient functionals. We +will explore further sets of uniqueness for concrete function spaces in the upcoming +examples and come back to b) in Section 5.7. +5.2.4. Definition (restriction space). Let G ⊂ E′ be a separating subspace +and U a set of uniqueness for (T K +m,F)m∈M. Let FG(U,E) be the space of functions +f∶U → E such that for every e′ ∈ G there is fe′ ∈ F(Ω) with T K +m(fe′)(x) = (e′ ○ +f)(m,x) for all (m,x) ∈ U. +5.2.5. Remark. Since U is a set of uniqueness, the functions fe′ are unique +and the map Rf∶E′ → F(Ω), Rf(e′) ∶= fe′, is well-defined and linear. The map Rf +resembles the map Rf defined above Lemma 3.2.1. +5.2.6. Remark. Let F(Ω) and F(Ω,E) be ε-into-compatible. Consider a set +of uniqueness U for (T K +m,F)m∈M, a separating subspace G ⊂ E′ and a strong, + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +77 +consistent family (T E +m,T K +m)m∈M for (F,E). For u ∈ F(Ω)εE set f ∶= S(u). Then +f ∈ F(Ω,E) by the ε-into-compatibility and we set ̃f∶U → E, ̃f(m,x) ∶= T E +m(f)(x). +It follows that +(e′ ○ ̃f)(m,x) = (e′ ○ T E +m(f))(x) = T K +m(e′ ○ f)(x) +for all (m,x) ∈ U and fe′ ∶= e′ ○f ∈ F(Ω) for all e′ ∈ E′ by the strength of the family. +We conclude that ̃f ∈ FG(U,E). +5.2.7. Remark. If U is a set of uniqueness for (T K +m,F)m∈M, then the existence +of operators (T E +m)m∈M such that (T E +m,T K +m)m∈M is a strong, consistent family for +(F,E) is often guaranteed by the Riesz–Markov–Kakutani representation theorems +in Section 4.3. +Under the assumptions of Remark 5.2.6 the map +RU,G∶S(F(Ω)εE) → FG(U,E), f ↦ (T E +m(f)(x))(m,x)∈U, +is well-defined. +The map RU,G is also linear since T E +m is linear for all m ∈ M. +Further, the strength of the defining family guarantees that RU,G is injective. +5.2.8. Proposition. Let F(Ω) and F(Ω,E) be ε-into-compatible, G ⊂ E′ a +separating subspace and U a set of uniqueness for (T K +m,F)m∈M. If (T E +m,T K +m)m∈M is +a strong family for (F,E), then the map +T E∶F(Ω,E) → EU, f ↦ (T E +m(f)(x))(m,x)∈U, +is injective, in particular, RU,G is injective. +Proof. Let f ∈ F(Ω,E) with T E(f) = 0. Then +0 = (e′ ○ T E(f))(m,x) = (e′ ○ T E +m(f))(x) = T K +m(e′ ○ f)(x), +(m,x) ∈ U, +and e′ ○ f ∈ F(Ω) for all e′ ∈ E′ by the strength of the family. Since U is a set of +uniqueness, we get that e′ ○ f = 0 for all e′ ∈ E′, which implies f = 0. +□ +5.2.9. Question. Let F(Ω) and F(Ω,E) be ε-into-compatible, G ⊂ E′ a sepa- +rating subspace, (T E +m,T K +m)m∈M a strong family for (F,E) and U a set of uniqueness +for (T K +m,F)m∈M. When is the injective restriction map +RU,G∶S(F(Ω)εE) → FG(U,E), f ↦ (T E +m(f)(x))(m,x)∈U, +surjective? +The Question 5.2.1 is a special case of this question if there is a set of uniqueness +U for (T K +m,F)m∈M with {T K +m,x ∣ (m,x) ∈ U} = {δx ∣ x ∈ Λ}, Λ ⊂ Ω. We observe that +a positive answer to the surjectivity of RΩ,G results in the following weak-strong +principle. +5.2.10. Proposition. Let F(Ω) and F(Ω,E) be ε-into-compatible, G ⊂ E′ a +separating subspace such that e′ ○ f ∈ F(Ω) for all e′ ∈ G and f ∈ F(Ω,E). If +RΩ,G∶S(F(Ω)εE) → FG(Ω,E), f ↦ f, +with the set of uniqueness Ω for (idKΩ,F) is surjective, then +F(Ω)εE ≅ F(Ω,E) +via S +and +F(Ω,E) = {f∶Ω → E ∣ ∀ e′ ∈ G ∶ e′ ○ f ∈ F(Ω)}. +Proof. From the ε-into-compatibility and the surjectivity of RΩ,G we obtain +{f∶Ω → E ∣ ∀ e′ ∈ G ∶ e′ ○ f ∈ F(Ω)} = FG(Ω,E) = S(F(Ω)εE) ⊂ F(Ω,E). +Further, the assumption that e′ ○ f ∈ F(Ω) for all e′ ∈ G and f ∈ F(Ω,E), implies +that F(Ω,E) is a subspace of the space on the left-hand side, which proves our +statement, in particular, the surjectivity of S. +□ + +78 +5. APPLICATIONS +To answer Question 5.2.9 for general sets of uniqueness we have to restrict to +a certain class of separating subspaces of E′. +5.2.11. Definition (determine boundedness [30, p. 230]). A linear subspace +G ⊂ E′ determines boundedness if every σ(E,G)-bounded set B ⊂ E is already +bounded in E. +In [67, p. 139] such a space G is called uniform boundedness deciding by Fer- +nández et al. and in [134, p. 63] w∗-thick by Nygaard if E is a Banach space. +5.2.12. Remark. +a) Let E be an lcHs. Then G ∶= E′ determines bound- +edness by [131, Mackey’s theorem 23.15, p. 268]. +b) Let X be a barrelled lcHs, Y an lcHs and E ∶= Lb(X,Y ). For x ∈ X and +y′ ∈ Y ′ we set δx,y′∶L(X,Y ) → K, T ↦ y′(T(x)), and G ∶= {δx,y′ ∣ x ∈ +X, y′ ∈ Y ′} ⊂ E′. Then the span of G determines boundedness (in E) by +Mackey’s theorem and the uniform boundedness principle. For Banach +spaces X,Y this is already observed in [30, Remark 11, p. 233] and, if in +addition Y = K, in [7, Remark 1.4 b), p. 781]. +c) Further examples and a characterisation of subspaces G ⊂ E′ that de- +termine boundedness can be found in [7, Remark 1.4, p. 781–782], [134, +Theorem 1.5, p. 63–64] and [134, Theorem 2.3, 2.4, p. 67–68] in the case +that E is a Banach space. +F(Ω) a semi-Montel space and E (sequentially) complete. Our next +results are in need of spaces F(Ω) such that closed graph theorems hold with +Banach spaces as domain spaces and F(Ω) as the range space. Let us formally +define this class of spaces. +5.2.13. Definition (BC-space [142, p. 395]). We call an lcHs F a BC-space +if for every Banach space X and every linear map f∶X → F with closed graph in +X × F, one has that f is continuous. +A characterisation of BC-spaces is given by Powell in [142, 6.1 Corollary, p. 400– +401]. Since every Banach space is ultrabornological and barrelled, the [131, Closed +graph theorem 24.31, p. 289] of de Wilde and the Pták–K¯omura–Adasch–Valdivia +closed graph theorem [103, §34, 9.(7), p. 46] imply that webbed spaces and Br- +complete spaces are BC-spaces. We recall that an lcHs F is said to be Br-complete +if every σ(F ′,F)-dense σf(F ′,F)-closed linear subspace of F ′ equals F ′ where +σf(F ′,F) is the finest topology coinciding with σ(F ′,F) on all equicontinuous sets +in F ′ (see [103, §34, p. 26]). An lcHs F is called B-complete if every σf(F ′,F)-closed +linear subspace of F ′ is weakly closed. In particular, B-complete spaces are Br- +complete and every Br-complete space is complete by [103, §34, 2.(1), p. 26]. These +definitions are equivalent to the original definitions of Br- and B-completeness by +Pták [143, Definition 2, 5, p. 50, 55] due to [103, §34, 2.(2), p. 26–27] and we +note that they are also called infra-Pták spaces and Pták spaces, respectively. In +particular, Fréchet spaces are B-complete by [89, 9.5.2 Krein–˘Smulian Theorem, p. +184] but we will encounter non-Fréchet B-complete spaces as well. +The following proposition is a modification of [94, Satz 10.6, p. 237] and uses +the map Rf∶e′ ↦ fe′ from Remark 5.2.5. +5.2.14. Proposition. Let U be a set of uniqueness for (T K +m,F)m∈M and F(Ω) +a BC-space. Then Rf(B○ +α) is bounded in F(Ω) for every f ∈ FE′(U,E) and α ∈ A +where Bα ∶= {x ∈ E ∣ pα(x) < 1}. In addition, if F(Ω) is a semi-Montel space, then +Rf(B○ +α) is relatively compact in F(Ω). + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +79 +Proof. Let f ∈ FE′(U,E) and α ∈ A. The polar B○ +α is compact in E′ +σ and +thus E′ +B○α is a Banach space by [131, Corollary 23.14, p. 268]. +We claim that +the restriction of Rf to E′ +B○α has closed graph. Indeed, let (e′ +τ) be a net in E′ +B○α +converging to e′ in E′ +B○α and Rf(e′ +τ) converging to g in F(Ω). For (m,x) ∈ U we +note that +T K +m,x(Rf(e′ +τ)) = T K +m(fe′τ )(x) = (e′ +τ ○ f)(m,x) → (e′ ○ f)(m,x) = T K +m(fe′)(x) += T K +m(Rf(e′))(x). +The left-hand side converges to T K +m,x(g) since T K +m,x ∈ F(Ω)′ for all (m,x) ∈ U. +Hence we have T K +m(g)(x) = T K +m(Rf(e′))(x) for all (m,x) ∈ U. From U being a +set of uniqueness follows that g = Rf(e′). Thus the restriction of Rf to E′ +B○α has +closed graph and is continuous since F(Ω) is a BC-space. This yields that Rf(B○ +α) +is bounded as B○ +α is bounded in E′ +B○α. If F(Ω) is also a semi-Montel space, then +Rf(B○ +α) is even relatively compact. +□ +Now, we are ready to prove our first extension theorem. Its proof of surjectivity +of RU,E′ is just an adaptation of the proof of surjectivity of S given in Theorem +3.2.4. Let U be a set of uniqueness for (T K +m,F)m∈M. For f ∈ FE′(U,E) we consider +the dual map +Rt +f∶F(Ω)′ → E′⋆, Rt +f(y)(e′) ∶= y(fe′), +where E′⋆ is the algebraic dual of E′. Further, we recall the notation J ∶E → E′⋆ +for the canonical injection. +5.2.15. Theorem. Let F(Ω) and F(Ω,E) be ε-into-compatible, (T E +m,T K +m)m∈M +a strong, consistent family for (F,E), F(Ω) a semi-Montel BC-space and U a set +of uniqueness for (T K +m,F)m∈M. If +(i) E is complete, or +(ii) E is sequentially complete and for every f ∈ FE′(U,E) and f ′ ∈ F(Ω)′ +there is a sequence (f ′ +n)n∈N in F(Ω)′ converging to f ′ in F(Ω)′ +κ such that +Rt +f(f ′ +n) ∈ J (E) for every n ∈ N, +then the restriction map RU,E′∶S(F(Ω)εE) → FE′(U,E) is surjective. +Proof. Let f ∈ FE′(U,E). +As in Theorem 3.2.4 we equip J (E) with the +system of seminorms given by +pB○α(J (x)) ∶= sup +e′∈B○α +∣J (x)(e′)∣ = pα(x), +x ∈ E, +(45) +for all α ∈ A where Bα ∶= {x ∈ E ∣ pα(x) < 1}. We claim Rt +f ∈ L(F(Ω)′ +κ,J (E)). +Indeed, we have for y ∈ F(Ω)′ +pB○α(Rt +f(y)) = sup +e′∈B○α +∣y(fe′)∣ = +sup +x∈Rf (B○α) +∣y(x)∣ ≤ sup +x∈Kα +∣y(x)∣ +(46) +where Kα ∶= Rf(B○α). +Due to Proposition 5.2.14 the set Rf(B○ +α) is absolutely +convex and relatively compact, implying that Kα is absolutely convex and compact +in F(Ω) by [89, 6.2.1 Proposition, p. 103]. Further, we have for all e′ ∈ E′ and +(m,x) ∈ U +Rt +f(T K +m,x)(e′) = T K +m,x(fe′) = (e′ ○ f)(m,x) = J (f(m,x))(e′) +(47) +and thus Rt +f(T K +m,x) ∈ J (E). +First, let condition (i) be satisfied, i.e. let E be complete, and f ′ ∈ F(Ω)′. The +span of {T K +m,x ∣ (m,x) ∈ U} is dense in F(Ω)′ +κ since U is a set of uniqueness for + +80 +5. APPLICATIONS +F(Ω). Thus there is a net (f ′ +τ) converging to f ′ in FV(Ω)′ +κ with Rt +f(f ′ +τ) ∈ J (E) +and +pB○α(Rt +f(f ′ +τ) − Rt +f(f ′)) ≤ +(46) sup +x∈Kα +∣(f ′ +τ − f ′)(x)∣ → 0 +(48) +for all α ∈ A. We gain that (Rt +f(f ′ +τ)) is a Cauchy net in the complete space J (E). +Hence it has a limit g ∈ J (E) which coincides with Rt +f(f ′) since +pB○α(g − Rt +f(f ′)) ≤ pB○α(g − Rt +f(f ′ +τ)) + pB○α(Rt +f(f ′ +τ) − Rt +f(f ′)) +≤ +(48)pB○α(g − Rt +f(f ′ +τ)) + sup +x∈Kα +∣(f ′ +τ − f ′)(x)∣ → 0 +for all α ∈ A. We conclude that Rt +f(f ′) ∈ J (E) for every f ′ ∈ F(Ω)′. +Second, let condition (ii) be satisfied and f ′ ∈ F(Ω)′. Then there is a sequence +(f ′ +n) in F(Ω)′ converging to f ′ in F(Ω)′ +κ such that Rt +f(f ′ +n) ∈ J (E) for every +n ∈ N. From (46) we derive that (Rt +f(f ′ +n)) is a Cauchy sequence in the sequentially +complete space J (E) converging to Rt +f(f ′) ∈ J (E). +Therefore we obtain in both cases that Rt +f ∈ L(F(Ω)′ +κ,J (E)). So we get for +all α ∈ A and y ∈ F(Ω)′ +pα((J −1 ○ Rt +f)(y)) = +(45) pB○α(J ((J −1 ○ Rt +f)(y))) = pB○α(Rt +f(y)) ≤ +(46) sup +x∈Kα +∣y(x)∣. +This implies J −1 ○ Rt +f ∈ L(F(Ω)′ +κ,E) = F(Ω)εE (as linear spaces). We set F ∶= +S(J −1 ○ Rt +f) and obtain from consistency that +T E +m(F)(x) = T E +mS(J −1 ○Rt +f)(x) = J −1(Rt +f(T K +m,x)) = +(47) J −1(J (f(m,x))) = f(m,x) +for every (m,x) ∈ U, which means RU,E′(F) = f. +□ +If E is complete and U a set of uniqueness for (T K +m,F)m∈M with {T K +m,x ∣ (m,x) ∈ +U} = {δx ∣ x ∈ Λ}, Λ ⊂ Ω, then we get [77, 0.1, p. 217] as a special case. Condition +(i) and (ii) are adaptations of Condition 3.2.3 a) and c) from FV(Ω,E) and Rf to +FE′(U,E) and Rf. We also treat an adaptation of Condition 3.2.3 e) in Theorem +5.2.52. Condition 3.2.3 b) and d) may be adapted as well but we restrict to the +ones we actually apply. First, we apply Theorem 5.2.15 to the space of bounded +zero-solutions of a hypoelliptic linear partial differential operator equipped with the +strict topology β from Proposition 4.2.24. +5.2.16. Proposition. Let Ω ⊂ Rd be open and P(∂)K a hypoelliptic linear +partial differential operator. +Then (C∞ +P (∂),b(Ω),β) is a B-complete semi-Montel +space. +Proof. Due to the proof of Proposition 4.2.24 we know that β coincides with +the mixed topology γ(τc,∥ ⋅ ∥∞). It is easy to check that the closed ∥ ⋅ ∥∞-unit ball +B∥⋅∥∞ is τc-compact in C∞ +P (∂),b(Ω). Thus [46, Section I.1, 1.13 Proposition, p. 11] +yields that (C∞ +P (∂),b(Ω),β) is a semi-Montel space. From [150, 2.9 Theorem, p. 185] +it follows that the space is B-complete. +□ +5.2.17. Corollary. Let Ω ⊂ Rd be open, E a complete lcHs, P(∂)K a hypoel- +liptic linear partial differential operator, (T E +m,T K +m)m∈M a strong, consistent family +for ((C∞ +P (∂),b(Ω),β),E) and U a set of uniqueness for (T K +m,(C∞ +P (∂),b(Ω),β))m∈M. +If f∶U → E is a function such that there is fe′ ∈ C∞ +P (∂),b(Ω) for each e′ ∈ E′ +with T K +m(fe′)(x) = (e′ ○ f)(m,x) for all (m,x) ∈ U, then there is a unique F ∈ +C∞ +P (∂),b(Ω,E) with T E +m(F)(x) = f(m,x) for all (m,x) ∈ U. + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +81 +Proof. The space (C∞ +P (∂),b(Ω),β) is a semi-Montel BC-space by Proposition +5.2.16. Moreover, (C∞ +P (∂),b(Ω),β) and (C∞ +P (∂),b(Ω,E),β) are ε-compatible by Propo- +sition 4.2.24, yielding our statement by Theorem 5.2.15 (i) and Proposition 5.2.8. +□ +Especially, for any m ∈ N0 the family ((∂β)E,(∂β)K)β∈Nd +0,∣β∣≤m is strong and con- +sistent for ((C∞ +P (∂),b(Ω),β),E) by the proof of Proposition 4.2.24. It is always pos- +sible to construct a strong, consistent family (T E +m,T K +m)m∈M for ((C∞ +P (∂),b(Ω),β),E) +from a given set of uniqueness (T K +m,(C∞ +P (∂),b(Ω),β))m∈M due to Remark 4.3.11 b) +and c). +Similarly, we may apply Theorem 5.2.15 to the space E{Mp}(Ω,E) of ultradiffer- +entiable functions of class {Mp} of Roumieu-type from Example 3.1.9 f). E{Mp}(Ω) +is a projective limit of a countable sequence of DFS-spaces by [99, Theorem 2.6, +p. 44] and thus webbed because being webbed is stable under the formation of +projective and inductive limits of countable sequences by [89, 5.3.3 Corollary, p. +92]. Further, if the sequence (Mp)p∈N0 satisfies Komatsu’s conditions (M.1) and +(M.3)’, then E{Mp}(Ω) is a Montel space by [99, Theorem 5.12, p. 65–66]. The +spaces E{Mp}(Ω) and E{Mp}(Ω,E) are ε-compatible if (M.1) and (M.3)’ hold and +E is complete by Example 3.2.11 b). Hence Theorem 5.2.15 (i) is applicable. +5.2.18. Remark. We remark that Remark 5.2.6 and Theorem 5.2.15 still hold +if the map S∶F(Ω)εE → F(Ω,E) is only a linear isomorphism into, i.e. an isomor- +phism into of linear spaces, since the topological nature of ε-into-compatibility is +not used in the proof. In particular, this means that it can be applied to the space +M(Ω,E) of meromorphic functions on an open, connected set Ω ⊂ C with values +in an lcHs E over C (see [29, p. 356]). The space M(Ω) is a Montel LF-space, +thus webbed by [89, 5.3.3 Corollary (b), p. 92], due to the proof of [80, Theorem +3 (a), p. 294–295] if it is equipped with the locally convex topology τML given in +[80, p. 292]. By [29, Proposition 6, p. 357] the map S∶M(Ω)εE → M(Ω,E) is +a linear isomorphism if E is locally complete and does not contain the space CN. +Therefore we can apply Theorem 5.2.15 if E is complete and does not contain CN. +This augments [92, Theorem 12, p. 12] where E is assumed to be locally complete +with suprabarrelled strong dual and (T E,T C) = (idEΩ,idCΩ). +F(Ω) a Fréchet–Schwartz space and E locally complete. We recall the +following abstract extension result. +5.2.19. Proposition ([30, Proposition 7, p. 231]). Let E be a locally complete +lcHs, Y a Fréchet–Schwartz space, X ⊂ Y ′ +b (= Y ′ +κ) dense and A∶X → E linear. Then +the following assertions are equivalent: +a) There is a (unique) extension ̂A ∈ Y εE of A. +b) (At)−1(Y ) (= {e′ ∈ E′ ∣ e′ ○ A ∈ Y }) determines boundedness in E. +Next, we generalise [30, Theorem 9, p. 232] using the preceding proposition. +The proof of the generalisation is simply obtained by replacing the set of uniqueness +in the proof of [30, Theorem 9, p. 232] by our more general set of uniqueness. +5.2.20. Theorem. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness and F(Ω) and F(Ω,E) be ε-into-compatible. Let (T E +m,T K +m)m∈M be a strong, +consistent family for (F,E), F(Ω) a Fréchet–Schwartz space and U a set of unique- +ness for (T K +m,F)m∈M. Then the restriction map RU,G∶S(F(Ω)εE) → FG(U,E) is +surjective. + +82 +5. APPLICATIONS +Proof. Let f ∈ FG(U,E). +We choose X ∶= span{T K +m,x ∣ (m,x) ∈ U} and +Y ∶= F(Ω). Let A∶X → E be the linear map generated by A(T K +m,x) ∶= f(m,x). +The map A is well-defined since G is σ(E′,E)-dense. Let e′ ∈ G and fe′ be the +unique element in F(Ω) such that T K +m(fe′)(x) = (e′ ○ A)(T K +m,x) for all (m,x) ∈ +U. +This equation allows us to consider fe′ as a linear form on X (by setting +fe′(T K +m,x) ∶= (e′ ○ A)(T K +m,x)), which yields e′ ○ A ∈ F(Ω) for all e′ ∈ G. It follows +that G ⊂ (At)−1(Y ), implying that (At)−1(Y ) determines boundedness. Applying +Proposition 5.2.19, there is an extension ̂A ∈ F(Ω)εE of A and we set F ∶= S(̂A). +We note that +T E +m(F)(x) = T E +mS(̂A)(x) = ̂A(T K +m,x) = A(T K +m,x) = f(m,x) +for all (m,x) ∈ U by consistency, yielding RU,G(F) = f. +□ +Let us apply the preceding theorem to our weighted spaces of continuously +partially differentiable functions and its subspaces from Example 3.1.9 and Example +4.2.22. +5.2.21. Corollary. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness, V∞ a directed family of weights which is locally bounded away from zero on +an open set Ω ⊂ Rd, let F(Ω) be a Fréchet–Schwartz space and U ⊂ Nd +0 × Ω a set +of uniqueness for (∂β,F)β∈Nd +0 where F stands for CV∞, CV∞ +0 , CV∞ +P (∂) or CV∞ +P (∂),0. +Then the following holds: +a) If f∶U → E is a function such that there is fe′ ∈ F(Ω) for each e′ ∈ G +with ∂βfe′(x) = (e′ ○ f)(β,x) for all (β,x) ∈ U, then there is a unique +F ∈ F(Ω,E) with (∂β)EF(x) = f(β,x) for all (β,x) ∈ U. +b) If U ⊂ Ω and f∶U → E is a function such that e′ ○ f admits an extension +fe′ ∈ F(Ω) for every e′ ∈ G, then there is a unique extension F ∈ F(Ω,E) +of f. +c) F(Ω,E) = {f∶Ω → E ∣ ∀ e′ ∈ G ∶ e′ ○ f ∈ F(Ω)}. +Proof. The strength and consistency of ((∂β)E,∂β)β∈Nd +0 for (F,E) and the ε- +compatibility of F(Ω) and F(Ω,E) follow from Example 3.2.7 e)+f) and Example +4.2.22 c)+f). This implies that part a) and its special case part b) hold by Theorem +5.2.20 and Proposition 5.2.8. Part c) follows from part b) and Proposition 5.2.10 +since U ∶= Ω is a set of uniqueness for (idKΩ,F). +□ +5.2.22. Remark. Let V∞ be a directed family of weights which is locally +bounded away from zero on an open set Ω ⊂ Rd. +a) Then any dense set U ⊂ Ω is a set of uniqueness for (idKΩ,F) with F = +CV∞, CV∞ +0 , CV∞ +P (∂) or CV∞ +P (∂),0 due to continuity. +b) Let Ω be connected and x0 ∈ Ω. +Then U ∶= {(en,x) ∣ 1 ≤ n ≤ d, x ∈ +Ω} ∪ {(0,x0)} is a set of uniqueness for (∂β,F)β∈N0 by the mean value +theorem with F from a). +c) Let K ∶= R, d ∶= 1, Ω ∶= (a,b) ⊂ R, g∶(a,b) → N and x0 ∈ (a,b). Then +U ∶= {(g(x),x) ∣ x ∈ (a,b)} ∪ {(n,x0) ∣ n ∈ N0} is a set of uniqueness for +(∂β,F)β∈N0 with F from a). Indeed, if f ∈ F(Ω) and 0 = ∂g(x)f(x) for all +x ∈ (a,b), then f is a polynomial by [56, Chap. 11, Theorem, p. 53]. If, in +addition, 0 = ∂nf(x0) for all n ∈ N0, then the polynomial f must vanish +on the whole interval Ω. +d) Let Ω ⊂ C be connected. Then any set U ⊂ Ω with an accumulation point +in Ω is a set of uniqueness for (idCΩ,CV∞ +∂ ) by the identity theorem for +holomorphic functions. + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +83 +e) Let Ω ⊂ C be connected and z0 ∈ Ω. Then U ∶= {(n,z0) ∣ n ∈ N0} is a set +of uniqueness for (∂n +C,CV∞ +∂ )n∈N0 by local power series expansion and the +identity theorem. +f) Let Ω ⊂ Rd be connected. +Then any non-empty open set U ⊂ Ω is a +set of uniqueness for (idKΩ,CV∞ +∆) by the identity theorem for harmonic +functions (see e.g. [85, Theorem 5, p. 218]). +g) Further examples of sets of uniqueness for (idKΩ,CV∞ +∆) are given in [98]. +In part e) a special case of Remark 5.2.3 b) is used, namely, that CW∞ +∂ (Dr(z0)) +has a Schauder basis with associated coefficient functionals (δz0 ○ ∂n +C)n∈N0 where +0 < r ≤ ∞ is such that Dr(z0) ⊂ Ω. In order to obtain some sets of uniqueness which +are more sensible w.r.t. the family of weights V∞, we turn to entire and harmonic +functions fulfilling some growth conditions. For a family V ∶= (νj)j∈N of continuous +weights on Rd set V∞ ∶= (νj,m)j∈N,m∈N0 where νj,m∶{β ∈ Nd +0 ∣ ∣β∣ ≤ m}×Rd → [0,∞), +νj,m(β,x) ∶= νj(x). We know that CVP (∂)(Rd,E) = CV∞ +P (∂)(Rd,E) as locally convex +spaces and that CVP (∂)(Rd) is a nuclear Fréchet space for P(∂) = ∂ or P(∂) = ∆ +by Proposition 4.2.19 if E is a locally complete lcHs and V fulfils Condition 4.2.18. +In particular, Condition 4.2.18 is fulfilled if νj(x) ∶= exp(−(τ + 1 +j )∣x∣), x ∈ Rd, for +all j ∈ N and some 0 ≤ τ < ∞ and thus we can apply Corollary 5.2.21 to the +spaces Aτ +∂(C,E) = CV∂(C,E) of entire and Aτ +∆(Rd,E) = CV∆(Rd,E) of harmonic +functions of exponential type τ by Remark 4.2.20. Hence we may complement our +list in Remark 5.2.22 by some more examples for spaces of functions of exponential +type 0 ≤ τ < ∞. +5.2.23. Remark. The following sets U ⊂ C are sets of uniqueness for (idCC,Aτ +∂). +a) If τ < π, then U ∶= N0 is a set of uniqueness by [21, 9.2.1 Carlson’s theorem, +p. 153]. +b) Let δ > 0 and (λn)n∈N ⊂ (0,∞) such that λn+1 − λn > δ for all n ∈ N. Then +U ∶= (λn)n∈N is a set of uniqueness if limsupr→∞ r−2τ/πψ(r) = ∞ where +ψ(r) ∶= exp(∑λn 0, by [21, 9.5.1 Fuchs’s theorem, p. 157–158]. +The following sets U are sets of uniqueness for (∂n +C,Aτ +∂)n∈N0. +c) Let (λn)n∈N0 ⊂ C with ∣λn∣ < 1 for all n ∈ N0. If τ < ln(2), then U ∶= +{(n,λn) ∣ n ∈ N0} is a set of uniqueness by [21, 9.11.1 Theorem, p. 172]. +If τ < ln(2 + +√ +3), then U ∶= {(2n + 1,0) ∣ n ∈ N0} ∪ {(2n,λn) ∣ n ∈ N0} is a +set of uniqueness by [21, 9.11.3 Theorem, p. 173]. +d) Let (λn)n∈N0 ⊂ C with limsupn→∞ n−1 ∑n +k=1 ∣λk∣ ≤ 1. +If τ < e−1, then +U ∶= {(n,λn) ∣ n ∈ N0} is a set of uniqueness by [21, 9.11.4 Theorem, p. +173]. +The following sets U ⊂ Rd are sets of uniqueness for (idRRd ,Aτ +∆). +e) Let d ∶= 2. If there is k ∈ N with τ < π/k, then U ∶= Z ∪ (Z + ik) is a set of +uniqueness by [22, Theorem 1, p. 425]. +f) Let d ∶= 2. If τ < π and θ ∉ πQ, then U ∶= Z ∪ (eiθZ) is a set of uniqueness +by [22, Theorem 2, p. 426]. +g) If τ < π, then U ∶= {0,1} × Zd−1 is a set of uniqueness by [145, Corollary +1.8, p. 312]. +h) If τ < π and a ∈ R with ∣a∣ ≤ +√ +1/(d − 1), then U ∶= Zd−1 × {0,a} is a set of +uniqueness by [185, Theorem A, p. 335]. +i) Further examples of sets of uniqueness can be found in [10]. +The following sets U are sets of uniqueness for ((∂β)R,Aτ +∆)β∈Nd +0. +j) If τ < π, then U ∶= {(β,(x,0)) ∣ β ∈ {0,ed}, x ∈ Zd−1} is a set of uniqueness +by [185, Theorem B, p. 335]. Further examples can be found in [10]. + +84 +5. APPLICATIONS +We need the following weak-strong principle in our last section for the space +E0(E) of E-valued infinitely continuously partially differentiable functions on (0,1) +such that all derivatives can be continuously extended to the boundary and vanish +at 1. +5.2.24. Corollary. Let E be a locally complete lcHs and G ⊂ E′ determine +boundedness. Then E0(E) = {f∶(0,1) → E ∣ ∀ e′ ∈ G ∶ e′ ○ f ∈ E0}. +Proof. By Example 4.2.29 E0 is a Fréchet–Schwartz space and E0 and E0(E) +are ε-compatible. We derive our statement from Theorem 5.2.20 and Proposition +5.2.10 with (T E,T K) ∶= (idEU ,idKU ) and U ∶= (0,1). +□ +Fν(Ω) a Banach space and E locally complete. In this subsection we +consider function spaces F(Ω,E) with a certain structure, namely, spaces FV(Ω,E) +from Definition 3.1.4 where the family of weights V = (νj,m)j∈J,m∈M only consists +of one weight function, i.e. the sets J and M can be chosen as singletons. So for +two non-empty sets Ω and ω, a weight ν∶ω → (0,∞), a linear operator T E∶EΩ ⊃ +domT E → Eω and a linear subspace AP(Ω,E) of EΩ we consider the space +Fν(Ω,E) = {f ∈ F(Ω,E) ∣ ∀ α ∈ A ∶ ∣f∣α < ∞} +where +F(Ω,E) = AP(Ω,E) ∩ domT E +and +∣f∣α = ∣f∣Fν(Ω),α = sup +x∈ω pα(T E(f)(x))ν(x). +For instance, if Ω ∶= ω, T E ∶= idEΩ and ν ∶= 1 on Ω, then Fν(Ω,E) is the +linear subspace of F(Ω,E) consisting of bounded functions. We use the methods +developed in [70, 93] where, in particular, the special case that Fν(Ω) is the space +of bounded smooth functions on an open set Ω ⊂ Rd in the kernel of a hypoelliptic +linear partial differential operator resp. a weighted space of holomorphic functions +on an open subset Ω of a Banach space is treated. The lack of compact subsets of an +infinite dimensional Banach space Fν(Ω) is compensated in [70, 93] by equipping +F(Ω) with a locally convex Hausdorff topology such that the closed unit ball of +Fν(Ω) is compact in F(Ω). Among others, the space F(Ω,E) ∶= (O(Ω,E),τc) of +holomorphic functions on an open set Ω ⊂ C with values in a locally complete space +E equipped with topology τc of compact convergence is used in [70] and the space +Fν(Ω,E) ∶= H∞(Ω,E) of E-valued bounded holomorphic functions on Ω. +5.2.25. Proposition. Let F(Ω) and F(Ω,E) be ε-into-compatible, (T E,T K) +a consistent family for (F,E) and a generator for (Fν,E) and the map i∶Fν(Ω) → +F(Ω), f ↦ f, continuous. We set +Fεν(Ω,E) ∶= S({u ∈ F(Ω)εE ∣ u(B○F (Ω)′ +Fν(Ω) ) is bounded in E}) +where B○F (Ω)′ +Fν(Ω) ∶= {y′ ∈ F(Ω)′ ∣ ∀ f ∈ BFν(Ω) ∶ ∣y′(f)∣ ≤ 1} and BFν(Ω) is the closed +unit ball of Fν(Ω). Then the following holds: +a) Fν(Ω) is a dom-space. +b) Let u ∈ F(Ω)εE. Then +sup +y′∈B○F (Ω)′ +Fν(Ω) +pα(u(y′)) = ∣S(u)∣Fν(Ω),α, +α ∈ A. +In particular, +Fεν(Ω,E) = S({u ∈ F(Ω)εE ∣ ∀ α ∈ A ∶ ∣S(u)∣Fν(Ω),α < ∞}). + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +85 +c) S(Fν(Ω)εE) ⊂ Fεν(Ω,E) ⊂ Fν(Ω,E) as linear spaces. +If F(Ω) and +F(Ω,E) are even ε-compatible, then Fεν(Ω,E) = Fν(Ω,E). +d) If Fν(Ω,E) is Hausdorff, then +(i) (T E,T K) is a consistent generator for (Fν,E). +(ii) Fν(Ω) and Fν(Ω,E) are ε-into-compatible. +(iii) (T E,T K) is a strong generator for (Fν,E) if it is a strong family for +(F,E). +Proof. Part a) follows from the continuity of the map i and the ε-into- +compatibility of F(Ω) and F(Ω,E). Let us turn to part b). As in Lemma 3.1.8 it +follows from the bipolar theorem that +B○F (Ω)′ +Fν(Ω) = acx{T K +x (⋅)ν(x) ∣ x ∈ ω}, +where acx denotes the closure w.r.t. κ(F(Ω)′,Fν(Ω)) of the absolutely convex hull +acx of the set D ∶= {T K +x (⋅)ν(x) ∣ x ∈ ω} on the right-hand side, and that +sup +y′∈B○F (Ω)′ +Fν(Ω) +pα(u(y′)) = +sup +y′∈acx(D) +pα(u(y′)) = sup +y′∈D +pα(u(y′)) = sup +x∈ω pα(u(T K +x ))ν(x) += sup +x∈ω pα(T E(S(u))(x))ν(x) = ∣S(u)∣Fν(Ω),α +by consistency, which proves part b). +Let us address part c). +The continuity of the map i implies the continuity +of the inclusion Fν(Ω)εE ↪ F(Ω)εE and thus we obtain u∣F (Ω)′ ∈ F(Ω)εE for +every u ∈ Fν(Ω)εE. If u ∈ Fν(Ω)εE and α ∈ A, then there are C0,C1 > 0 and an +absolutely convex compact set K ⊂ Fν(Ω) such that K ⊂ C1BFν(Ω) and +sup +y′∈B○F (Ω)′ +Fν(Ω) +pα(u(y′)) ≤ C0 +sup +y′∈B○F (Ω)′ +Fν(Ω) +sup +f∈K +∣y′(f)∣ ≤ C0C1, +which implies S(Fν(Ω)εE) ⊂ Fεν(Ω,E). If f ∶= S(u) ∈ Fεν(Ω,E) and α ∈ A, then +S(u) ∈ F(Ω,E) and +∣f∣Fν(Ω),α = sup +x∈ω pα(u(T K +x )ν(x)) < ∞ +by consistency, yielding Fεν(Ω,E) ⊂ Fν(Ω,E). If F(Ω) and F(Ω,E) are even +ε-compatible, then S(F(Ω)εE) = F(Ω,E), which yields Fεν(Ω,E) = Fν(Ω,E) by +part b). +Let us turn to part d). By part a) Fν(Ω) is a dom-space. Since Fν(Ω,E) is +Hausdorff, it is also a dom-space due to Remark 3.1.6 c). We have u∣F (Ω)′ ∈ F(Ω)εE +for every u ∈ Fν(Ω)εE and +SFν(Ω)(u)(x) = u(δx) = u∣F (Ω)′(δx) = SF (Ω)(u∣F (Ω)′)(x), +x ∈ Ω. +In combination with S(F(Ω)εE) ⊂ F(Ω,E) and the consistency of (T E,T K) for +(F,E) this yields that (T E,T K) is a consistent generator for (Fν,E). Thus part (i) +holds and implies part (ii) by Theorem 3.1.12. If (T E,T K) is in addition a strong +family for (F,E), then the inclusion Fν(Ω,E) ⊂ F(Ω,E) implies that e′ ○ f ∈ +F(Ω,E) and T K(e′ ○f)(x) = (e′ ○T E(f))(x) for all e′ ∈ E′, f ∈ Fν(Ω,E) and x ∈ ω. +It follows that (T E,T K) is a strong generator for (Fν,E). +□ +The canonical situation in part c) is that Fεν(Ω,E) and Fν(Ω,E) coincide +as linear spaces for locally complete E as we will encounter in the forthcoming +examples, e.g. if Fν(Ω,E) ∶= H∞(Ω,E) and F(Ω,E) ∶= (O(Ω,E),τc) for an open +set Ω ⊂ C. That all three spaces in part c) coincide is usually only guaranteed +by Corollary 3.2.5 (iii) if E is a semi-Montel space. Therefore the ‘mingle-mangle’ +space Fεν(Ω,E) is a good replacement for S(Fν(Ω)εE) for our purpose. + +86 +5. APPLICATIONS +5.2.26. Remark. Let (T E,T K) be a strong, consistent family for (F,E) and +a generator for (Fν,E). +Let F(Ω) and F(Ω,E) be ε-into-compatible and the +inclusion Fν(Ω) ↪ F(Ω) continuous. Consider a set of uniqueness U for (T K,Fν) +and a separating subspace G ⊂ E′. For u ∈ F(Ω)εE such that u(B○F (Ω)′ +Fν(Ω) ) is bounded +in E, i.e. S(u) ∈ Fεν(Ω,E), we set f ∶= S(u). Then f ∈ F(Ω,E) by the ε-into- +compatibility and we define ̃f∶U → E, ̃f(x) ∶= T E(f)(x). This yields +(e′ ○ ̃f)(x) = (e′ ○ T E(f))(x) = T K(e′ ○ f)(x) +(49) +for all x ∈ U and fe′ ∶= e′ ○ f ∈ F(Ω) for each e′ ∈ E′ by the strength of the family. +Moreover, T K +x (⋅)ν(x) ∈ B○F (Ω)′ +Fν(Ω) for every x ∈ ω, which implies that for every e′ ∈ E′ +there are α ∈ A and C > 0 such that +∣fe′∣Fν(Ω) = sup +x∈ω +∣e′(u(T K +x (⋅)ν(x))∣ ≤ C +sup +y′∈B○F (Ω)′ +Fν(Ω) +pα(u(y′)) < ∞ +by strength and consistency. Hence fe′ ∈ Fν(Ω) for every e′ ∈ E′ and ̃f ∈ FνG(U,E). +Under the assumptions of Remark 5.2.26 the map +RU,G∶Fεν(Ω,E) → FνG(U,E), f ↦ (T E(f)(x))x∈U, +(50) +is well-defined and linear. In addition, we derive from (49) that RU,G is injective +since U is a set of uniqueness and G ⊂ E′ separating. The replacement of Question +5.2.9 reads as follows. +5.2.27. Question. Let the assumptions of Remark 5.2.26 be fulfilled. When is +the injective restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E), f ↦ (T E(f)(x))x∈U, +surjective? +Due to Proposition 5.2.25 c) the Question 5.2.1 is a special case of this question +if Λ ⊂ Ω =∶ ω and U ∶= Λ is a set of uniqueness for (idKΩ,Fν). We recall the following +extension result for continuous linear operators. +5.2.28. Proposition ([70, Proposition 2.1, p. 691]). Let E be a locally complete +lcHs, G ⊂ E′ determine boundedness, Z a Banach space whose closed unit ball BZ +is a compact subset of an lcHs Y and X ⊂ Y ′ be a σ(Y ′,Z)-dense subspace. If +A∶X → E is a σ(X,Z)-σ(E,G)-continuous linear map, then there exists a (unique) +extension ̂A ∈ Y εE of A such that ̂A(B○Y ′ +Z +) is bounded in E where B○Y ′ +Z +∶= {y′ ∈ +Y ′ ∣ ∀ z ∈ BZ ∶ ∣y′(z)∣ ≤ 1}. +Now, we are able to generalise [70, Theorem 2.2, p. 691] and [93, Theorem 10, +p. 5]. +5.2.29. Theorem. Let E be a locally complete lcHs, G ⊂ E′ determine bounded- +ness and F(Ω) and F(Ω,E) be ε-into-compatible. Let (T E,T K) be a generator for +(Fν,E) and a strong, consistent family for (F,E), Fν(Ω) a Banach space whose +closed unit ball BFν(Ω) is a compact subset of F(Ω) and U a set of uniqueness for +(T K,Fν). Then the restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E) +is surjective. +Proof. Let f ∈ FνG(U,E). We set X ∶= span{T K +x ∣ x ∈ U}, Y ∶= F(Ω) and +Z ∶= Fν(Ω). The consistency of (T E,T K) for (F,E) yields that X ⊂ Y ′. From +U being a set of uniqueness of Z follows that X is σ(Z′,Z)-dense. Since BZ is a +compact subset of Y , it follows that Z is a linear subspace of Y and the inclusion + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +87 +Z ↪ Y is continuous, which yields y′ +∣Z ∈ Z′ for every y′ ∈ Y ′. Thus X is σ(Y ′,Z)- +dense. Let A∶X → E be the linear map determined by A(T K +x ) ∶= f(x). The map A +is well-defined since G is σ(E′,E)-dense. Due to +e′(A(T K +x )) = (e′ ○ f)(x) = T K +x (fe′) +for every e′ ∈ G and x ∈ U we have that A is σ(X,Z)-σ(E,G)-continuous. We +apply Proposition 5.2.28 and gain an extension ̂A ∈ Y εE of A such that ̂A(B○Y ′ +Z +) is +bounded in E. We set ̃F ∶= S(̂A) ∈ Fεν(Ω,E) and get for all x ∈ U that +T E( ̃F)(x) = T ES(̂A)(x) = ̂A(T K +x ) = f(x) +by consistency for (F,E), implying RU,G( ̃F) = f. +□ +Let Ω ⊂ Rd be open, E an lcHs and P(∂)E∶C∞(Ω,E) → C∞(Ω,E) a linear +partial differential operator which is hypoelliptic if E = K. We consider the weighted +space CVP (∂)(Ω,E) of zero solutions from Proposition 4.2.14 where the familiy of +weights V only consists of one continuous weight ν∶Ω → (0,∞), i.e. the space +CνP (∂)(Ω,E) = {f ∈ C∞ +P (∂)(Ω,E) ∣ ∀ α ∈ A ∶ ∣f∣ν,α ∶= sup +x∈Ω +pα(f(x))ν(x) < ∞}. +5.2.30. Corollary. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness, Ω ⊂ Rd open, P(∂)K a hypoelliptic linear partial differential operator, +ν∶Ω → (0,∞) continuous and U a set of uniqueness for (idKΩ,CνP (∂)). If f∶U → E +is a function such that e′ ○ f admits an extension fe′ ∈ CνP (∂)(Ω) for every e′ ∈ G, +then there exists a unique extension F ∈ CνP (∂)(Ω,E) of f. +Proof. We choose F(Ω) ∶= (C∞ +P (∂)(Ω),τc) and F(Ω,E) ∶= (C∞ +P (∂)(Ω,E),τc). +Then we have Fν(Ω) = CνP (∂)(Ω) and Fν(Ω,E) = CνP (∂)(Ω,E) with the generator +(T E,T K) ∶= (idEΩ,idKΩ) for (Fν,E). +We note that F(Ω) and F(Ω,E) are ε- +compatible and (T E,T K) is a strong, consistent family for (F,E) by Proposition +4.2.17. We observe that Fν(Ω) is a Banach space by Proposition 4.2.14 and for +every compact K ⊂ Ω we have +sup +x∈K +∣f(x)∣ ≤ +(22) sup +z∈K +ν(z)−1∣f∣ν ≤ sup +z∈K +ν(z)−1, +f ∈ BFν(Ω), +yielding that BFν(Ω) is bounded in F(Ω). The space F(Ω) = (C∞ +P (∂)(Ω),τc) is a +Fréchet–Schwartz space, thus a Montel space, and it is easy to check that BFν(Ω) is +τc-closed. Hence the bounded and τc-closed set BFν(Ω) is compact in F(Ω). Finally, +we remark that the ε-compatibility of F(Ω) and F(Ω,E) in combination with the +consistency of (idEΩ,idKΩ) for (F,E) gives Fεν(Ω,E) = Fν(Ω,E) as linear spaces +by Proposition 5.2.25 c). From Theorem 5.2.29 follows our statement. +□ +If Ω = D ⊂ C is the open unit disc, P(∂) = ∂ the Cauchy–Riemann operator +and ν = 1 on D, then CνP (∂)(Ω,E) = H∞(D,E) and a sequence U ∶= (zn)n∈N ⊂ D of +distinct elements is a set of uniqueness for (idCD,H∞) if and only if it satisfies the +Blaschke condition ∑n∈N(1 − ∣zn∣) = ∞ (see e.g. [149, 15.23 Theorem, p. 303]). +For a continuous function ν∶D → (0,∞) and a complex lcHs E we define the +Bloch type spaces +Bν(D,E) ∶= {f ∈ O(D,E) ∣ ∀ α ∈ A ∶ ∣f∣ν,α < ∞} +with +∣f∣ν,α ∶= max(pα(f(0)),sup +z∈D +pα((∂1 +C)Ef(z))ν(z)). +If E = C, we write f ′(z) ∶= (∂1 +C)Cf(z) for z ∈ D and f ∈ O(D). +5.2.31. Proposition. If ν∶D → (0,∞) is continuous, then Bν(D) is a Banach +space. + +88 +5. APPLICATIONS +Proof. Let f ∈ Bν(D). From the estimates +∣f(z)∣ ≤ ∣f(0)∣ + ∣ +z +∫ +0 +f ′(ζ)dζ∣ ≤ ∣f(0)∣ + +∣z∣ +minξ∈[0,z] ν(ξ) sup +ζ∈[0,z] +∣f ′(ζ)∣ν(ζ) +≤ 2max(1, +∣z∣ +minξ∈[0,z] ν(ξ))∣f∣ν +for every z ∈ D and +max +∣z∣≤r ∣f(z)∣ ≤ 2max(1, +r +min∣z∣≤r ν(z))∣f∣ν +(51) +for all 0 < r < 1 and f ∈ Bν(D) it follows that Bν(D) is a Banach space by using +the completeness of (O(D),τc) analogously to the proof of Proposition 4.2.14. +□ +5.2.32. Proposition. Let Ω ⊂ C be open and E a locally complete lcHs over +C. Then ((∂n +C)E,(∂n +C)C)n∈N0 is a strong, consistent family for ((O(Ω),τc),E). +Proof. We recall from (5) that the real and complex derivatives are related +by +(∂β)Ef(z) = iβ2(∂∣β∣ +C )Ef(z), +z ∈ Ω, +(52) +for every f ∈ O(Ω,E) and β = (β1,β2) ��� N2 +0. Further, the Fréchet space (O(Ω),τc) +is barrelled. Due to Proposition 3.1.11 c) and (52) we have for all u ∈ (O(Ω),τc)εE +(∂n +C)ES(u)(z) = u(δz ○ (∂n +C)C), +n ∈ N0, z ∈ Ω, +which means that ((∂n +C)E,(∂n +C)C)n∈N0 is consistent. +Moreover, we have +(∂n +C)C(e′ ○ f)(z) = e′((∂n +C)Ef(z)), +n ∈ N0, z ∈ Ω, +for all e′ ∈ E′ and f ∈ O(Ω,E), implying the strength of ((∂n +C)E,(∂n +C)C)n∈N0. +□ +Let E be an lcHs and ν∶D → (0,∞) be continuous. We set ω ∶= {0}∪{(1,z) ∣ z ∈ +D}, define the operator T E∶O(D,E) → Eω by +T E(f)(0) ∶= f(0) +and +T E(f)(1,z) ∶= (∂1 +C)Ef(z), z ∈ D, +and the weight ν∗∶ω → (0,∞) by +ν∗(0) ∶= 1 +and +ν∗(1,z) ∶= ν(z), z ∈ D. +Then we have for every α ∈ A that +∣f∣ν,α = sup +x∈ω pα(T E(f)(x))ν∗(x), +f ∈ Bν(D,E), +and with F(D,E) ∶= O(D,E) we observe that Fν∗(D,E) = Bν(D,E) with generator +(T E,T C). +5.2.33. Corollary. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness, ν∶D → (0,∞) continuous and U∗ ⊂ D have an accumulation point in D. If +f∶{0} ∪ ({1} × U∗) → E is a function such that there is fe′ ∈ Bν(D) for each e′ ∈ G +with fe′(0) = e′(f(0)) and f ′ +e′(z) = e′(f(1,z)) for all z ∈ U∗, then there exists a +unique F ∈ Bν(D,E) with F(0) = f(0) and (∂1 +C)EF(z) = f(1,z) for all z ∈ U∗. +Proof. We take F(D) ∶= (O(D),τc) and F(D,E) ∶= (O(D,E),τc). Then we +have Fν∗(D) = Bν(D) and Fν∗(Ω,E) = Bν(D,E) with the weight ν∗ and generator +(T E,T C) for (Fν∗,E) described above. +The spaces F(D) and F(D,E) are ε- +compatible by Proposition 4.2.17 in combination with (23), and the generator is +a strong, consistent family for (F,E) by Proposition 5.2.32. Due to Proposition +5.2.31 Fν∗(D) = Bν(D) is a Banach space and we deduce from (51) that BFν∗(D) is +compact in the Montel space (O(D),τc). We note that the ε-compatibility of F(Ω) + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +89 +and F(Ω,E) in combination with the consistency of (T E,T C) for (F,E) gives +Fεν∗(D,E) = Fν∗(D,E) as linear spaces by Proposition 5.2.25 c). In addition, +U ∶= {0} ∪ {(1,z) ∣ z ∈ U∗} is a set of uniqueness for (T C,Fν∗) by the identity +theorem, proving our statement by Theorem 5.2.29. +□ +E a Fréchet space. In this section we restrict to the case that E is a Fréchet +space and G ⊂ E′ is generated by a sequence that fixes the topology in E. +5.2.34. Definition ([30, Definition 12, p. 8]). Let Y be a Fréchet space. An +increasing sequence (Bn)n∈N of bounded subsets of Y ′ +b fixes the topology in Y if +(B○ +n)n∈N is a fundamental system of zero neighbourhoods of Y . +5.2.35. Remark. Let Y be a Banach space. If B ⊂ Y ′ +b is bounded, i.e. bounded +w.r.t. the operator norm, such that B fixes the topology in Y , i.e. B○ is bounded +in Y , then B is called an almost norming subset. Examples of almost norming +subspaces are given in [7, Remark 1.2, p. 780–781]. For instance, the set of point +evaluations B ∶= {δ1/n ∣ n ∈ N} is almost norming for the Y ∶= H∞(D) ∶= C∞ +∂,b(D). +5.2.36. Definition (sb-restriction space). Let E be a Fréchet space, (Bn) fix +the topology in E and G ∶= span(⋃n∈N Bn). Let FV(Ω) be a dom-space, U a set of +uniqueness for (T K +m,FV)m∈M and set +FVG(U,E)sb ∶= {f ∈ FVG(U,E) ∣ ∀ n ∈ N ∶ {fe′ ∣ e′ ∈ Bn} is bounded in FV(Ω)}. +Let E be a Fréchet space, (Bn) fix the topology in E, G ∶= span(⋃n∈N Bn), +(T E +m,T K +m)m∈M be a strong, consistent generator for (FV,E) and U a set of unique- +ness for (T K +m,FV)m∈M. For u ∈ FV(Ω)εE we have RU,G(f) ∈ FVG(U,E) with +f ∶= S(u) by Remark 5.2.6 and for j ∈ J and m ∈ M +sup +e′∈Bn +∣fe′∣j,m = sup +e′∈Bn +sup +x∈ωm +∣e′(T E +m(f)(x)νj,m(x))∣ = sup +e′∈Bn +sup +y∈Nj,m(f) +∣e′(y)∣ +with Nj,m(f) ∶= {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm}. This set is bounded in E since +sup +y∈Nj,m(f) +pα(f) = ∣f∣j,m,α < ∞ +for all α ∈ A, implying supe′∈Bn ∣fe′∣j,m < ∞ and RU,G(f) ∈ FVG(U,E)sb. Hence the +injective linear map +RU,G∶S(FV(Ω)εE) → FVG(U,E)sb, f ↦ (T E +m(f)(x))(m,x)∈U, +is well-defined. +5.2.37. Question. Let E be a Fréchet space, (Bn) fix the topology in E +and G ∶= span(⋃n∈N Bn). Let (T E +m,T K +m)m∈M be a strong, consistent generator for +(FV,E) and U a set of uniqueness for (T K +m,FV)m∈M. When is the injective restric- +tion map +RU,G∶S(FV(Ω)εE) → FVG(U,E)sb, f ↦ (T E +m(f)(x))(m,x)∈U, +surjective? +5.2.38. Remark. Let E be a Fréchet space with increasing system of seminorms +(pαn)n∈N, Bn ∶= B○ +αn where Bαn ∶= {x ∈ E ∣ pαn(x) < 1}, (T E +m,T K +m)m∈M a strong, +consistent generator for (FV,E) and U a set of uniqueness for (T K +m,FV)m∈M. If +FV(Ω) is a BC-space, then FVE′(U,E)sb = FVE′(U,E) by Proposition 5.2.14. +Hence Theorem 5.2.15 (i) answers Question 5.2.37 in this case. +Let us turn to the case where G need not coincide with E′. + +90 +5. APPLICATIONS +FV(Ω) a Fréchet–Schwartz space and E a Fréchet space. We recall the +following result. +5.2.39. Proposition ([69, Lemma 9, p. 504]). Let E be a Fréchet space, (Bn) +fix the topology in E, Y a Fréchet–Schwartz space and X ⊂ Y ′ +b (= Y ′ +κ) a dense +subspace. +Set G ∶= span(⋃n∈N Bn) and let A∶X → E be a linear map which is +σ(X,Y )-σ(E,G)-continuous and satisfies that At(Bn) is bounded in Y for each +n ∈ N. Then A has a (unique) extension ̂A ∈ Y εE. +Next, we improve [69, Theorem 1 ii), p. 501]. +5.2.40. Theorem. Let E be a Fréchet space, (Bn) fix the topology in E and G ∶= +span(⋃n∈N Bn), (T E +m,T K +m)m∈M a strong, consistent generator for (FV,E), FV(Ω) +a Fréchet–Schwartz space and U a set of uniqueness for (T K +m,FV)m∈M. Then the +restriction map RU,G∶S(FV(Ω)εE) → FVG(U,E)sb is surjective. +Proof. Let f ∈ FVG(U,E)sb. +We set X ∶= span{T K +m,x ∣ (m,x) ∈ U} and +Y ∶= FV(Ω). Let A∶X → E be the linear map determined by A(T K +m,x) ∶= f(m,x) +which is well-defined since G is σ(E′,E)-dense. From +e′(A(T K +m,x)) = (e′ ○ f)(m,x) = T K +m,x(fe′) +for every e′ ∈ G and (m,x) ∈ U it follows that A is σ(X,Y )-σ(E,G)-continuous and +sup +e′∈Bn +∣At(e′)∣j,k = sup +e′∈Bn +∣fe′∣j,k < ∞ +for all j ∈ J, k ∈ M and n ∈ N. Due to Proposition 5.2.39 there is an extension +̂A ∈ FV(Ω)εE of A. We set F ∶= S(̂A) and get for all (m,x) ∈ U that +T E +m(F)(x) = T E +mS(̂A)(x) = ̂A(T K +m,x) = f(m,x) +by consistency, which means RU,G(F) = f. +□ +5.2.41. Corollary. Let E be a Fréchet space, (Bn) fix the topology in E and +G ∶= span(⋃n∈N Bn). +Let V ∶= (νj)j∈N be an increasing family of weights which +is locally bounded away from zero on an open set Ω ⊂ Rd, P(∂)K a hypoelliptic +linear partial differential operator, CVP (∂)(Ω) a Schwartz space and U ⊂ Ω a set of +uniqueness for (idKΩ,CVP (∂)). If f∶U → E is a function such that e′ ○ f admits +an extension fe′ ∈ CVP (∂)(Ω) for each e′ ∈ G and {fe′ ∣ e′ ∈ Bn} is bounded in +CVP (∂)(Ω) for each n ∈ N, then there is a unique extension F ∈ CVP (∂)(Ω,E) of f. +Proof. CVP (∂)(Ω) is a Fréchet–Schwartz space and (idEΩ,idKΩ) a strong, +consistent generator for (CVP (∂),E) by Proposition 4.2.14 and the proof of Example +4.2.16 b). Now, Theorem 5.2.40 and Proposition 5.2.8 prove our statement. +□ +We already mentioned examples of families of weights V such that CVP (∂)(Rd) +is a nuclear Fréchet space and sets of uniqueness for (idKRd ,CVP (∂)) in Remark +4.2.20 and Remark 5.2.23 and if P(∂) = ∂ or P(∂) = ∆. Further sets of uniqueness +are given in Remark 5.2.66. If E is a Banach space, then an almost norming set +fixes the topology and examples can be found via Remark 5.2.35. +Fν(Ω) a Banach space and E a Fréchet space. Let E be a Fréchet space, +(Bn) fix the topology in E and recall the assumptions of Remark 5.2.26. +Let +(T E,T K) be a strong, consistent family for (F,E) and a generator for (Fν,E). +Let F(Ω) and F(Ω,E) be ε-into-compatible and the inclusion Fν(Ω) ↪ F(Ω) +continuous. Consider a set of uniqueness U for (T K,Fν) and G ∶= span(⋃n∈N Bn) ⊂ + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +91 +E′. For u ∈ F(Ω)εE such that u(B○F (Ω)′ +Fν(Ω) ) is bounded in E we have RU,G(f) ∈ +FνG(U,E) with f ∶= S(u) ∈ Fεν(Ω,E) by (50). We note that +sup +e′∈Bn +∣fe′∣Fν(Ω) = sup +e′∈Bn +sup +x∈ω ∣e′(T E(f)(x)ν(x))∣ = sup +e′∈Bn +sup +y∈Nω(f) +∣e′(y)∣ +with the bounded set Nω(f) ∶= {T E(f)(x)ν(x) ∣ x ∈ ω} ⊂ E, implying RU,G(f) ∈ +FVG(U,E)sb. Thus the injective linear map +RU,G∶Fεν(Ω,E) → FνG(U,E)sb, f ↦ (T E(f)(x))x∈U, +is well-defined. +5.2.42. Question. Let the assumptions of Remark 5.2.26 be fulfilled, E be a +Fréchet space, (Bn) fix the topology in E and G ∶= span(⋃n∈N Bn). When is the +injective restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E)sb, f ↦ (T E(f)(x))x∈U, +surjective? +Now, we can generalise [70, Corollary 2.4, p. 692] and [93, Theorem 11, p. 5]. +5.2.43. Corollary. Let E be a Fréchet space, (Bn) fix the topology in E, set +G ∶= span(⋃n∈N Bn) and let F(Ω) and F(Ω,E) be ε-into-compatible. Let (T E,T K) +be a generator for (Fν,E) and a strong, consistent family for (F,E), Fν(Ω) a +Banach space whose closed unit ball BFν(Ω) is a compact subset of F(Ω) and U a +set of uniqueness for (T K,Fν). Then the restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E)sb +is surjective. +Proof. Let f ∈ FνG(U,E)sb. +Then {fe′ ∣ e′ ∈ Bn} is bounded in Fν(Ω) +for each n ∈ N. We deduce for each n ∈ N, (ak)k∈N ∈ ℓ1 and (e′ +k)k∈N ⊂ Bn that +(∑k∈N ake′ +k)○f admits the extension ∑k∈N akfe′ +k in Fν(Ω). Due to [69, Proposition +7, p. 503] the LB-space E′((Bn)n∈N) ∶= lim +←� +n∈N +E′(Bn), where +E′(Bn) ∶= {��� +k∈N +ake′ +k ∣ (ak)k∈N ∈ ℓ1, (e′ +k)k∈N ⊂ Bn} +is endowed with its Banach space topology for n ∈ N, determines boundedness in +E. +Hence we conclude that f ∈ FνE′((Bn)n∈N)(U,E), which yields that there is +u ∈ F(Ω)εE with bounded u(B○F (Ω)′ +Fν(Ω) ) ⊂ E such that RU,G(S(u)) = f by Theorem +5.2.29. +□ +As an application we directly obtain the following two corollaries of Corol- +lary 5.2.43 since its assumptions are fulfilled by the proof of Corollary 5.2.30 and +Corollary 5.2.33, respectively. +5.2.44. Corollary. Let E be a Fréchet space, (Bn) fix the topology in E and +G ∶= span(⋃n∈N Bn), Ω ⊂ Rd open, P(∂)K a hypoelliptic linear partial differential +operator, ν∶Ω → (0,∞) continuous and U a set of uniqueness for (idKΩ,CνP (∂)). If +f∶U → E is a function such that e′ ○f admits an extension fe′ ∈ CνP (∂)(Ω) for each +e′ ∈ G and {fe′ ∣ e′ ∈ Bn} is bounded in CνP (∂)(Ω) for each n ∈ N, then there exists +a unique extension F ∈ CνP (∂)(Ω,E) of f. +5.2.45. Corollary. Let E be a Fréchet space, (Bn) fix the topology in E and +G ∶= span(⋃n∈N Bn), ν∶D → (0,∞) continuous and U∗ ⊂ D have an accumulation +point in D. If f∶{0} ∪ ({1} × U∗) → E is a function such that there is fe′ ∈ Bν(D) +for each e′ ∈ G with fe′(0) = e′(f(0)) and f ′ +e′(z) = e′(f(1,z)) for all z ∈ U∗ and + +92 +5. APPLICATIONS +{fe′ ∣ e′ ∈ Bn} is bounded in Bν(D) for each n ∈ N, then there exists a unique +F ∈ Bν(D,E) with F(0) = f(0) and (∂1 +C)EF(z) = f(1,z) for all z ∈ U∗. +5.2.2. Extension from thick sets. In order to obtain an affirmative answer +to Question 5.2.9 for general separating subspaces of E′ we have to restrict to the +spaces FV(Ω) from Definition 3.1.4 and a certain class of sets of uniqueness. +5.2.46. Definition (fix the topology). Let FV(Ω) be a dom-space. We say +that U ⊂ ⋃m∈M({m}×ωm) fixes the topology in FV(Ω) if for every j ∈ J and m ∈ M +there are i ∈ J, k ∈ M and C > 0 such that +∣f∣j,m ≤ C +sup +x∈ωk +(k,x)∈U +∣T K +k (f)(x)∣νi,k(x), +f ∈ FV(Ω). +In particular, U is a set of uniqueness if it fixes the topology. The present +definition of fixing the topology is a generalisation of [30, Definition 13, p. 234]. +Sets that fix the topology appear under several different notions. Rubel and Shields +call them dominating in [148, 4.10 Definition, p. 254] in the context of bounded +holomorphic functions. In the context of the space of holomorphic functions with +the topology of compact convergence studied by Grosse-Erdmann [81, p. 401] they +are said to determine locally uniform convergence. Ehrenpreis [61, p. 3,4,13] (cf. +[156, Definition 3.2, p. 166]) refers to them as sufficient sets when he considers +inductive limits of weighted spaces of entire resp. holomorphic functions, including +the case of Banach spaces. In the case of Banach spaces sufficient sets coincide +with weakly sufficient sets defined by Schneider [156, Definition 2.1, p. 163] (see e.g. +[102, §7, 1), p. 547]) and these notions are extended beyond spaces of holomorphic +functions by Korobe˘ınik [102, p. 531]. Seip [162, p. 93] uses the term sampling sets +in the context of weighted Banach spaces of holomorphic functions whereas Beurling +uses the term balayage in [14, p. 341] and [14, Definition, p. 343]. Leibowitz [122, +Exercise 4.1.4, p. 53], Stout [170, 7.1 Definition, p. 36] and Globevnik [76, p. 291– +292] call them boundaries in the context of subalgebras of the algebra C(Ω,C) of +complex-valued continuous functions on a compact Hausdorff space Ω with sup- +norm. Fixing the topology is also connected to the notion of frames used by Bonet +et al. in [31]. Let us set +ℓV(U,E) ∶= {f∶U → E ∣ ∀ j ∈ J,m ∈ M,α ∈ A ∶ ∥f∥j,m,α < ∞} +(53) +with +∥f∥j,m,α ∶= +sup +x∈ωm +(m,x)∈U +pα(f(m,x))νj,m(x) +for an lcHs E and a set U which fixes the topology in FV(Ω). If M is a singleton, +ωm = Ω = U, then ℓV(U,E) coincides with the space defined right above Example +4.2.2. If U is countable, then the inclusion ℓV(U) ↪ KU continuous where KU is +equipped with the topology of pointwise convergence and ℓV(U) contains the space +of sequences (on U) with compact support as a linear subspace, then (T K +k,x)(k,x)∈U +is an ℓV(U)-frame in the sense of [31, Definition 2.1, p. 3]. +5.2.47. Definition (lb-restriction space). Let FV(Ω) be a dom-space, U fix +the topology in FV(Ω) and G ⊂ E′ a separating subspace. We set +NU,i,k(f) ∶= {f(k,x)νi,k(x) ∣ x ∈ ωk, (k,x) ∈ U} +for i ∈ J, k ∈ M and f ∈ FVG(U,E) and +FVG(U,E)lb ∶={f ∈ FVG(U,E) ∣ ∀ i ∈ J, k ∈ M ∶ NU,i,k(f) bounded in E} +=FVG(U,E) ∩ ℓV(U,E). + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +93 +Consider a set U which fixes the topology in FV(Ω), a separating subspace +G ⊂ E′ and a strong, consistent family (T E +m,T K +m)m∈M for (FV,E). For u ∈ FV(Ω)εE +set f ∶= S(u) ∈ FV(���,E) by Theorem 3.1.12. Then we have RU,G(f) ∈ FVG(U,E) +with f ∶= S(u) by Remark 5.2.6 and for i ∈ J and k ∈ M +sup +y∈NU,i,k(RU,G(f)) +pα(y) = +sup +x∈ωk +(k,x)∈U +pα(T E +k (f)(x))νi,k(x) ≤ ∣f∣i,k,α < ∞ +for all α ∈ A, implying the boundedness of NU,i,k(RU,G(f)) in E. Thus RU,G(f) ∈ +FVG(U,E)lb and the injective linear map +RU,G∶S(FV(Ω)εE) → FVG(U,E)lb, f ↦ (T E +m(f)(x))(m,x)∈U, +is well-defined. +5.2.48. Question. Let G ⊂ E′ be a separating subspace, (T E +m,T K +m)m∈M a +strong, consistent generator for (FV,E) and U fix the topology in FV(Ω). When +is the injective restriction map +RU,G∶S(FV(Ω)εE) → FVG(U,E)lb, f ↦ (T E +m(f)(x))(m,x)∈U, +surjective? +If G ⊂ E′ determines boundedness and U fixes the topology in FV(Ω), then +the preceding question and Question 5.2.9 coincide. +5.2.49. Remark. Let G ⊂ E′ determine boundedness, (T E +m,T K +m)m∈M a strong, +consistent generator for (FV,E) and U fix the topology in FV(Ω). Then +FVG(U,E)lb = FVG(U,E). +Proof. We only need to show that the inclusion ‘⊃’ holds. Let f ∈ FVG(U,E). +Then there is fe′ ∈ FV(Ω) with T K +m(fe′)(x) = (e′ ○ f)(m,x) for all (m,x) ∈ U and +sup +y∈NU,i,k(f) +∣e′(y)∣ = +sup +x∈ωk +(k,x)∈U +∣(e′ ○ f)(k,x)∣νi,k(x) ≤ ∣fe′∣i,k < ∞ +for each e′ ∈ G, i ∈ J and k ∈ M. Since G ⊂ E′ determines boundedness, this means +that NU,i,k(f) is bounded in E and hence f ∈ FVG(U,E)lb. +□ +FV(Ω) arbitrary and E a semi-Montel space. +5.2.50. Definition (generalised Schwartz space). We call an lcHs E a gener- +alised Schwartz space if every bounded set in E is already precompact. +In particular, semi-Montel spaces and Schwartz spaces are generalised Schwartz +spaces by [89, 10.4.3 Corollary, p. 202]. Conversely, a generalised Schwartz space is +a Schwartz space if it is quasi-normable by [89, 10.7.3 Corollary, p. 215]. Moreover, +looking at the proof of Lemma 3.2.2 b), we see that this lemma not only holds for +semi-Montel or Schwartz spaces but for all generalised Schwartz spaces. +5.2.51. Proposition. Let E be an lcHs, FV(Ω) a dom-space and U fix the +topology in FV(Ω). Then Rf ∈ L(E′ +b,FV(Ω)) and Rf(B○ +α) is bounded in FV(Ω) +for every f ∈ FVE′(U,E)lb and α ∈ A where Bα ∶= {x ∈ E ∣ pα(x) < 1} and Rf is +the map from Remark 5.2.5. In addition, if E is a generalised Schwartz space, then +Rf ∈ L(E′ +γ,FV(Ω)) and Rf(B○ +α) is relatively compact in FV(Ω). +Proof. Let f ∈ FVE′(U,E)lb, j ∈ J and m ∈ M. Then there are i ∈ J, k ∈ M +and C > 0 such that for every e′ ∈ E′ +∣Rf(e′)∣j,m = ∣fe′∣j,m ≤ C +sup +x∈ωk +(k,x)∈U +∣T K +k (fe′)(x)∣νi,k(x) + +94 +5. APPLICATIONS += C +sup +x∈ωk +(k,x)∈U +∣(e′ ○ f)(k,x)∣νi,k(x) = C +sup +y∈NU,i,k(f) +∣e′(y)∣, +which proves the first part because NU,i,k(f) is bounded in E. Let us consider the +second part. The bounded set NU,i,k(f) is already precompact in E because E is +a generalised Schwartz space. Therefore we have Rf ∈ L(E′ +γ,FV(Ω)). The polar +B○ +α is relatively compact in E′ +γ for every α ∈ A by the Alaoğlu–Bourbaki theorem +and thus Rf(B○ +α) in FV(Ω) as well. +□ +5.2.52. Theorem. Let E be a semi-Montel space, (T E +m,T K +m)m∈M a strong, con- +sistent generator for (FV,E) and U fix the topology in FV(Ω). Then the restriction +map RU,E′∶S(FV(Ω)εE) → FVE′(U,E)lb is surjective. +Proof. Let f ∈ FVE′(Ω,E)lb and e′ ∈ E′. For every f ′ ∈ FV(Ω)′ there are +j ∈ J, m ∈ M and C0 > 0 with +∣Rt +f(f ′)(e′)∣ = ∣f ′(fe′)∣ ≤ C0∣fe′∣j,m. +By the proof of Proposition 5.2.51 there are i ∈ J, k ∈ M and C > 0 such that +∣Rt +f(f ′)(e′)∣ ≤ C0C +sup +y∈NU,i,k(f) +∣e′(y)∣ ≤ C0C +sup +y∈acx(NU,i,k(f)) +∣e′(y)∣. +The set acx(NU,i,k(f)) is absolutely convex and compact by [89, 6.2.1 Proposi- +tion, p. 103] and [89, 6.7.1 Proposition, p. 112] because E is a semi-Montel space. +Therefore Rt +f(f ′) ∈ (E′ +κ)′ = J (E) by the Mackey–Arens theorem. As in Theorem +5.2.15 we obtain J −1○Rt +f ∈ FV(Ω)εE by (45), (46) and Proposition 5.2.51. Setting +F ∶= S(J −1 ○ Rt +f), we conclude T E +m(F)(x) = f(m,x) for all (m,x) ∈ U by (47) and +so RU,E′(F) = f. +□ +5.2.53. Remark. Let E be a Fréchet space with increasing system of seminorms +(pαn)n∈N, Bn ∶= B○ +αn where Bαn ∶= {x ∈ E ∣ pαn(x) < 1}, (T E +m,T K +m)m∈M a strong, +consistent generator for (FV,E) and U a set of uniqueness for (T K +m,FV)m∈M. If U +fixes the topology of FV(Ω), then FVE′(U,E)sb = FVE′(U,E) by Remark 5.2.49 +and Proposition 5.2.51. Hence Theorem 5.2.52 answers Question 5.2.37 if E is a +Fréchet–Montel space. +Our first application of Theorem 5.2.52 concerns the space Cbu(Ω,E) of bounded +uniformly continuous functions from a metric space Ω to an lcHs E from Example +4.2.7. +5.2.54. Corollary. Let Ω be a metric space, U ⊂ Ω a dense subset and E a +semi-Montel space. If f∶U → E is a function such that e′ ○ f admits an extension +fe′ ∈ Cbu(Ω) for each e′ ∈ E′, then there is a unique extension F ∈ Cbu(Ω,E) of f. +In particular, +Cbu(Ω,E) = {f∶Ω → E ∣ ∀ e′ ∈ E′ ∶ e′ ○ f ∈ Cbu(Ω)}. +Proof. (idEΩ,idKΩ) is a strong, consistent generator for (Cbu,E) and we have +Cbu(Ω)εE ≅ Cbu(Ω,E) via S by Example 4.2.7. Due to Theorem 5.2.52, Proposition +5.2.8 and Remark 5.2.49 with G = E′ the extension F exists and is unique because +the dense set U ⊂ Ω fixes the topology in Cbu(Ω). The rest follows from Proposition +5.2.10. +□ +Next, we consider the space A(Ω,E) of continuous functions from Ω to an lcHs +E over C which are holomorphic on an open and bounded set Ω ⊂ C from Example +4.2.13. + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +95 +5.2.55. Corollary. Let Ω ⊂ C be open and bounded, U ⊂ Ω fix the topology in +A(Ω) and E a semi-Montel space over C. If f∶U → E is a function such that e′ ○f +admits an extension fe′ ∈ A(Ω) for each e′ ∈ E′, then there is a unique extension +F ∈ A(Ω,E) of f. In particular, +A(Ω,E) = {f∶Ω → E ∣ ∀ e′ ∈ E′ ∶ e′ ○ f ∈ A(Ω)}. +Proof. (idEΩ,idCΩ) is a strong, consistent generator for (A,E) and A(Ω)εE ≅ +A(Ω,E) via S by Example 4.2.13. Due to Theorem 5.2.52, Proposition 5.2.8 and +Remark 5.2.49 with G = E′ the extension F exists and is unique. The remaining +part follows from Proposition 5.2.10. +□ +If Ω ⊂ C is connected, then the boundary ∂Ω of Ω fixes the topology in A(Ω) +by the maximum principle. If Ω = D, then ∂D is the intersection of all sets that fix +the topology in A(D) by [170, 7.7 Example, p. 39]. +If E is a generalised Schwartz space which is not a semi-Montel space, we do not +know whether the extension results in Corollary 5.2.54 and Corollary 5.2.55 hold +but we still have a weak-strong principle due to the following observation which is +based on [87, Chap. 3, §9, Proposition 2, p. 231] with σ(E,E′) replaced by σ(E,G). +5.2.56. Proposition. If +(i) E is a semi-Montel space and G ⊂ E′ a separating subspace, or +(ii) E is a generalised Schwartz space and G ⊂ ̂E′ a separating subspace, i.e. +separates the points of the completion ̂E, +then the initial topology of E and the topology σ(E,G) coincide on the bounded sets +of E. +Proof. (i) Let B ⊂ E be a bounded set. If E is a semi-Montel space, then the +closure B is compact in E. The topology induced by σ(E,G) on B is Hausdorff and +weaker than the initial topology induced by E. Thus the two topologies coincide +on B and so on B by the remarks above [87, Chap. 3, §9, Proposition 2, p. 231]. +(ii) Let B ⊂ E be a bounded set. If E is a generalised Schwartz space, then B is +precompact in E and relatively compact in the completion ̂E by [89, 3.5.1 Theorem, +p. 64]. Hence the closure B is compact in ̂E. The topology induced by σ( ̂E,G) on +B is Hausdorff and weaker than the initial topology induced by ̂E, implying that +the two topologies coincide on B as in part (i). This yields that σ(E,G) and the +initial topology of E coincide on B because σ(E,G) = σ( ̂E,G) on B and the initial +topologies of E and ̂E coincide on B as well. +□ +Concerning (ii), we note that a separating subspace G ⊂ E′ of E need not +separate the points of ̂E by [79, 5.4 Example, p. 36] (even though E′ = ̂E′ by [89, +3.4.2 Theorem, p. 61–62]). Next, we apply Proposition 5.2.56 to the space A(Ω,E). +5.2.57. Remark. Let E be an lcHs over C and Ω ⊂ C open and bounded. If +(i) E is a semi-Montel space and G ⊂ E′ determines boundedness, or +(ii) E is a generalised Schwartz space and G ⊂ ̂E′ a separating subspace which +determines boundedness in E, +then +A(Ω,E) = {f∶Ω → E ∣ ∀ e′ ∈ G ∶ e′ ○ f ∈ A(Ω)}. +Indeed, let us denote the right-hand side by A(Ω,E)σ and set Eσ ∶= (E,σ(E,G)). +Then A(Ω,E)σ = A(Ω,Eσ) and f(Ω) is bounded for every f ∈ A(Ω,E)σ as G +determines boundedness in E. The initial topology of E and σ(E,G) coincide on +the bounded range f(Ω) of f ∈ A(Ω,E)σ by Proposition 5.2.56. Hence we deduce +that +A(Ω,E)σ = A(Ω,Eσ) = A(Ω,E). + +96 +5. APPLICATIONS +In this way Bierstedt proves his weak-strong principles for weighted continuous +functions in [17, 2.10 Lemma, p. 140] with G = E′ = ̂E′. +FV(Ω) a Fréchet–Schwartz space and E locally complete. +5.2.58. Definition (chain-structured). Let FV(Ω) be a dom-space. We say +that U ⊂ ⋃m∈N({m} × ωm) is chain-structured if +(i) (k,x) ∈ U +⇒ ∀ m ∈ N, m ≥ k ∶ (m,x) ∈ U, +(ii) ∀ (k,x) ∈ U, m ∈ N, m ≥ k, f ∈ FV(Ω) ∶ T K +k (f)(x) = T K +m(f)(x). +5.2.59. Remark. Let Ω ⊂ Rd be open and V∞ a directed family of weights. +Concerning the operators (T K +m)m∈N0 of CV∞(Ω) from Example 3.1.9 a) where ωm ∶= +{β ∈ Nd +0 ∣ ∣β∣ ≤ m} × Ω resp. ωm ∶= Nd +0 × Ω, we have for all k ∈ N0 and f ∈ CV∞(Ω) +that +T K +k (f)(β,x) = ∂βf(x) = T K +m(f)(β,x), +β ∈ Nd +0, ∣β∣ ≤ k, x ∈ Ω, +for all m ∈ N0, m ≥ k. Hence condition (ii) of Definition 5.2.58 is fulfilled for any +U ⊂ ⋃m∈N0({m} × ωm) in this case. Condition (i) says that once a ‘link’ (k,β,x) +belongs to U for some order k, then the ‘link’ (m,β,x) belongs to U for any higher +order m as well. +5.2.60. Definition (diagonally dominated, increasing). We say that a family +V ∶= (νj,m)j,m∈N of weights on Ω is diagonally dominated and increasing if ωm ⊂ +ωm+1 for all m ∈ N and νj,m ≤ νmax(j,m),max(j,m) on ωmin(j,m) for all j,m ∈ N as well +as νj,j ≤ νj+1,j+1 on ωj for all j ∈ N. +5.2.61. Remark. Let FV(Ω) be a dom-space, U ⊂ ⋃m∈N({m} × ωm) chain- +structured, G ⊂ E′ a separating subspace and V diagonally dominated and increas- +ing. +a) If U fixes the topology in FV(Ω), then +FVG(U,E)lb = {f ∈ FVG(U,E) ∣ ∀ i ∈ N ∶ NU,i(f) bounded in E} +with NU,i(f) ∶= NU,i,i(f). +b) Let FV(Ω) be a Fréchet space. We set Um ∶= {(m,x) ∈ U ∣ x ∈ ωm} and +Bj ∶= ⋃j +m=1{T K +m,x(⋅)νm,m(x) ∣ (m,x) ∈ Um} ⊂ FV(Ω)′ for j ∈ N. Then +U fixes the topology in FV(Ω) in the sense of Definition 5.2.46 if and +only if the sequence (Bj)j∈N fixes the topology in FV(Ω) in the sense of +Definition 5.2.34. +Proof. Let us begin with a). We only need to show that the inclusion ‘⊃’ holds. +Let f be an element of the right-hand side and i,k ∈ N. We set m ∶= max(i,k) and +observe that for (k,x) ∈ U we have (m,x) ∈ U by (i) and +(e′ ○ f)(k,x) = T K +k (fe′)(x) = +(ii) T K +m(fe′)(x) = (e′ ○ f)(m,x) +for each e′ ∈ G with (i) and (ii) from the definition of U being chain-structured. +Since G is separating, it follows that f(k,x) = f(m,x). Hence we get for all α ∈ A +sup +y∈NU,i,k(f) +pα(y) = +sup +x∈ωk +(k,x)∈U +pα(f(k,x))νi,k(x) ≤ +(i) +sup +x∈ωm +(m,x)∈U +pα(f(k,x))νm,m(x) += +sup +x∈ωm +(m,x)∈U +pα(f(m,x))νm,m(x) < ∞ +using that ωk ⊂ ωm and V is diagonally dominated. +Let us turn to part b). ‘⇒’: Let j ∈ N and A ⊂ FV(Ω) be bounded. Then +sup +y∈Bj +sup +f∈A +∣y(f)∣ = +sup +1≤m≤j +(m,x)∈Um +sup +f∈A +∣T K +m(f)(x)∣νm,m(x) ≤ sup +f∈A +sup +1≤m≤j +∣f∣m,m < ∞ + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +97 +since A is bounded, implying that Bj is bounded in FV(Ω)′ +b. Further, (Bj) is +increasing by definition. Additionally, for all j ∈ N +B○ +j = +j +⋂ +m=1 +{f ∈ FV(Ω) ∣ +sup +x∈ωm +(m,x)∈U +∣T K +m(f)(x)∣νm,m(x) ≤ 1} += {f ∈ FV(Ω) ∣ +sup +x∈ωj +(j,x)∈U +∣T K +j (f)(x)∣νj,j(x) ≤ 1} +because U is chain-structured and V increasing. Thus (B○ +j) is a fundamental system +of zero neighbourhoods of FV(Ω) if U fixes the topology. +‘⇐’: Let j,m ∈ N. Then there are i ∈ N and ε > 0 such that +εB○ +i ⊂ {f ∈ FV(Ω) ∣ ∣f∣j,m ≤ 1} =∶ Dj,m +which follows from fixing the topology in the sense of Definition 5.2.34. Let f ∈ Dj,m +and set +∣f∣Ui ∶= +sup +(i,x)∈Ui +∣T K +i (f)(x)∣νi,i(x). +If ∣f∣Ui = 0, then tf ∈ εB○ +i for all t > 0 and hence t∣f∣j,m = ∣tf∣j,m ≤ 1 for all t > 0, +which yields ∣f∣j,m = 0 = ∣f∣Ui. If ∣f∣Ui ≠ 0, then +f +∣f∣Ui ∈ B○ +i and thus ε +f +∣f∣Ui ∈ Dj,m, +implying +∣f∣j,m = 1 +ε∣f∣Ui∣ε f +∣f∣Ui +∣j,m ≤ 1 +ε∣f∣Ui. +The inequality ∣f∣j,m ≤ 1 +ε∣f∣Ui still holds if ∣f∣Ui = 0. +□ +5.2.62. Theorem ([30, Theorem 16, p. 236]). Let Y be a Fréchet–Schwartz +space, (Bj)j∈N fix the topology in Y and A∶X ∶= span(⋃j∈N Bj) → E be a linear map +which is bounded on each Bj. If +a) (At)−1(Y ) is dense in E′ +b and E locally complete, or +b) (At)−1(Y ) is dense in E′ +σ and E is Br-complete, +then A has a (unique) extension ̂A ∈ Y εE. +Now, we generalise [30, Theorem 17, p. 237]. +5.2.63. Theorem. Let E be an lcHs and G ⊂ E′ a separating subspace. Let +(T E +m,T K +m)m∈M be a strong, consistent generator for (FV,E), FV(Ω) a Fréchet– +Schwartz space, V diagonally dominated and increasing and U be chain-structured +and fix the topology in FV(Ω). If +a) G is dense in E′ +b and E locally complete, or +b) E is Br-complete, +then the restriction map RU,G∶S(FV(Ω)εE) → FVG(U,E)lb is surjective. +Proof. Let f ∈ FVG(U,E)lb. We set X ∶= span(⋃j∈N Bj) with Bj from Re- +mark 5.2.61 b) and Y ∶= FV(Ω). Let A∶X → E be the linear map determined +by +A(T K +m,x(⋅)νm,m(x)) ∶= f(m,x)νm,m(x) +for 1 ≤ m ≤ j and (m,x) ∈ Um with Um from Remark 5.2.61 b). The map A is +well-defined since G is σ(E′,E)-dense, and bounded on each Bj because A(Bj) = +⋃j +m=1 NU,m(f). +Let e′ ∈ G and fe′ be the unique element in FV(Ω) such that +T K +m(fe′)(x) = (e′ ○ f)(m,x) for all (m,x) ∈ U, which implies T K +m(fe′)(x)νm,m(x) = +(e′○A)(T K +m,x(⋅)νm,m(x)) for all (m,x) ∈ Um. This equation allows us to consider fe′ +as a linear form on X (by fe′(T K +m,x(⋅)��m,m(x)) ∶= (e′ ○ A)(T K +m,x(⋅)νm,m(x))), which +yields e′ ○ A ∈ FV(Ω) for all e′ ∈ G. It follows that G ⊂ (At)−1(Y ). Noting that G +is σ(E′,E)-dense, we apply Theorem 5.2.62 and obtain an extension ̂A ∈ FV(Ω)εE + +98 +5. APPLICATIONS +of A. We set F ∶= S(̂A) and observe that for all (m,x) ∈ U there is j ∈ N, j ≥ m, +such that (j,x) ∈ Uj and νj,j(x) > 0 by (6) and because U is chain-structured and +V diagonally dominated and increasing. Due to the proof of Remark 5.2.61 a) we +have f(j,x) = f(m,x) and thus +T E +m(F)(x) = T E +mS(̂A)(x) = ̂A(T K +m,x) = +1 +νj,j(x) +̂A(T K +m,x(⋅)νj,j(x)) += +1 +νj,j(x) +̂A(T K +j,x(⋅)νj,j(x)) = f(j,x) = f(m,x) +by consistency, yielding RU,G(F) = f. +□ +In particular, condition a) is fulfilled if E is semi-reflexive. Indeed, if E is semi- +reflexive, then E is quasi-complete by [153, Chap. IV, 5.5, Corollary 1, p. 144] and +G +b(E′,E) = G +τ(E′,E) = E′ by [89, 11.4.1 Proposition, p. 227] and the bipolar theorem. +For instance, condition b) is satisfied if E is a Fréchet space or E = (C∞ +∂,b(D),β) +which is a Br-complete space by Proposition 5.2.16 and is not a Fréchet space by +Remark 4.2.23. +As stated, our preceding theorem generalises [30, Theorem 17, p. 237] where +FV(Ω) is a closed subspace of CW∞(Ω) for open, connected Ω ⊂ Rd. A characteri- +sation of sets that fix the topology in the space CW∞ +∂ (Ω) of holomorphic functions +on an open, connected set Ω ⊂ C is given in [30, Remark 14, p. 235]. The characteri- +sation given in [30, Remark 14 (b), p. 235] is still valid and applied in [30, Corollary +18, p. 238] for closed subspaces of CW∞ +P (∂)(Ω) where P(∂)K is a hypoelliptic linear +partial differential operator which satisfies the maximum principle, namely, that +U ⊂ Ω fixes the topology if and only if there is a sequence (Ωn)n∈N of relatively +compact, open subsets of Ω with ⋃n∈N Ωn = Ω such that ∂Ωn ⊂ U ∩ Ωn+1 for all +n ∈ N. Among the hypoelliptic operators P(∂)K satisfying the maximum principle +are the Cauchy–Riemann operator ∂ and the Laplacian ∆. Further examples can +be found in [74, Corollary 3.2, p. 33]. The statement of [30, Corollary 18, p. 238] +for the space of holomorphic functions is itself a generalisation of [81, Theorem 2, +p. 401] with [81, Remark 2 (a), p. 406] where E is Br-complete and of [92, Theo- +rem 6, p. 10] where E is semi-reflexive. The case that G is dense in E′ +b and E is +sequentially complete is covered by [77, 3.3 Satz, p. 228–229], not only for spaces +of holomorphic functions, but for several classes of function spaces. +Let us turn to other families of weights than W∞. Due to Proposition 4.2.19 +we already know that U ∶= {0} × C fixes the topology in CV∞ +∂ (C) = CV∂(C) and +U ∶= {0} × Rd in CV∞ +∆(Rd) = CV∆(Rd) if V ∶= (νj)j∈N fulfils Condition 4.2.18 and +V∞ ∶= (νj,m)j∈N,m∈N0 where νj,m∶{β ∈ Nd +0 ∣ ∣β∣ ≤ m} × Rd → [0,∞), νj,m(β,x) ∶= +νj(x). +5.2.64. Corollary. Let E be an lcHs, G ⊂ E′ a separating subspace, V ∶= +(νj)j∈N an increasing family of weights which is locally bounded away from zero +on an open set Ω ⊂ Rd, P(∂)K a hypoelliptic linear partial differential operator, +CVP (∂)(Ω) a Schwartz space and U ⊂ Ω fix the topology of CVP (∂)(Ω). If +a) G is dense in E′ +b and E locally complete, or +b) E is Br-complete, +and f∶U → E is a function in ℓV(U) such that e′ ○ f admits an extension fe′ ∈ +CVP (∂)(Ω) for each e′ ∈ G, then there is a unique extension F ∈ CVP (∂)(Ω,E) of f. +Proof. The existence of F follows from Proposition 4.2.14, Example 4.2.16 b) +and Theorem 5.2.63 with (T E +m,T K +m)m∈M ∶= (idEΩ,idKΩ). The uniqueness of F is a +result of Proposition 5.2.8. +□ + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +99 +We have the following sufficient conditions on a family of weights V which +guarantee the existence of a countable set U ⊂ C that fixes the topology of CV∂(C) +due to Abanin and Varziev [2]. +5.2.65. Proposition. Let V ∶= (νj)j∈N where νj(z) ∶= exp(ajµ(z)−ϕ(z)), z ∈ C, +with some continuous, subharmonic function µ∶C → [0,∞), a continuous function +ϕ∶C → R and a strictly increasing, positive sequence (aj)j∈N with a ∶= limj→∞ aj ∈ +(0,∞]. Let there be +(i) s ≥ 0 and C > 0 such that ∣ϕ(z) − ϕ(ζ)∣ ≤ C and ∣µ(z) − µ(ζ)∣ ≤ C for all +z,ζ ∈ C with ∣z − ζ∣ ≤ (1 + ∣z∣)−s, +(ii) max(ϕ(z),µ(z)) ≤ ∣z∣q + C0 for some q,C0 > 0 and +(iii) ln(∣z∣) = O(µ(z)) as ∣z∣ → ∞ if a = ∞, or ln(∣z∣) = o(µ(z)) as ∣z∣ → ∞ if +0 < a < ∞. +Let (λk)k∈N be the sequence of simple zeros of a function L ∈ C̃V∂(C) having no +other zeros where ̃V ∶= (ν2 +j /νmj)j∈N for some sequence (mj)j∈N in N. Suppose that +there are j0 ∈ N and a sequence of circles {z ∈ C ∣ ∣z∣ = Rm} with Rm ↗ ∞ such that +∣L(z)∣νj0(z) ≥ Cm, +m ∈ N, z ∈ C, ∣z∣ = Rm, +for some Cm ↗ ∞ and +∣L′(λk)∣νj0(λk) ≥ 1 +for all sufficiently large k ∈ N. +Then CV∂(C) is a nuclear Fréchet space for all a ∈ (0,∞] and U ∶= (λk)k∈N fixes the +topology of CV∂(C) if a = ∞. If µ is a radial function, i.e. µ(z) = µ(∣z∣), z ∈ C, with +µ(2z) ∼ µ(z) as ∣z∣ → ∞, then U fixes the topology of CV∂(C) for all a ∈ (0,∞]. +Proof. First, we check that Condition 4.2.18 is satisfied, which implies that +CV∂(C) is a nuclear Fréchet space by Proposition 4.2.19. We set k ∶= max(s,2) and +observe that (i) is also fulfilled with k instead of s. Let z ∈ C and ∥ζ∥∞,∥η∥∞ ≤ +(1/ +√ +2)(1 + ∣z∣)−k =∶ r(z). From ∣ ⋅ ∣ ≤ +√ +2∥ ⋅ ∥∞ and (i) it follows +∣µ(z + ζ) − µ(z + η)∣ ≤ ∣µ(z + ζ) − µ(z)∣ + ∣µ(z) − µ(z + η)∣ ≤ 2C +and thus µ(z+ζ) ≤ 2C+µ(z+η). In the same way we obtain −ϕ(z+ζ) ≤ 2C−ϕ(z+η). +Hence we have +ajµ(z + ζ) − ϕ(z + ζ) ≤ 2C(aj + 1) + ajµ(z + η) − ϕ(z + η) +for j ∈ N, implying +νj(z + ζ) ≤ e2C(aj+1)νj(z + η), +which means that (α.1) of Condition 4.2.18 holds. By (iii) there are ε > 0 and +R > 0 such that ln(∣z∣) ≤ εµ(z) for all z ∈ C with ∣z∣ ≥ R if a = ∞. This yields for all +∣z∣ ≥ max(2,R) that +ajµ(z) + k ln(1 + ∣z∣) ≤ ajµ(z) + k ln(∣z∣2) = ajµ(z) + 2k ln(∣z∣) ≤ ajµ(z) + 2kεµ(z). +Since a = ∞, there is n ∈ N such that an ≥ aj + 2kε, resulting in +ajµ(z) + k ln(1 + ∣z∣) ≤ anµ(z) +for all ∣z∣ ≥ max(2,R). Therefore we derive +ajµ(z) + k ln(1 + ∣z∣) ≤ anµ(z) + k ln(1 + max(2,R)) +(54) +for all z ∈ C, which means that (α.2) and (α.3) hold with ψj(z) ∶= r(z). If 0 < a < ∞, +for every ε > 0 there is R > 0 such that ln(∣z∣) ≤ εµ(z) for all z ∈ C with ∣z∣ ≥ R by +(iii). Thus we may choose ε > 0 such that aj+1 −aj ≥ 2kε > 0 because (aj) is strictly +increasing. We deduce that (54) with n ∶= j + 1 holds in this case as well and (α.2) +and (α.3), too. +Observing that the condition that U = (λk)k∈N is the sequence of simple zeros +of a function L ∈ C̃V∂(C) means that L ∈ L (Φa +ϕ,µ;U) and (i) that ϕ and µ vary + +100 +5. APPLICATIONS +slowly w.r.t. r(z) ∶= (1 + ∣z∣)−s in the notation of [2, Definition, p. 579, 584] and [2, +p. 585], respectively, the statement that U fixes the topology is a consequence of +[2, Theorem 2, p. 585–586]. +□ +5.2.66. Remark. +a) Let D ⊂ C be convex, bounded and open with 0 ∈ D. +Let ϕ(z) ∶= HD(z) ∶= supζ∈D Re(zζ), z ∈ C, be the supporting function of +D, µ(z) ∶= ln(1 + ∣z∣), z ∈ C, and aj ∶= j, j ∈ N. Then ϕ and µ fulfil the +conditions of Proposition 5.2.65 with a = ∞ by [2, p. 586] and the existence +of an entire function L which fulfils the conditions of Proposition 5.2.65 +is guaranteed by [3, Theorem 1.6, p. 1537]. Thus there is a countable +set U ∶= (λk)k∈N ⊂ C which fixes the topology in A−∞ +D +∶= CV∂(C) with +V ∶= (exp(ajµ − ϕ))j∈N. +b) An explicit construction of a set U ∶= (λk)k∈N ⊂ C which fixes the topology +in A−∞ +D +is given in [1, Algorithm 3.2, p. 3629]. This construction does not +rely on the entire function L. In particular (see [31, p. 15]), for D ∶= D +we have ϕ(z) = ∣z∣, for each k ∈ N we may take lk ∈ N, lk > 2πk2, and set +λk,j ∶= krk,j, 1 ≤ j ≤ lk, where rk,j denote the lk-roots of unity. Ordering +λk,j in a sequence of one index appropriately, we obtain a sequence which +fixes the topology of A−∞ +D . +c) Let µ∶C → [0,∞) be a continuous, subharmonic, radial function which +increases with ∣z∣ and satisfies +(i) supζ∈C,∥ζ∥∞≤r(z) µ(z + ζ) ≤ infζ∈C,∥ζ∥∞≤r(z) µ(z + ζ) + C for some con- +tinuous function r∶C → (0,1] and C > 0, +(ii) ln(1 + ∣z∣2) = o(µ(∣z∣)) as ∣z∣ → ∞, +(iii) µ(2∣z∣) = O(µ(∣z∣)) as ∣z∣ → ∞. +Then V ∶= (exp(−(1/j)µ))j∈N fulfils Condition 4.2.18 where (α.1) follows +from (i) and (α.2), (α.3) as in the proof of Proposition 5.2.65. +Thus +CV∂(C) is a nuclear Fréchet space by Proposition 4.2.19. If µ(∣z∣) = o(∣z∣2) +as ∣z∣ → ∞ or µ(∣z∣) = ∣z∣2, z ∈ C, then U ∶= {αn + iβm ∣ n,m ∈ Z} fixes the +topology in the space A0 +µ ∶= CV∂(C) for any α,β > 0 by [31, Corollary 4.6, +p. 20] and [31, Proposition 4.7, p. 20], respectively. +d) For instance, the conditions on µ in c) are fulfilled for µ(z) ∶= ∣z∣γ, z ∈ +C, with 0 < γ ≤ 2 by [130, 1.5 Examples (3), p. 205]. +If γ = 1, then +A0 +µ = A0 +∂(C) is the space of entire functions of exponential type zero (see +Remark 4.2.20). +e) More general characterisations of countable sets that fix the topology of +CV∂(C) can be found in [2, Theorem 1, p. 580] and [31, Theorem 4.5, p. +17]. +The spaces A0 +µ from c) are known as Hörmander algebras and the space A−∞ +D +considered in a) is isomorphic to the strong dual of the Korenblum space A−∞(D) +via Laplace transform by [132, Proposition 4, p. 580]. +Fν(Ω) a Banach space and E locally complete. For a dom-space Fν(Ω), +a set U that fixes the topology in Fν(Ω) and a separating subspace G ⊂ E′ we have +FνG(U,E)lb ={f ∈ FνG(U,E) ∣ NU(f) bounded in E} +=FνG(U,E) ∩ ℓν(U,E) +where NU(f) ∶= {f(x)ν(x) ∣ x ∈ U}. Let us recall the assumptions of Remark 5.2.26 +but now U fixes the topology. Let (T E,T K) be a strong, consistent family for (F,E) +and a generator for (Fν,E). Let F(Ω) and F(Ω,E) be ε-into-compatible and the +inclusion Fν(Ω) ↪ F(Ω) continuous. Consider a set U which fixes the topology in +Fν(Ω) and a separating subspace G ⊂ E′. For u ∈ F(Ω)εE such that u(B○F (Ω)′ +Fν(Ω) ) is + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +101 +bounded in E we have RU,G(f) ∈ FνG(U,E) with f ∶= S(u) ∈ Fεν(Ω,E) by (50). +Further, T K +x (⋅)ν(x) ∈ B○F (Ω)′ +Fν(Ω) for every x ∈ ω, which implies that +sup +x∈U +pα(RU,G(f)(x))ν(x) = sup +x∈U +pα(u(T K +x (⋅)ν(x))) ≤ +sup +y′∈B○F (Ω)′ +Fν(Ω) +pα(u(y′)) < ∞ +for all α ∈ A by consistency. Hence RU,G(f) ∈ FνG(U,E)lb. Therefore the injective +linear map +RU,G∶Fεν(Ω,E) → FνG(U,E)lb, f ↦ (T E(f)(x))x∈U, +is well-defined and the question we want to answer is: +5.2.67. Question. Let the assumptions of Remark 5.2.26 be fulfilled and U fix +the topology in Fν(Ω). When is the injective restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E)lb, f ↦ (T E(f)(x))x∈U, +surjective? +5.2.68. Proposition ([70, Proposition 3.1, p. 692]). Let E be a locally complete +lcHs, G ⊂ E′ a separating subspace and Z a Banach space whose closed unit ball +BZ is a compact subset of an lcHs Y . Let B1 ⊂ B○Y ′ +Z +such that B○Z +1 +∶= {z ∈ Z ∣ ∀ y′ ∈ +B1 ∶ ∣y′(z)∣ ≤ 1} is bounded in Z. If A∶X ∶= spanB1 → E is a linear map which is +bounded on B1 such that there is a σ(E′,E)-dense subspace G ⊂ E′ with e′ ○ A ∈ Z +for all e′ ∈ G, then there exists a (unique) extension ̂A ∈ Y εE of A such that ̂A(B○Y ′ +Z +) +is bounded in E. +The following theorem is a generalisation of [70, Theorem 3.2, p. 693] and [93, +Theorem 12, p. 5]. +5.2.69. Theorem. Let E be a locally complete lcHs, G ⊂ E′ a separating sub- +space and F(Ω) and F(Ω,E) be ε-into-compatible. Let (T E,T K) be a generator +for (Fν,E) and a strong, consistent family for (F,E), Fν(Ω) a Banach space +whose closed unit ball BFν(Ω) is a compact subset of F(Ω) and U fix the topology +in Fν(Ω). Then the restriction map +RU,G∶Fεν(Ω,E) → FνG(U,E)lb +is surjective. +Proof. Let f ∈ FνG(U,E)lb. We set B1 ∶= {T K +x (⋅)ν(x) ∣ x ∈ U}, X ∶= spanB1, +Y ∶= F(Ω) and Z ∶= Fν(Ω). We have B1 ⊂ Y ′ since (T E,T K) is a consistent family +for (F,E). If f ∈ BZ, then +∣T K +x (f)ν(x)∣ ≤ ∣f∣Fν(Ω) ≤ 1 +for all x ∈ U and thus B1 ⊂ B○Y ′ +Z +. Further on, there is C > 0 such that for all f ∈ B○Z +1 +∣f∣Fν(Ω) ≤ C sup +x∈U +∣T K +x (f)∣ν(x) ≤ C +as U fixes the topology in Z, implying the boundedness of B○Z +1 +in Z. Let A∶X → E +be the linear map determined by +A(T K +x (⋅)ν(x)) ∶= f(x)ν(x). +The map A is well-defined since G is σ(E′,E)-dense, and bounded on B1 be- +cause A(B1) = NU(f). +Let e′ ∈ G and fe′ be the unique element in Fν(Ω) +such that T K(fe′)(x) = (e′ ○ f)(x) for all x ∈ U, which implies T K(fe′)(x)ν(x) = +(e′○A)(T K +x (⋅)ν(x)). Again, this equation allows us to consider fe′ as a linear form on +X (by setting fe′(T K +x (⋅)ν(x)) ∶= (e′○A)(T K +x (⋅)ν(x))), which yields e′○A ∈ Fν(Ω) = Z +for all e′ ∈ G. +Hence we can apply Proposition 5.2.68 and obtain an extension + +102 +5. APPLICATIONS +̂A ∈ Y εE of A such that ̂A(B○Y ′ +Z +) is bounded in E. We set ̃F ∶= S(̂A) ∈ Fεν(Ω,E) +and get for all x ∈ U that +T E( ̃F)(x) = T ES(̂A)(x) = ̂A(T K +x ) = +1 +ν(x)A(T K +x (⋅)ν(x)) = f(x) +by consistency for (F,E), yielding RU,G( ̃F) = f. +□ +5.2.70. Corollary. Let E be a locally complete lcHs, G ⊂ E′ a separating +subspace, Ω ⊂ Rd open, P(∂)K a hypoelliptic linear partial differential operator, +ν∶Ω → (0,∞) continuous and U fix the topology in CνP (∂)(Ω). If f∶U → E is a +function in ℓν(U,E) such that e′ ○ f admits an extension fe′ ∈ CνP (∂)(Ω) for every +e′ ∈ G, then there exists a unique extension F ∈ CνP (∂)(Ω,E) of f. +Proof. Observing that f ∈ FνG(U,E)lb with Fν(Ω) = CνP (∂)(Ω), our state- +ment follows directly from Theorem 5.2.69 whose conditions are fulfilled by the +proof of Corollary 5.2.30. +□ +Sets that fix the topology in CνP (∂)(Ω) for different weights ν are well-studied +if P(∂) = ∂ is the Cauchy–Riemann operator. If Ω ⊂ C is open, P(∂) = ∂ and ν = 1, +then CνP (∂)(Ω) = H∞(Ω) is the space of bounded holomorphic functions on Ω. +Brown, Shields and Zeller characterise the countable discrete sets U ∶= (zn)n∈N ⊂ Ω +that fix the topology in H∞(Ω) with C = 1 and equality in Definition 5.2.46 for +Jordan domains Ω in [36, Theorem 3, p. 167]. In particular, they prove for Ω = D +that a discrete U = (zn)n∈N fixes the topology in H∞(D) if and only if almost every +boundary point is a non-tangential limit of a sequence in U. Bonsall obtains the +same characterisation for bounded harmonic functions, i.e. P(∂) = ∆ and ν = 1, on +Ω = D by [32, Theorem 2, p. 473]. An example of such a set U = (zn)n∈N ⊂ D is +constructed in [36, Remark 6, p. 172]. Probably the first example of a countable +discrete set U ⊂ D that fixes the topology in H∞(D) is given by Wolff in [183, p. +1327] (cf. [81, Theorem (Wolff), p. 402]). In [148, 4.14 Theorem, p. 255] Rubel and +Shields give a charaterisation of sets U ⊂ Ω that fix the topology in H∞(Ω) by means +of bounded complex measures where Ω ⊂ C is open and connected. The existence +of a countable U fixing the topology in H∞(Ω) is shown in [148, 4.15 Proposition, +p. 256]. In the case of several complex variables the existence of such a countable +U is treated by Sibony in [167, Remarques 4 b), p. 209] and by Massaneda and +Thomas in [128, Theorem 2, p. 838]. +If Ω = C and P(∂) = ∂, then CνP (∂)(Ω) =∶ F ∞ +ν (C) is a generalised L∞-version +of the Bargmann–Fock space. In the case that ν(z) = exp(−α∣z∣2/2), z ∈ C, for +some α > 0, Seip and Wallstén show in [162, Theorem 2.3, p. 93] that a countable +discrete set U ⊂ C fixes the topology in F ∞ +ν (C) if and only if U contains a uniformly +discrete subset U ′ with lower uniform density D−(U ′) > α/π (the proof of sufficiency +is given in [165] and the result was announced in [161, Theorem 1.3, p. 324]). A +generalisation of this result using lower angular densities is given by Lyubarski˘ı and +Seip in [124, Theorem 2.2, p. 162] to weights of the form ν(z) = exp(−φ(arg z)∣z∣2/2), +z ∈ C, with a 2π-periodic 2-trigonometrically convex function φ such that φ ∈ +C2([0,2π]) and φ(θ) + (1/4)φ′′(θ) > 0 for all θ ∈ [0,2π]. An extension of the results +in [162] to weights of the form ν(z) = exp(−φ(z)), z ∈ C, with a subharmonic +function φ such that ∆φ(z) ∼ 1 is given in [135, Theorem 1, p. 249] by Ortega- +Cerdà and Seip. +Here, f(x) ∼ g(x) for two functions f,g∶Ω → R means that +there are C1,C2 > 0 such that C1g(x) ≤ f(x) ≤ C2g(x) for all x ∈ Ω. +Marco, +Massaneda and Ortega-Cerdà describe sets that fix the topology in F ∞ +ν (C) with +ν(z) = exp(−φ(z)), z ∈ C, for some subharmonic function φ whose Laplacian ∆φ is +a doubling measure (see [126, Definition 5, p. 868]), e.g. φ(z) = ∣z∣β for some β > 0, + +5.2. EXTENSION OF VECTOR-VALUED FUNCTIONS +103 +in [126, Theorem A, p. 865]. The case of several complex variables is handled by +Ortega-Cerdà, Schuster and Varolin in [136, Theorem 2, p. 81]. +If Ω = D and P(∂) = ∂, then CνP (∂)(Ω) =∶ A∞ +ν (D) is a generalised L∞-version +of the weighted Bergman space (and of H∞(D)). For ν(z) = (1 − ∣z∣2)n, z ∈ D, for +some n ∈ N, Seip proves that a countable discrete set U ⊂ D fixes the topology in +A∞ +ν (D) if and only if U contains a uniformly discrete subset U ′ with lower uniform +density D−(U ′) > n by [163, Theorem 1.1, p. 23], and gives a typical example +in [163, p. 23]. +Later on, this is extended by Seip in [164, Theorem 2, p. 718] +to weights ν(z) = exp(−φ(z)), z ∈ D, with a subharmonic function φ such that +∆φ(z) ∼ (1 − ∣z∣2)−2, e.g. φ(z) = −β ln(1 − ∣z∣2), z ∈ D, for some β > 0. Domański +and Lindström give necessary resp. sufficient conditions for fixing the topology in +A∞ +ν (D) in the case that ν is an essential weight on D, i.e. there is C > 0 with +ν(z) ≤ ̃ν(z) ≤ Cν(z) for each z ∈ D where ̃ν(z) ∶= (sup{∣f(z)∣ ∣ f ∈ BA∞ +ν (D)})−1 is +the associated weight. In [55, Theorem 29, p. 260] they describe necessary resp. +sufficient conditions for fixing the topology if the upper index Uν is finite (see [55, +p. 242]), and necessary and sufficient conditions in [55, Corollary 31, p. 261] if +0 < Lν = Uν < ∞ holds where Lν is the lower index (see [55, p. 243]), which for +example can be applied to ν(z) = (1 − ∣z∣2)n(ln( +e +1−∣z∣))β, z ∈ D, for some n > 0 and +β ∈ R. The case of simply connected open Ω ⊂ C is considered in [55, Corollary 32, +p. 261–262]. +Borichev, Dhuez and Kellay treat A∞ +ν (D) and F ∞ +ν (C) simultaneously. +Let +ΩR ∶= D, if R = 1, and ΩR ∶= C if R = ∞. They take ν(z) = exp(−φ(z)), z ∈ ΩR, +where φ∶[0,R) → [0,∞) is an increasing function such that φ(0) = 0, limr→R φ(r) = +∞, φ is extended to ΩR by φ(z) ∶= φ(∣z∣), φ ∈ C2(ΩR), and, in addition ∆φ(z) ≥ 1 +if R = ∞ (see [33, p. 564–565]). Then they set ρ∶[0,R) → R, ρ(r) ∶= [∆φ(r)]−1/2, +and suppose that ρ decreases to 0 near R, ρ′(r) → 0, r → R, and either (ID) the +function r ↦ ρ(r)(1 − r)−C increases for some C ∈ R and for r close to 1, resp. (IC) +the function r ↦ ρ(r)rC increases for some C ∈ R and for large r, or (IIΩR) that +ρ′(r)ln(1/ρ(r)) → 0, r → R (see [33, p. 567–569]). Typical examples for (ID) are +φ(r) = ln(ln( 1 +1−r))ln( 1 +1−r) +or +φ(r) = +1 +1−r, +a typical example for (IID) is φ(r) = exp( 1 +1−r), for (IC) +φ(r) = r2 ln(ln(r)) +or +φ(r) = rp, for some p > 2, +and a typical example for (IIC) is φ(r) = exp(r). Sets that fix the topology in +A∞ +ν (D) are described by densities in [33, Theorem 2.1, p. 568] and sets that fix the +topology in F ∞ +ν (C) in [33, Theorem 2.5, p. 569]. +Wolf uses sets that fix the topology in A∞ +ν (D) for the characterisation of +weighted composition operators on A∞ +ν (D) with closed range in [182, Theorem +1, p. 36] for bounded ν. +5.2.71. Corollary. Let E be a locally complete lcHs, G ⊂ E′ a separating +subspace, ν∶D → (0,∞) continuous and U ∶= {0} ∪ ({1} × U∗) fix the topology in +Bν(D) with U∗ ⊂ D. If f∶U → E is a function in ℓν∗(U,E) such that there is fe′ ∈ +Bν(D) for each e′ ∈ G with fe′(0) = e′(f(0)) and f ′ +e′(z) = e′(f(1,z)) for all z ∈ U∗, +then there exists a unique F ∈ Bν(D,E) with F(0) = f(0) and (∂1 +C)EF(z) = f(1,z) +for all z ∈ U∗. +Proof. As in Corollary 5.2.70 but with Fν∗(D) = Bν(D) and Corollary 5.2.33 +instead of Corollary 5.2.30. +□ +Sets that fix the topology in Bν(D) play an important role in the characteri- +sation of composition operators on Bν(D) with closed range. Chen and Gauthier +give a characterisation in [42] for weights of the form ν(z) = (1 − ∣z∣2)α, z ∈ D, + +104 +5. APPLICATIONS +for some α ≥ 1. We recall the following definitions which are needed to phrase +this characterisation. For a continuous function ν∶D → (0,∞) and a non-constant +holomorphic function φ∶D → D we set +τ ν +φ(z) ∶= ν(z)∣φ′(z)∣ +ν(φ(z)) , z ∈ D, +and +Ων +ε ∶= {z ∈ D ∣ τ ν +φ(z) ≥ ε}, ε > 0, +and define the pseudohyperbolic distance +ρ(z,w) ∶= ∣ z − w +1 − zw∣, z,w ∈ D. +For 0 < r < 1 a set B ⊂ D is called a pseudo r-net if for every w ∈ D there is z ∈ B +with ρ(z,w) ≤ r (see [42, p. 195]). A set U⋆ ⊂ D is a sampling set for Bν(D) with ν +as above in the sense of [42, p. 198] if and only if {0}∪({1}×U⋆) fixes the topology +in Bν(D) (see the definitions above Corollary 5.2.33). +5.2.72. Theorem ([42, Theorem 3.1, p. 199, Theorem 4.3, p. 202]). Let φ∶D → +D be a non-constant holomorphic function and ν(z) = (1 − ∣z∣2)α, z ∈ D, for some +α ≥ 1. Then the following statements are equivalent. +(i) The composition operator Cφ∶Bν(D) → Bν(D), Cφ(f) ∶= f ○ φ, is bounded +below (i.e. has closed range). +(ii) There is ε > 0 such that {0} ∪ ({1} × φ(Ων +ε)) fixes the topology in Bν(D). +(iii) There are ε > 0 and 0 < r < 1 such that φ(Ων +ε) is a pseudo r-net. +This theorem has some predecessors. The implications (i)⇒(iii) and (iii), r < +1/4 ⇒(i) for α = 1 are due to Ghatage, Yan and Zheng by [72, Proposition 1, +p. 2040] and [72, Theorem 2, p. 2043]. This was improved by Chen to (i)⇔(iii) +for α = 1 by removing the restriction r < 1/4 in [41, Theorem 1, p. 840]. +The +proof of the equivalence (i)⇔(ii) given in [73, Theorem 1, p. 1372] for α = 1 is +due to Ghatage, Zheng and Zorboska. A non-trivial example of a sampling set +for α = 1 can be found in [73, Example 2, p. 1376] (cf. [42, p. 203]). In the case +of several complex variables a characterisation corresponding to Theorem 5.2.72 +is given by Chen in [41, Theorem 2, p. 844] and Deng, Jiang and Ouyang in [50, +Theorem 1–3, p. 1031–1032, 1034] where Ω is the unit ball of Cd. Giménez, Malavé +and Ramos-Fernández extend Theorem 5.2.72 by [75, Theorem 3, p. 112] and [75, +Corollary 1, p. 113] to more general weights of the form ν(z) = µ(1−∣z∣2) with some +continuous function µ∶(0,1] → (0,∞) such that µ(r) → 0, r → 0+, which can be +extended to a holomorphic function µ0 on D1(1) without zeros in D1(1) and fulfilling +µ(1−∣1−z∣) ≤ C∣µ0(z)∣ for all z ∈ D1(1) and some C > 0 (see [75, p. 109]). Examples +of such functions µ are µ1(r) ∶= rα, α > 0, µ2 ∶= r ln(2/r) and µ3(r) ∶= rβ ln(1 − r), +β > 1, for r ∈ (0,1] (see [75, p. 110]) and with ν(z) = µ1(1 − ∣z∣2) = (1 − ∣z∣2)α, z ∈ D, +one gets Theorem 5.2.72 back for α ≥ 1. For 0 < α < 1 and ν(z) = µ1(1−∣z∣2), z ∈ D, +the equivalence (i)⇔(ii) is given in [184, Proposition 4.4, p. 14] of Yoneda as well and +a sufficient condition implying (ii) in [184, Corollary 4.5, p. 15]. Ramos-Fernández +generalises the results given in [75] to bounded essential weights ν on D by [144, +Theorem 4.3, p. 85] and [144, Remark 4.2, p. 84]. In [141, Theorem 2.4, p. 3106] +Pirasteh, Eghbali and Sanatpour use sets that fix the topology in Bν(D) for radial +essential ν to characterise Li–Stević integral-type operators on Bν(D) with closed +range instead of composition operators. The composition operator on the harmonic +variant of the Bloch type space Bν(D) with ν(z) = (1 −∣z∣2)α, z ∈ D, for some α > 0 +is considered by Esmaeili, Estaremi and Ebadian, who give a corresponding result +in [64, Theorem 2.8, p. 542]. +5.3. Weak-strong principles for differentiability of finite order +This section is dedicated to Ck-weak-strong principles for differentiable func- +tions. So the question is: + +5.3. WEAK-STRONG PRINCIPLES FOR DIFFERENTIABILITY OF FINITE ORDER +105 +5.3.1. Question. Let E be an lcHs, G ⊂ E′ a separating subspace, Ω ⊂ Rd +open and k ∈ N0 ∪ {∞}. If f∶Ω → E is such that e′ ○ f ∈ Ck(Ω) for each e′ ∈ G, does +f ∈ Ck(Ω,E) hold? +An affirmative answer to the preceding question is called a Ck-weak-strong +principle. It is a result of Bierstedt [17, 2.10 Lemma, p. 140] that for k = 0 the +C0-weak-strong principle holds if Ω ⊂ Rd is open (or more general a Hausdorff kR- +space), G = E′ and E is such that every bounded set is already precompact in E, +i.e. E is a generalised Schwartz space (see Definition 5.2.50 and Remark 5.2.57). +For instance, the last condition is fulfilled if E is a semi-Montel or Schwartz space. +The C0-weak-strong principle does not hold for general E by [94, Beispiel, p. 232]. +Grothendieck sketches in a footnote [82, p. 39] (cf. [84, Chap. 3, Sect. 8, Corol- +lary 1, p. 134]) the proof that for k < ∞ a weakly-Ck+1 function f∶Ω → E on an +open set Ω ⊂ Rd with values in a quasi-complete lcHs E is already Ck, i.e. that +from e′ ○ f ∈ Ck+1(Ω) for all e′ ∈ E′ it follows f ∈ Ck(Ω,E). A detailed proof of this +statement is given by Schwartz in [158], simultaneously weakening the condition +from quasi-completeness of E to sequential completeness and from weakly-Ck+1 to +weakly-Ck,1 +loc . +5.3.2. Theorem ([158, Appendice, Lemme II, Remarques 10), p. 146–147]). +Let E be a sequentially complete lcHs, Ω ⊂ Rd open and k ∈ N0. +a) If f∶Ω → E is such that e′ ○ f ∈ Ck,1 +loc (Ω) for all e′ ∈ E′, then f ∈ Ck(Ω,E). +b) If f∶Ω → E is such that e′ ○ f ∈ Ck+1(Ω) for all e′ ∈ E′, then f ∈ Ck(Ω,E). +Here Ck,1 +loc (Ω) denotes the space of functions in Ck(Ω) whose partial derivatives +of order k are locally Lipschitz continuous. +Part b) clearly implies a C∞-weak- +strong principle for open Ω ⊂ Rd, G = E′ and sequentially complete E. This can +be generalised to locally complete E. Waelbroeck has shown in [177, Proposition +2, p. 411] and [176, Definition 1, p. 393] that the C∞-weak-strong principle holds if +Ω is a manifold, G = E′ and E is locally complete. It is a result of Bonet, Frerick +and Jordá that the C∞-weak-strong principle still holds by [30, Theorem 9, p. 232] +if Ω ⊂ Rd is open, G ⊂ E′ determines boundedness and E is locally complete. Due +to [104, 2.14 Theorem, p. 20] of Kriegl and Michor an lcHs E is locally complete if +and only if the C∞-weak-strong principle holds for Ω = R and G = E′. +One of the goals of this section is to improve Theorem 5.3.2. We recall the +following definition from Example 4.2.28. For k ∈ N0 the space of k-times con- +tinuously partially differentiable E-valued functions on an open set Ω ⊂ Rd whose +partial derivatives up to order k are continuously extendable to the boundary of Ω +is +Ck(Ω,E) = {f ∈ Ck(Ω,E) ∣ (∂β)Ef cont. extendable on Ω for all β ∈ Nd +0, ∣β∣ ≤ k} +equipped with the system of seminorms given by +∣f∣Ck(Ω),α = +sup +x∈Ω +β∈Nd +0,∣β∣≤k +pα((∂β)Ef(x)), +f ∈ Ck(Ω,E), α ∈ A. +The space of functions in Ck(Ω,E) such that all its k-th partial derivatives are +γ-Hölder continuous with 0 < γ ≤ 1 is given by +Ck,γ(Ω,E) ∶= {f ∈ Ck(Ω,E) ∣ ∀ α ∈ A ∶ ∣f∣Ck,γ(Ω),α < ∞} +where +∣f∣Ck,γ(Ω),α ∶= max(∣f∣Ck(Ω),α, +sup +β∈Nd +0,∣β∣=k +∣(∂β)Ef∣C0,γ(Ω),α) + +106 +5. APPLICATIONS +with +∣f∣C0,γ(Ω),α ∶= sup +x,y∈Ω +x≠y +pα(f(x) − f(y)) +∣x − y∣γ +. +We set Ck,γ(Ω) ∶= Ck,γ(Ω,K) and +ω1 ∶= {β ∈ Nd +0 ∣ ∣β∣ ≤ k} × Ω +and +ω2 ∶= {β ∈ Nd +0 ∣ ∣β∣ = k} × (Ω2 ∖ {(x,x) ∣ x ∈ Ω}) +as well as ω ∶= ω1 ∪ ω2. We define the operator T E∶Ck(Ω,E) → Eω by +T E(f)(β,x) ∶=(∂β)E(f)(x) +, (β,x) ∈ ω1, +T E(f)(β,(x,y)) ∶=(∂β)E(f)(x) − (∂β)E(f)(y) +, (β,(x,y)) ∈ ω2. +and the weight ν∶ω → (0,∞) by +ν(β,x) ∶= 1, (β,x) ∈ ω1, +and +ν(β,(x,y)) ∶= +1 +∣x − y∣γ , (β,(x,y)) ∈ ω2. +By setting F(Ω,E) ∶= Ck(Ω,E) and observing that +∣f∣Ck,γ(Ω),α = sup +x∈ω pα(T E(f)(x))ν(x), +f ∈ Ck,γ(Ω,E), α ∈ A, +we have Fν(Ω,E) = Ck,γ(Ω,E) with generator (T E,T K). +5.3.3. Corollary. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness, Ω ⊂ Rd open and bounded, k ∈ N0 and 0 < γ ≤ 1. In the case k ≥ 1, assume +additionally that Ω has Lipschitz boundary. If f∶Ω → E is such that e′ ○f ∈ Ck,γ(Ω) +for all e′ ∈ G, then f ∈ Ck,γ(Ω,E). +Proof. We take F(Ω) ∶= Ck(Ω) and F(Ω,E) ∶= Ck(Ω,E) and have Fν(Ω) = +Ck,γ(Ω) and Fν(Ω,E) = Ck,γ(Ω,E) with the weight ν and generator (T E,T K) for +(Fν,E) described above. Due to the proof of Example 4.2.28 and Theorem 3.1.12 +the spaces F(Ω) and F(Ω,E) are ε-into-compatible for any lcHs E (the condi- +tion that E has metric ccp in Example 4.2.28 is only needed for ε-compatibility). +Another consequence of Example 4.2.28 is that +T E(S(u))(β,x) = (∂β)E(S(u))(x) = u(δx ○ (∂β)K) = u(T K +β,x), +(β,x) ∈ ω1, +holds for all u ∈ F(Ω)εE, implying +T E(S(u))(β,(x,y)) = T E(S(u))(β,x) − T E(S(u))(β,y) = u(T K +β,x) − u(T K +β,y) += u(T K +β,(x,y)), +(β,(x,y)) ∈ ω2. +Thus (T E,T K) is a consistent family for (F,E) and its strength is easily seen. In +addition, Fν(Ω) = Ck,γ(Ω) is a Banach space by [57, Theorem 9.8, p. 110] (cf. [4, +1.7 Hölderstetige Funktionen, p. 46]) whose closed unit ball is compact in F(Ω) = +Ck(Ω) by [4, 8.6 Einbettungssatz in Hölder-Räumen, p. 338]. Moreover, the ε-into- +compatibility of F(Ω) and F(Ω,E) in combination with the consistency of (T E,T K) +for (F,E) implies Fεν(Ω,E) ⊂ Fν(Ω,E) as linear spaces by Proposition 5.2.25 +c). Hence our statement follows from Theorem 5.2.29 with the set of uniqueness +U ∶= {0} × Ω for (T K,Fν). +□ +5.3.4. Remark. We point out that Corollary 5.3.3 corrects our result [117, +Corollary 5.3, p. 16] by adding the missing assumption that Ω should additionally +have Lipschitz boundary in the case k ≥ 1. This is needed to deduce that the closed +unit ball of Ck,γ(Ω) is compact in Ck(Ω) by [4, 8.6 Einbettungssatz in Hölder- +Räumen, p. 338] (in the notation of [74] Ω having Lipschitz boundary means that +it is a C0,1 domain, see [74, Lemma 6.36, p. 136] and the comments below and +above this lemma). This additional assumption is missing in [57, Theorem 14.32, +p. 232], which is our main reference in [117] for the compact embedding, but it is + +5.3. WEAK-STRONG PRINCIPLES FOR DIFFERENTIABILITY OF FINITE ORDER +107 +needed due to [4, U8.1 Gegenbeispiel zu Einbettungssätzen, p. 365] (cf. [74, p. 53]). +However, this only affects the result [117, Corollary 6.3, p. 21–22] where we have to +add this missing assumption as well (see Corollary 5.4.4 for this). The other results +of [117] derived from [117, Corollary 5.3, p. 16] are not affected by this missing +assumption since they are all a consequence of [117, Corollary 5.4, p. 17] and [117, +Corollary 6.4, p. 22], whose proofs can be adjusted without additional assumptions +(see Corollary 5.3.5 and Corollary 5.4.5 for this). +Next, we use the preceding corollary to generalise the theorem of Grothendieck +and Schwartz on weakly Ck+1-functions. For k ∈ N0 and 0 < γ ≤ 1 we define the +space of k-times continuously partially differentiable E-valued functions with locally +γ-Hölder continuous partial derivatives of k-th order on an open set Ω ⊂ Rd by +Ck,γ +loc (Ω,E) ∶= {f ∈ Ck(Ω,E) ∣ ∀ K ⊂ Ω compact, α ∈ A ∶ ∣f∣K,α < ∞} +where +∣f∣K,α ∶= max(∣f∣Ck(K),α, +sup +β∈Nd +0,∣β∣=k +∣(∂β)Ef∣C0,γ(K),α) +with +∣f∣Ck(K),α ∶= +sup +x∈K +β∈Nd +0,∣β∣≤k +pα((∂β)Ef(x)) +and +∣f∣C0,γ(K),α ∶= sup +x,y∈K +x≠y +pα(f(x) − f(y)) +∣x − y∣γ +. +Further, we set Ck,γ +loc (Ω) ∶= Ck,γ +loc (Ω,K). Using Corollary 5.3.3, we are able to +improve Theorem 5.3.2 to the following form. +5.3.5. Corollary. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness, Ω ⊂ Rd open, k ∈ N0 and 0 < γ ≤ 1. +a) If f∶Ω → E is such that e′ ○f ∈ Ck,γ +loc (Ω) for all e′ ∈ G, then f ∈ Ck,γ +loc (Ω,E). +b) If f∶Ω → E is such that e′ ○f ∈ Ck+1(Ω) for all e′ ∈ G, then f ∈ Ck,1 +loc (Ω,E). +Proof. Let us start with a). Let f∶Ω → E be such that e′ ○ f ∈ Ck,γ +loc (Ω) for +all e′ ∈ G. Let (Ωn)n∈N be an exhaustion of Ω with open, relatively compact sets +Ωn ⊂ Ω with Lipschitz boundaries ∂Ωn (e.g. choose each Ωn as the interior of a +finite union of closed axis-parallel cubes, see the proof of [168, Theorem 1.4, p. 7] +for the construction) that satisfies Ωn ⊂ Ωn+1 for all n ∈ N. Then the restriction of +e′ ○ f to Ωn is an element of Ck,γ(Ωn) for every e′ ∈ G and n ∈ N. Due to Corollary +5.3.3 we obtain that f ∈ Ck,γ(Ωn,E) for every n ∈ N. Thus f ∈ Ck,γ +loc (Ω,E) since +differentiability is a local property and for each compact K ⊂ Ω there is n ∈ N such +that K ⊂ Ωn. +Let us turn to b), i.e. let f∶Ω → E be such that e′ ○ f ∈ Ck+1(Ω) for all e′ ∈ G. +Since Ω ⊂ Rd is open, for every x ∈ Ω there is εx > 0 such that Bεx(x) ⊂ Ω. For all +e′ ∈ G, β ∈ Nd +0 with ∣β∣ = k and w,y ∈ Bεx(x), w ≠ y, it holds that +∣(∂β)K(e′ ○ f)(w) − (∂β)K(e′ ○ f)(y)∣ +∣w − y∣ +≤ Cd max +1≤n≤d +max +z∈Bεx(x) +∣(∂β+en)K(e′ ○ f)(z)∣ +by the mean value theorem applied to the real and imaginary part where Cd ∶= +√ +d +if K = R, and Cd ∶= 2 +√ +d if K = C. +Thus e′ ○ f ∈ Ck,1 +loc (Ω) for all e′ ∈ G since +for each compact set K ⊂ Ω there are m ∈ N and xi ∈ Ω, 1 ≤ i ≤ m, such that +K ⊂ ⋃m +i=1 Bεxi (xi). It follows from part a) that f ∈ Ck,1 +loc (Ω,E). +□ + +108 +5. APPLICATIONS +If Ω = R, γ = 1 and G = E′, then part a) of Corollary 5.3.5 is already known +by [104, 2.3 Corollary, p. 15]. A ‘full’ Ck-weak-strong principle for k < ∞, i.e. the +conditions of part b) imply f ∈ Ck+1(Ω,E), does not hold in general (see [104, p. +11–12]) but it holds if we restrict the class of admissible lcHs E. +5.3.6. Theorem. Let E be a semi-Montel space, G ⊂ E′ determine boundedness, +Ω ⊂ Rd open and k ∈ N. If f∶Ω → E is such that e′ ○ f ∈ Ck(Ω) for all e′ ∈ G, then +f ∈ Ck(Ω,E). +Proof. Let f∶Ω → E be such that e′○f ∈ Ck(Ω) for all e′ ∈ G. Due to Corollary +5.3.5 b) we already know that f ∈ Ck−1,1 +loc +(Ω,E) since semi-Montel spaces are quasi- +complete and thus locally complete. Now, let x ∈ Ω, εx > 0 such that Bεx(x) ⊂ Ω, +β ∈ Nd +0 with ∣β∣ = k − 1 and n ∈ N with 1 ≤ n ≤ d. The set +B ∶= {(∂βf)E(x + hen) − (∂βf)Ef(x) +h +∣ h ∈ R, 0 < h ≤ εx} +is bounded in E because f ∈ Ck−1,1 +loc +(Ω,E). As E is a semi-Montel space, the closure +B is compact in E. Let (hm)m∈N be a sequence in R such that 0 < hm ≤ εx for all +m ∈ N. From the compactness of B we deduce that there is a subnet (hmι)ι∈I of +(hm)m∈N and yx ∈ B with +yx = lim +ι∈I +(∂βf)E(x + hmιen) − (∂βf)Ef(x) +hmι +=∶ lim +ι∈I yι. +Further, we note that the limit +(∂β+en)K(e′ ○ f)(x) = +lim +h→0 +h∈R,h≠0 +∂β(e′ ○ f)(x + hen) − ∂β(e′ ○ f)(x) +h +(55) +exists for all e′ ∈ G and that (e′(yι))ι∈I is a subnet of the net of difference quotients +on the right-hand side of (55) as (∂β)K(e′ ○ f) = e′ ○ (∂β)Ef. Therefore +(∂β+en)K(e′ ○ f)(x) = +lim +h→0 +h∈R,h≠0 +e′((∂β)Ef(x + hen) − (∂β)Ef(x) +h +) += +lim +h→0 +h∈R,0 1? +(iii) For every ε > 0 does there exist a function g ∈ Ck(R,E) such that λ({x ∈ +Ω ∣ f(x) ≠ g(x)}) < ε in Corollary 5.3.8 where λ is the one-dimensional +Lebesgue measure. In the case that E = Rn this is true by [66, Theorem +3.1.15, p. 227]. + +110 +5. APPLICATIONS +(iv) Is there a ‘Radon–Nikodým type’ characterisation of generalised Gelfand +spaces as in the Banach case? +5.4. Vector-valued Blaschke theorems +In this section we prove several convergence theorems for Banach-valued func- +tions in the spirit of Blaschke’s convergence theorem [38, Theorem 7.4, p. 219] +as it is done in [7, Theorem 2.4, p. 786] and [7, Corollary 2.5, p. 786–787] for +bounded holomorphic functions and more general in [70, Corollary 4.2, p. 695] for +bounded functions in the kernel of a hypoelliptic linear partial differential operator. +Blaschke’s convergence theorem says that if (zn)n∈N ⊂ D is a sequence of distinct +elements with ∑n∈N(1 − ∣zn∣) = ∞ and if (fk)k∈N is a bounded sequence in H∞(D) +such that (fk(zn))k converges in C for each n ∈ N, then there is f ∈ H∞(D) such +that (fk)k converges uniformly to f on the compact subsets of D, i.e. w.r.t. to τc. +5.4.1. Proposition ([70, Proposition 4.1, p. 695]). Let (E,∥ ⋅ ∥) be a Banach +space, Z a Banach space whose closed unit ball BZ is a compact subset of an lcHs +Y and let (Aι)ι∈I be a net in Y εE such that +sup +ι∈I +{∥Aι(y)∥ ∣ y ∈ B○Y ′ +Z +} < ∞. +Assume further that there exists a σ(Y ′,Z)-dense subspace X ⊂ Y ′ such that +limι Aι(x) exists for each x ∈ X. Then there is A ∈ Y εE with A(B○Y ′ +Z +) bounded +and limι Aι = A uniformly on the equicontinuous subsets of Y ′, i.e. for all equicon- +tinuous B ⊂ Y ′ and ε > 0 there exists ς ∈ I such that +sup +y∈B +∥Aι(y) − A(y)∥ < ε +for each ι ≥ ς. +Next, we generalise [70, Corollary 4.2, p. 695]. +5.4.2. Corollary. Let (E,∥ ⋅ ∥) be a Banach space and F(Ω) and F(Ω,E) be +ε-into-compatible. Let (T E,T K) be a generator for (Fν,E) and a strong, consistent +family for (F,E), Fν(Ω) a Banach space whose closed unit ball BFν(Ω) is a compact +subset of F(Ω) and U a set of uniqueness for (T K,Fν). +If (fι)ι∈I ⊂ Fεν(Ω,E) is a bounded net in Fν(Ω,E) such that limι T E(fι)(x) +exists for all x ∈ U, then there is f ∈ Fεν(Ω,E) such that (fι)ι∈I converges to f in +F(Ω,E). +Proof. We set X ∶= span{T K +x ∣ x ∈ U}, Y ∶= F(Ω) and Z ∶= Fν(Ω). As in +the proof of Theorem 5.2.29 we observe that X is σ(Y ′,Z)-dense in Y ′. +From +(fι)ι∈I ⊂ Fεν(Ω,E) follows that there are Aι ∈ F(Ω)εE with S(Aι) = fι for all ι ∈ I. +Since (fι)ι∈I is a bounded net in Fν(Ω,E), we note that +sup +ι∈I +sup +x∈ω ∥Aι(T K +x (⋅)ν(x))∥ = sup +ι∈I +sup +x∈ω ∥T ES(Aι)(x)∥ν(x) = sup +ι∈I +sup +x∈ω ∥T Efι(x)∥ν(x) += sup +ι∈I +∣fι∣Fν(Ω,E) < ∞ +by consistency. Further, limι S(Aι)(T K +x ) = limι T E(fι)(x) exists for each x ∈ U, +implying the existence of limι S(Aι)(x) for each x ∈ X by linearity. +We apply +Proposition 5.4.1 and obtain f ∶= S(A) ∈ Fεν(Ω,E) such that (Aι)ι∈I converges +to A in F(Ω)εE. From F(Ω) and F(Ω,E) being ε-into-compatible it follows that +(fι)ι∈I converges to f in F(Ω,E). +□ + +5.4. VECTOR-VALUED BLASCHKE THEOREMS +111 +First, we apply the preceding corollary to the space C[γ] +z (Ω,E) of γ-Hölder +continuous functions on Ω that vanish at a fixed point z ∈ Ω from Example 4.2.9 +a). We recall that for a metric space (Ω,d), z ∈ Ω, an lcHs E and 0 < γ ≤ 1 we have +C[γ] +z (Ω,E) = {f ∈ EΩ ∣ f(z) = 0 and ∀ α ∈ A ∶ ∣f∣C0,γ(Ω),α < ∞}. +Further, we set ω ∶= Ω2 ∖ {(x,x) ∣ x ∈ Ω}, F(Ω,E) ∶= {f ∈ C(Ω,E) ∣ f(z) = 0} and +T E∶F(Ω,E) → Eω, T E(f)(x,y) ∶= f(x) − f(y), and +ν∶ω → [0,∞), ν(x,y) ∶= +1 +d(x,y)γ . +Then we have for every α ∈ A that +∣f∣C0,γ(Ω),α = sup +x∈ω pα(T E(f)(x))ν(x), +f ∈ C[γ] +z (Ω,E), +and observe that Fν(Ω,E) = C[γ] +z (Ω,E) with generator (T E,T K). +5.4.3. Corollary. Let E be a Banach space, (Ω,d) a metric space, z ∈ Ω and +0 < γ ≤ 1. If (fι)ι∈I is a bounded net in C[γ] +z (Ω,E) such that limι fι(x) exists for all +x in a dense subset U ⊂ Ω, then there is f ∈ C[γ] +z (Ω,E) such that (fι)ι∈I converges +to f in C(Ω,E) uniformly on compact subsets of Ω. +Proof. We choose F(Ω) ∶= {f ∈ C(Ω) ∣ f(z) = 0} and F(Ω,E) ∶= {f ∈ +C(Ω,E) ∣ f(z) = 0}. Then we have Fν(Ω) = C[γ] +z (Ω) and Fν(Ω,E) = C[γ] +z (Ω,E) +with the weight ν and generator (T E,T K) for (Fν,E) described above. Due to [17, +3.1 Bemerkung, p. 141] the spaces F(Ω) and F(Ω,E), equipped with the topology +τc of compact convergence, are ε-compatible. Obviously, (T E,T K) is a strong, con- +sistent family for (F,E). In addition, Fν(Ω) = C[γ] +z (Ω) is a Banach space by [179, +Proposition 1.6.2, p. 20]. For all f from the closed unit ball BFν(Ω) of Fν(Ω) we +have +∣f(x) − f(y)∣ ≤ d(x,y)γ, +x,y ∈ Ω, +and +∣f(x)∣ = ∣f(x) − f(z)∣ ≤ d(x,z)γ, +x ∈ Ω. +It follows that BFν(Ω) is (uniformly) equicontinuous and {f(x) ∣ f ∈ BFν(Ω)} is +bounded in K for all x ∈ Ω. Ascoli’s theorem (see e.g. [133, Theorem 47.1, p. 290]) +implies the compactness of BFν(Ω) in F(Ω) (see also [118, 3.7 Theorem (a), p. 10]). +Furthermore, the ε-compatibility of F(Ω) and F(Ω,E) in combination with the +consistency of (T E,T K) for (F,E) gives Fεν(Ω,E) = Fν(Ω,E) as linear spaces by +Proposition 5.2.25 c). We note that limι fι(x) = limι T E(fι)(x,z) for all x in U, +proving our claim by Corollary 5.4.2. +□ +The space C[γ] +z (Ω) is named Lip0(Ωγ) in [179] (see [179, Definition 1.6.1 (b), p. +19] and [179, Definition 1.1.2, p. 2]). Corollary 5.4.3 generalises [179, Proposition +2.1.7, p. 38] (in combination with [179, Proposition 1.2.4, p. 5]) where Ω is compact, +U = Ω and E = K. +5.4.4. Corollary. Let E be a Banach space, Ω ⊂ Rd open and bounded, k ∈ N0 +and 0 < γ ≤ 1. In the case k ≥ 1, assume additionally that Ω has Lipschitz boundary. +If (fι)ι∈I is a bounded net in Ck,γ(Ω,E) such that +(i) limι fι(x) exists for all x in a dense subset U ⊂ Ω, or if +(ii) limι(∂en)Efι(x) exists for all 1 ≤ n ≤ d and x in a dense subset U ⊂ Ω, Ω +is connected and there is x0 ∈ Ω such that limι fι(x0) exists and k ≥ 1, +then there is f ∈ Ck,γ(Ω,E) such that (fι)ι∈I converges to f in Ck(Ω,E). + +112 +5. APPLICATIONS +Proof. As in Corollary 5.3.3 we take F(Ω) ∶= Ck(Ω) and F(Ω,E) ∶= Ck(Ω,E) +as well as Fν(Ω) ∶= Ck,γ(Ω) and Fν(Ω,E) ∶= Ck,γ(Ω,E) with the weight ν and +generator (T E,T K) for (Fν,E) described above of Corollary 5.3.3. By the proof +of Corollary 5.3.3 all conditions of Corollary 5.4.2 are satisfied, which implies our +statement. +□ +We recall that CWk(Ω,E) is the space Ck(Ω,E) equipped with its usual topol- +ogy for an open set Ω ⊂ Rd, k ∈ N∞ ∪ {0} and an lcHs E (see Example 3.1.9 b) for +k ∈ N∞ and the definition above Proposition 3.1.11 for k = 0). +5.4.5. Corollary. Let E be a Banach space, Ω ⊂ Rd open, k ∈ N0 and 0 < γ ≤ 1. +If (fι)ι∈I is a bounded net in Ck,γ +loc (Ω,E) such that +(i) limι fι(x) exists for all x in a dense subset U ⊂ Ω, or if +(ii) limι(∂en)Efι(x) exists for all 1 ≤ n ≤ d and x in a dense subset U ⊂ Ω, Ω +is connected and there is x0 ∈ Ω such that limι fι(x0) exists and k ≥ 1, +then there is f ∈ Ck,γ +loc (Ω,E) such that (fι)ι∈I converges to f in CWk(Ω,E). +Proof. Let (Ωn)n∈N be an exhaustion of Ω with open, relatively compact sets +Ωn ⊂ Ω such that Ωn has Lipschitz boundary, Ωn ⊂ Ωn+1 for all n ∈ N and, in +addition, x0 ∈ Ω1 and Ωn is connected for each n ∈ N in case (ii) (see the proof of +Corollary 5.3.5). The restriction of (fι)ι∈I to Ωn is a bounded net in Ck,γ(Ωn,E) +for each n ∈ N. By Corollary 5.4.4 there is Fn ∈ Ck,γ(Ωn,E) for each n ∈ N such +that the restriction of (fι)ι∈I to Ωn converges to Fn in Ck(Ωn,E) since U ∩ Ωn is +dense in Ωn due to Ωn being open and x0 being an element of the connected set +Ωn in case (ii). The limits Fn+1 and Fn coincide on Ωn for each n ∈ N. Thus the +definition f ∶= Fn on Ωn for each n ∈ N gives a well-defined function f ∈ Ck,γ +loc (Ω,E), +which is a limit of (fι)ι∈I in CWk(Ω,E). +□ +5.4.6. Corollary. Let E be a Banach space, Ω ⊂ Rd open and k ∈ N0. If +(fι)ι∈I is a bounded net in Ck+1(Ω,E) such that +(i) limι fι(x) exists for all x in a dense subset U ⊂ Ω, or if +(ii) limι(∂en)Efι(x) exists for all 1 ≤ n ≤ d and x in a dense subset U ⊂ Ω, Ω +is connected and there is x0 ∈ Ω such that limι fι(x0) exists, +then there is f ∈ Ck,1 +loc (Ω,E) such that (fι)ι∈I converges to f in CWk(Ω,E). +Proof. By Corollary 5.3.5 b) (fι)ι∈I is a bounded net in Ck,1 +loc (Ω,E). Hence +our statement is a consequence of Corollary 5.4.5. +□ +The preceding result directly implies a C∞-smooth version. +5.4.7. Corollary. Let E be a Banach space and Ω ⊂ Rd open. If (fι)ι∈I is a +bounded net in C∞(Ω,E) such that +(i) limι fι(x) exists for all x in a dense subset U ⊂ Ω, or if +(ii) limι(∂en)Efι(x) exists for all 1 ≤ n ≤ d and x in a dense subset U ⊂ Ω, Ω +is connected and there is x0 ∈ Ω such that limι fι(x0) exists, +then there is f ∈ C∞(Ω,E) such that (fι)ι∈I converges to f in CW∞(Ω,E). +Now, we turn to weighted kernels of hypoelliptic linear partial differential op- +erators. +5.4.8. Corollary. Let E be a Banach space, Ω ⊂ Rd open, P(∂)K a hypoelliptic +linear partial differential operator, ν∶Ω → (0,∞) continuous and U ⊂ Ω a set of +uniqueness for (idKΩ,CνP (∂)). If (fι)ι∈I is a bounded net in (CνP (∂)(Ω,E),∣ ⋅ ∣ν) +such that limι fι(x) exists for all x ∈ U, then there is f ∈ CνP (∂)(Ω,E) such that +(fι)ι∈I converges to f in (C∞ +P (∂)(Ω,E),τc). + +5.4. VECTOR-VALUED BLASCHKE THEOREMS +113 +Proof. Our statement follows from Corollary 5.4.2 since by the proof of Corol- +lary 5.2.30 all conditions needed are fulfilled. +□ +For ν = 1 on Ω the preceding corollary is included in [70, Corollary 4.2, p. +695] but then an even better result is available, whose proof we prepare next. +We recall the definition of the space (C∞ +P (∂),b(Ω,E),β) with the strict topology β +from Proposition 4.2.24. For an open set Ω ⊂ Rd, an lcHs E and a linear partial +differential operator P(∂)E which is hypoelliptic if E = K the space of bounded +zero solutions is +C∞ +P (∂),b(Ω,E) = {f ∈ C∞ +P (∂)(Ω,E) ∣ ∀ α ∈ A ∶ ∥f∥∞,α = sup +x∈Ω +pα(f(x)) < ∞}. +We equip this space with strict topology β induced by the seminorms +∣f∣̃ν,α ∶= sup +x∈Ω +pα(f(x))∣̃ν(x)∣, +f ∈ C∞ +P (∂),b(Ω,E), +for ̃ν ∈ C0(Ω). Now, we phrase for C∞ +P (∂),b(Ω,E) = CνP (∂)(Ω,E) with ν = 1 on Ω +the improved version of Corollary 5.4.8. +5.4.9. Corollary. Let E be a Banach space, Ω ⊂ Rd open, P(∂)K a hypoelliptic +linear partial differential operator and U ⊂ Ω a set of uniqueness for (idKΩ,C∞ +P (∂),b). +If (fι)ι∈I is a bounded net in (C∞ +P (∂),b(Ω,E),∥ ⋅ ∥∞) such that limι fι(x) exists +for all x ∈ U, then there is f ∈ C∞ +P (∂),b(Ω,E) such that (fι)ι∈I converges to f in +(C∞ +P (∂),b(Ω,E),β). +Proof. We take F(Ω) ∶= (C∞ +P (∂),b(Ω),β) and F(Ω,E) ∶= (C∞ +P (∂),b(Ω,E),β) as +well as Fν(Ω) ∶= (C∞ +P (∂),b(Ω),∥ ⋅ ∥∞) and Fν(Ω,E) ∶= (C∞ +P (∂),b(Ω,E),∥ ⋅ ∥∞) with +the weight ν(x) ∶= 1, x ∈ Ω, and generator (idEΩ,idΩK) for (Fν,E). The generator +is strong and consistent for (F,E) and F(Ω) and F(Ω,E) are ε-compatible by +Proposition 4.2.24. The space Fν(Ω) is a Banach space as a closed subspace of +the Banach space (Cb(Ω),∥ ⋅ ∥∞). Its closed unit ball BFν(Ω) is τc-compact because +(C∞ +P (∂)(Ω),τc) is a Fréchet–Schwartz space, in particular, a Montel space. Thus +BFν(Ω) is ∥ ⋅ ∥∞-bounded and τc-compact, which implies that it is also β-compact +by [45, Proposition 1 (viii), p. 586] and [45, Proposition 3, p. 590]. In addition, +the ε-compatibility of F(Ω) and F(Ω,E) in combination with the consistency of +(idEΩ,idKΩ) for (F,E) gives Fεν(Ω,E) = Fν(Ω,E) as linear spaces by Proposition +5.2.25 c), verifying our statement by Corollary 5.4.2. +□ +A direct consequence of Corollary 5.4.9 is the following remark. +5.4.10. Remark. Let E be a Banach space, Ω ⊂ Rd open, P(∂)K a hypoel- +liptic linear partial differential operator and (fι)ι∈I a bounded net in the space +(C∞ +P (∂),b(Ω,E),∥ ⋅ ∥∞). Then the following statements are equivalent: +(i) (fι) converges pointwise, +(ii) (fι) converges uniformly on compact subsets of Ω, +(iii) (fι) is β-convergent. +In the case of complex-valued bounded holomorphic functions of one variable, +i.e. E = C, Ω ⊂ C is open and P(∂) = ∂ is the Cauchy–Riemann operator, conver- +gence w.r.t. β is known as bounded convergence (see [147, p. 13–14, 16]) and the +preceding remark is included in [148, 3.7 Theorem, p. 246] for connected sets Ω. +A similar improvement of Corollary 5.4.3 for the space C[γ] +z (Ω,E) of γ-Hölder +continuous functions on a metric space (Ω,d) that vanish at a given point z ∈ Ω is + +114 +5. APPLICATIONS +possible, using the strict topology β on C[γ] +z (Ω) given by the seminorms +∣f∣ν ∶= sup +x,y∈Ω +x≠y +∣f(x) − f(y)∣ +∣x − y∣γ +∣ν(x,y)∣, +f ∈ C[γ] +z (Ω), +for ν ∈ C0(ω) with ω = Ω2 ∖ {(x,x) ∣ x ∈ Ω}. If Ω is compact and E a Banach +space, this follows as in Corollary 5.4.9 from the observation that β is the mixed +topology γ(∣ ⋅ ∣C0,γ(Ω),τc) by [90, Theorem 3.3, p. 645], that a set is β-compact if +and only if it is ∣⋅∣C0,γ(Ω)-bounded and τc-compact by [90, Theorem 2.1 (6), p. 642], +the ε-compatibility (C[γ] +z (Ω),β)εE ≅ (C[γ] +z (Ω,E),γτγ) by [90, Theorem 4.4, p. 648] +where the topology γτγ is described in [90, Definition 4.1, p. 647] and coincides +with β if E = K by [90, Proposition 4.3 (i), p. 647]. +Let us turn to Bloch type spaces. The result corresponding to Corollary 5.4.8 +for Bloch type spaces reads as follows. +5.4.11. Corollary. Let E be a Banach space, ν∶D → (0,∞) continuous and +U∗ ⊂ D have an accumulation point in D. If (fι)ι∈I is a bounded net in Bν(D,E) +such that limι fι(0) and limι(∂1 +C)Efι(z) exist for all z ∈ U∗, then there is f ∈ +Bν(D,E) such that (fι)ι∈I converges to f in (O(D,E),τc). +Proof. Due to the proof of Corollary 5.2.33 all conditions needed to apply +Corollary 5.4.2 are fulfilled, which proves our statement. +□ +5.5. Wolff type results +The following theorem gives us a Wolff type description of the dual of F(Ω) and +generalises [70, Theorem 3.3, p. 693] and [70, Corollary 3.4, p. 694] whose proofs +only need a bit of adaptation. Wolff’s theorem [183, p. 1327] (cf. [81, Theorem +(Wolff), p. 402]) phrased in a functional analytic way (see [70, p. 240]) says: if +Ω ⊂ C is a domain (i.e. open and connected), then for each µ ∈ O(Ω)′ there are +a sequence (zn)n∈N which is relatively compact in Ω and a sequence (an)n∈N in ℓ1 +such that µ = ∑∞ +n=1 anδzn. +5.5.1. Theorem. Let F(Ω) and F(Ω,E) be ε-into-compatible, (T E,T K) be a +generator for (Fν,E) and a strong, consistent family for (F,E) for every Banach +space E. Let F(Ω) be a nuclear Fréchet space and Fν(Ω) a Banach space whose +closed unit ball BFν(Ω) is a compact subset of F(Ω) and (xn)n∈N fixes the topology +in Fν(Ω). +a) Then there is 0 < λ ∈ ℓ1, i.e. λ ∈ ℓ1 and λn > 0 for all n ∈ N, such that for +every bounded B ⊂ F(Ω)′ +b there is C ≥ 1 with +{µ∣Fν(Ω) ∣ µ ∈ B} ⊂ { +∞ +∑ +n=1 +anν(xn)T K +xn ∈ Fν(Ω)′ ∣ a ∈ ℓ1, ∀ n ∈ N ∶ ∣an∣ ≤ Cλn}. +b) Let (∥ ⋅ ∥k)k∈N denote the system of seminorms generating the topology of +F(Ω). Then there is a decreasing zero sequence (εn)n∈N such that for all +k ∈ N there is C ≥ 1 with +∥f∥k ≤ C sup +n∈N +∣T K(f)(xn)∣ν(xn)εn, +f ∈ Fν(Ω). +Proof. We start with part a). +Let B1 ∶= {T K +xn(⋅)ν(xn) ∣ n ∈ N} ⊂ F(Ω)′, +X ∶= spanB1, Y ∶= F(Ω), Z ∶= Fν(Ω) and E1 ∶= {∑∞ +n=1 anν(xn)T K +xn ∣ a ∈ ℓ1}. From +∣j1(a)(f)∣ ∶= ∣ +∞ +∑ +n=1 +anν(xn)T K +xn(f)∣ ≤ sup +n∈N +∣T K(f)(xn)∣ν(xn)∥a∥ℓ1 ≤ ∣f∣Fν(Ω)∥a∥ℓ1 +for all f ∈ Fν(Ω) and a ∈ ℓ1 it follows that E1 is a linear subspace of Fν(Ω)′ +and the continuity of the map j1∶ℓ1 → Fν(Ω)′ where Fν(Ω)′ is equipped with the + +5.5. WOLFF TYPE RESULTS +115 +operator norm. In addition, we deduce that the linear map j∶ℓ1/kerj1 → Fν(Ω)′, +j([a]) ∶= j1(a), where [a] denotes the equivalence class of a ∈ ℓ1 in the quotient +space ℓ1/kerj1, is continuous w.r.t. the quotient norm since +∥j([a])∥Fν(Ω)′ ≤ +inf +b∈ℓ1,[b]=[a]∥b∥ℓ1 = ∥[a]∥ℓ1/ ker j1. +By setting E ∶= j(ℓ1/kerj1) and ∥j([a])∥E ∶= ∥[a]∥ℓ1/ ker j1, a ∈ ℓ1, and observing +that ℓ1/kerj1 is a Banach space, we obtain that E is also a Banach space, which is +continuously embedded in Fν(Ω)′. +We denote by A∶X → E the restriction to Z = Fν(Ω) determined by +A(T K +xn(⋅)ν(xn)) ∶= T K +xn(⋅)∣Fν(Ω)ν(xn) = j([en]) +where en is the n-th unit sequence in ℓ1. We consider Fν(Ω) as a subspace of E′ +via +f(j([a])) ∶= j([a])(f) = +∞ +∑ +n=1 +anν(xn)T K(f)(xn), +a ∈ ℓ1, +for f ∈ Fν(Ω). The space G ∶= Fν(Ω) clearly separates the points of E, thus is +σ(E′,E)-dense and +(f ○ A)(T K +xn(⋅)ν(xn)) = A(T K +xn(⋅)ν(xn))(f) = j([en])(f) = f(j([en])) +for all n ∈ N. Hence we may consider f ○ A by identification with f as an element +of Z = Fν(Ω) for all f ∈ G = Fν(Ω). It follows from Proposition 5.2.68 that there +is a unique extension ̂A ∈ F(Ω)εE of A such that S(̂A) ∈ Fεν(Ω,E). +For each e′ ∈ E′ there are C0,C1 > 0 and an absolutely convex compact set +K ⊂ F(Ω) such that +∣(e′ ○ ̂A)(µ)∣ ≤ C0∥̂A(µ)∥E ≤ C0C1 sup +f∈K +∣µ(f)∣ +for all µ ∈ F(Ω)′, implying e′ ○ ̂A ∈ (F(Ω)′ +b)′. Due to the reflexivity of the nuclear +Fréchet space F(Ω) we obtain e′ ○ ̂A ∈ F(Ω) for each e′ ∈ E′. Further, for each +e′ ∈ E′ we have +∥e′ ○ ̂A∥Fν(Ω) = sup +x∈ω ∣T K(e′ ○ ̂A)(x)∣ν(x) = sup +x∈ω ∣(e′ ○ ̂A)(T K +x (⋅)ν(x))∣ +≤ C0 sup +x∈ω ∥̂A(T K +x (⋅)ν(x))∥E < ∞ +since ̂A(B○F (Ω)′ +Fν(Ω) ) is bounded in E. This yields e′ ○ ̂A ∈ Fν(Ω) for each e′ ∈ E′. In +particular, we get that ̂A is σ(F(Ω)′,Fν(Ω))-σ(E,E′) continuous. The restriction +r∶F(Ω)′ → Fν(Ω)′, r(µ) ∶= µ∣Fν(Ω), is σ(F(Ω)′,Fν(Ω))-σ(Fν(Ω)′,Fν(Ω)) contin- +uous and coincides with ̂A on the σ(F(Ω)′,Fν(Ω))-dense subspace X = spanB1 ⊂ +F(Ω)′. Therefore ̂A(µ) = r(µ) = µ∣Fν(Ω) for all µ ∈ F(Ω)′. +Let B be an absolutely convex, closed and bounded subset of F(Ω)′ +b. +We +endow W ∶= spanB with the Minkowski functional of B. Due to the nuclearity +of F(Ω), there are an absolutely convex, closed and bounded subset V ⊂ F(Ω)′ +b, +(w′ +k)k∈N ⊂ BW ′, (µk)k∈N ⊂ V and 0 ≤ γ ∈ ℓ1, i.e. γ ∈ ℓ1 and γn ≥ 0 for all n ∈ N, such +that +µ = +∞ +∑ +k=1 +γkw′ +k(µ)µk, +µ ∈ B, +by [24, 2.9.1 Theorem, p. 134, 2.9.2 Definition, p. 135]. The boundedness of ̂A(V ) +in E and the definition of E give us a bounded sequence ([β(k)])k∈N ⊂ E with +µk∣Fν(Ω) = ̂A(µk) = +∞ +∑ +n=1 +β(k) +n ν(xn)T K +xn + +116 +5. APPLICATIONS +for all k ∈ N. The sequence (β(k))k∈N ⊂ ℓ1 is also bounded by [131, Remark 5.11, p. +36] and we set ρn ∶= ∑∞ +k=1 γk∣β(k) +n ∣ for n ∈ N. With ρ ∶= (ρn)n∈N we have +∥ρ∥ℓ1 = +∞ +∑ +n=1 +∞ +∑ +k=1 +γk∣β(k) +n ∣ ≤ +∞ +∑ +n=1 +sup +l∈N +∣β(l) +n ∣ +∞ +∑ +k=1 +γk = sup +l∈N +∥β(l)∥ℓ1∥γ∥ℓ1 < ∞, +which means that ρ ∈ ℓ1. For every µ ∈ B we set an ∶= ∑∞ +k=1 γkw′ +k(µ)β(k) +n , n ∈ N, +and conclude that a ∈ ℓ1 with ∣an∣ ≤ ρn for all n ∈ N and +µ∣Fν(Ω) = +∞ +∑ +n=1 +anν(xn)T K +xn. +(57) +The strong dual F(Ω)′ +b of the Fréchet–Schwartz space F(Ω) is a DFS-space and +thus there is a fundamental sequence of bounded (closed, absolutely convex) sets +(Bl)l∈N in F(Ω)′ +b by [131, Proposition 25.19, p. 303]. Due to our preceding results +there is ρ(l) ∈ ℓ1 with (57) for each l ∈ N. Finally, part a) follows from choosing +0 < λ ∈ ℓ1 such that each ρ(l) is componentwise smaller than a multiple of λ, i.e. +we choose λ in a way that for each l ∈ N there is Cl ≥ 1 with ρ(l) +n +≤ Clλn for all +n ∈ N (w.l.o.g. we may assume (the worst case) that limn→∞ ρ(l+1) +n +/ρ(l) +n = ∞ for each +l ∈ N. Then the construction of a suitable 0 < λ ∈ ℓ1 is given in [97, Chap. IX, §41, +7., p. 301–302]: set c(l) +n ∶= ρ(l) +n +for all l,n ∈ N and define λn ∶= cn + +1 +n2 for all n ∈ N +with the (cn) ∈ ℓ1 constructed there. Then set C1 ∶= 1 and Cl ∶= (max{c(l) +n ∣ 1 ≤ n ≤ +nl−1}/min{λn ∣ 1 ≤ n ≤ nl−1}) + 1, l ≥ 2, for the sequence of indices (nl)l∈N from the +construction of (cn).). +Let us turn to part b). We choose λ ∈ ℓ1 from part a) and a decreasing zero +sequence (εn)n∈N such that ( λn +εn )n∈N still belongs to ℓ1 (e.g. take εn ∶= (∑∞ +k=n λk)1/2 +for n ∈ N by [97, Chap. IX, §39, Theorem of Dini, p. 293]). For k ∈ N we set +̃Bk ∶= {f ∈ F(Ω) ∣ ∥f∥k ≤ 1} +and note that the polar ̃B○ +k is bounded in F(Ω)′ +b. Due to part a) there exists C ≥ 1 +such that +̂A( ̃B○ +k) ⊂ { +∞ +∑ +n=1 +anν(xn)T K +xn ∈ Fν(Ω)′ ∣ a ∈ ℓ1, ∀ n ∈ N ∶ ∣an∣ ≤ Cλn}. +By [131, Proposition 22.14, p. 256] the formula +∥f∥k = sup +y′∈ ̃ +B○ +k +∣y′(f)∣, +f ∈ F(Ω), +is valid and hence +∥f∥k = sup +y′∈ ̃ +B○ +k +∣r(y′)(f)∣ = sup +y′∈ ̃ +B○ +k +∣̂A(y′)(f)∣ ≤ C sup +a∈ℓ1 +∣an∣≤λn +∣ +∞ +∑ +n=1 +anν(xn)T K(f)(xn)∣ +≤ C∥(λn +εn +) +n∥ +ℓ1 sup +n∈N +∣T K(f)(xn)∣ν(xn)εn +for all f ∈ Fν(Ω). +□ +5.5.2. Remark. The proof of Theorem 5.5.1 shows it is not needed that the +assumption that F(Ω) and F(Ω,E) are ε-into-compatible, (T E,T K) is a generator +for (Fν,E) and a strong, consistent family for (F,E) is fulfilled for every Banach +space E. It is sufficient that it is fulfilled for the Banach space E ∶= j(ℓ1/kerj1). +We recall from (53) that for a positive sequence ν ∶= (νn)n∈N and an lcHs E we +have +ℓν(N,E) = {x = (xn)n∈N ∈ EN ∣ ∀ α ∈ A ∶ ∥x∥α = sup +n∈N +pα(xn)νn < ∞}. + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +117 +Further, we equip the space EN of all sequences in E from Example 4.2.1 with the +topology of pointwise convergence, i.e. the topology generated by the seminorms +∣x∣k,α ∶= sup +1≤n≤k +pα(xn), +x = (xn)n∈N ∈ EN, +for k ∈ N and α ∈ A. +5.5.3. Corollary. Let ν ∶= (νn)n∈N be a positive sequence. +a) Then there is 0 < λ ∈ ℓ1 such that for every bounded B ⊂ (KN)′ +b there is +C ≥ 1 with +{µ∣ℓν(N) ∣ µ ∈ B} ⊂ { +∞ +∑ +n=1 +anνnδn ∈ ℓν(N)′ ∣ a ∈ ℓ1, ∀ n ∈ N ∶ ∣an∣ ≤ Cλn}. +b) Then there is a decreasing zero sequence (εn)n∈N such that for all k ∈ N +there is C ≥ 1 with +sup +1≤n≤k +∣xn∣ ≤ C sup +n∈N +∣xn∣νnεn, +x = (xn)n∈N ∈ ℓν(N). +Proof. We take F(N) ∶= KN and F(N,E) ∶= EN as well as Fν(N) ∶= ℓν(N) and +Fν(N,E) ∶= ℓν(N,E) where (T E,T K) ∶= (idEN,idKN) is the generator for (Fν,E). +We remark that F(N) and F(N,E) are ε-compatible and (T E,T K) is a strong, +consistent family for (F,E) by Example 4.2.1 for every Banach space E. Moreover, +Fν(N) = ℓν(N) is a Banach space by [131, Lemma 27.1, p. 326] since ℓν(N) = λ∞(A) +with the Köthe matrix A ∶= (an,j)n,j∈N given by an,j ∶= νn for all n,j ∈ N. +In +addition, we have for every k ∈ N +sup +1≤n≤k +∣xn∣ ≤ sup +1≤n≤k +ν−1 +n ∣x∣ν ≤ sup +1≤n≤k +ν−1 +n , +x = (xn)n∈N ∈ BFν(N), +which means that BFν(N) is bounded in F(N). The space F(N) = KN is a nuclear +Fréchet space and BFν(N) is obviously closed in KN. Thus the bounded and closed +set BFν(N) is compact in F(N), implying our statement by Theorem 5.5.1. +□ +5.5.4. Corollary. Let Ω ⊂ Rd be open, P(∂)K a hypoelliptic linear partial +differential operator, ν∶Ω → (0,∞) continuous and (xn)n∈N fix the topology in +CνP (∂)(Ω). +a) Then there is 0 < λ ∈ ℓ1 such that for every bounded B ⊂ (C∞ +P (∂)(Ω),τc)′ +b +there is C ≥ 1 with +{µ∣CνP (∂)(Ω) ∣ µ ∈ B} ⊂ { +∞ +∑ +n=1 +anν(xn)δxn ∈ CνP (∂)(Ω)′ ∣ a ∈ ℓ1, ∀ n ∈ N ∶ ∣an∣ ≤ Cλn}. +b) Then there is a decreasing zero sequence (εn)n∈N such that for all compact +K ⊂ Ω there is C ≥ 1 with +sup +x∈K +∣f(x)∣ ≤ C sup +n∈N +∣f(xn)∣ν(xn)εn, +f ∈ CνP (∂)(Ω). +Proof. Due to the proof of Corollary 5.2.30 and the observation that the space +F(Ω) = (C∞ +P (∂)(Ω),τc) is a nuclear Fréchet space all conditions of Theorem 5.5.1 +are fulfilled, which yields our statement. +□ +5.6. Series representation of vector-valued functions via Schauder +decompositions +The purpose of this section is to lift series representations known from scalar- +valued functions to vector-valued functions and its underlying idea was derived from +the classical example of the (local) power series representation of a holomorphic +function. We recall that a C-valued function f on the open disc Dr(0) around zero + +118 +5. APPLICATIONS +with radius r > 0 belongs to the space O(Dr(0)) of holomorphic functions on Dr(0) +if the limit +f (1)(z) ∶= +lim +h→0 +h∈C,h≠0 +f(z + h) − f(z) +h +, +z ∈ Dr(0), +(58) +exists in C. It is well-known that every f ∈ O(Dr(0)) can be written as +f(z) = +∞ +∑ +n=0 +f (n)(0) +n! +zn, +z ∈ Dr(0), +where the power series on the right-hand side converges uniformly on every com- +pact subset of Dr(0) and f (n)(0) is the n-th complex derivative of f at 0 which +is defined from (58) by the recursion f (0) ∶= f and f (n) ∶= (f (n−1))(1) for n ∈ N. +By [79, 2.1 Theorem and Definition, p. 17–18] and [79, 5.2 Theorem, p. 35], this +series representation remains valid if f is a holomorphic function on Dr(0) with +values in a locally complete locally convex Hausdorff space E over C where holo- +morphy means that the limit (58) exists in E and the higher complex derivatives +are defined recursively as well. Analysing this example, we observe that O(Dr(0)), +equipped with the topology τc of uniform convergence on compact subsets of Dr(0), +is a Fréchet space, in particular, barrelled, with a Schauder basis formed by the +monomials z ↦ zn. Further, the formulas for the complex derivatives of a C-valued +resp. an E-valued function f on Dr(0) are built up in the same way by (58) (see +Chapter 2). +Our goal is to derive a mechanism which uses these observations and transfers +known series representations for other spaces of scalar-valued functions to their +vector-valued counterparts. Let us describe the general setting. We recall from [89, +14.2, p. 292] that a sequence (fn) in a locally convex Hausdorff space F over a field +K is called a topological basis, or simply a basis, if for every f ∈ F there is a unique +sequence of coefficients (λK +n(f)) in K such that +f = +∞ +∑ +n=1 +λK +n(f)fn +(59) +where the series converges in F. +Due to the uniqueness of the coefficients the +map λK +n∶f ↦ λK +n(f) is well-defined, linear and called the n-th coefficient functional +associated to (fn). Further, for each k ∈ N the map +Pk∶F → F, Pk(f) ∶= +k +∑ +n=1 +λK +n(f)fn, +is a linear projection whose range is span{f1,...,fn} and it is called the k-th ex- +pansion operator associated to (fn). A basis (fn) of F is called equicontinuous if +the expansion operators Pk form an equicontinuous sequence in the linear space +L(F,F) of continuous linear maps from F to F (see [89, 14.3, p. 296]). A basis +(fn) of F is called a Schauder basis if the coefficient functionals are continuous, i.e. +λK +n ∈ F ′ for each n ∈ N. In particular, this is already fulfilled if F is a Fréchet space +by [131, Corollary 28.11, p. 351]. If F is barrelled, then a Schauder basis of F is +already equicontinuous and F has the (bounded) approximation property by the +uniform boundedness principle. +The starting point for our approach is equation (59). Let F and E be non- +trivial locally convex Hausdorff spaces over a field K where F has an equicontinuous +Schauder basis (fn) with associated coefficient functionals (λK +n). The expansion +operators (Pk) form a so-called Schauder decomposition of F (see [27, p. 77]), i.e. +they are continuous projections on F such that +(i) PkPj = Pmin(j,k) for all j,k ∈ N, +(ii) Pk ≠ Pj for k ≠ j, + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +119 +(iii) (Pkf) converges to f for each f ∈ F. +This operator theoretic definition of a Schauder decomposition is equivalent to the +usual definition in terms of closed subspaces of F given in [96, p. 377] (see [123, +p. 219]). In our main Theorem 5.6.1 of this section we prove that (PkεidE) is a +Schauder decomposition of Schwartz’ ε-product FεE and each u ∈ FεE has the +series representation +u(f ′) = +∞ +∑ +n=1 +u(λK +n)f ′(fn), +f ′ ∈ F ′. +Now, suppose that F = F(Ω) is a space of K-valued functions on a set Ω with a +topology such that the point-evaluation functionals δx, x ∈ Ω, belong to F(Ω)′ and +that there is a locally convex Hausdorff space F(Ω,E) of functions from Ω to E +such that the map +S∶F(Ω)εE → F(Ω,E), u �→ [x ↦ u(δx)], +is an isomorphism, i.e. suppose that F(Ω) and F(Ω,E) are ε-compatible. Assuming +that for each n ∈ N and u ∈ F(Ω)εE there is λE +n (S(u)) ∈ E with +λE +n (S(u)) = u(λK +n), +(60) +i.e. (λE,λK) is consistent, we obtain in Corollary 5.6.5 that (S ○ (PkεidE) ○ S−1)k +is a Schauder decomposition of F(Ω,E) and +f = lim +k→∞(S ○ (PkεidE) ○ S−1)(f) = +∞ +∑ +n=1 +λE +n (f)fn, +f ∈ F(Ω,E), +which is the desired series representation in F(Ω,E). In particular, the consis- +tency condition (60) guarantees in the case of E-valued holomorphic functions on +Dr(0) that the complex derivatives at 0 appear in the Schauder decomposition of +O(Dr(0),E) since (∂n +C)ES(u)(0) = u(δ0 ○(∂n +C)C) for all u ∈ O(Dr(0))εE and n ∈ N0 +by Proposition 5.2.32 if E is locally complete. We apply our result to sequence +spaces, spaces of continuously differentiable functions on a compact interval, the +space of holomorphic functions, the Schwartz space and the space of smooth func- +tions which are 2π-periodic in each variable. +As a byproduct of Theorem 5.6.1 we obtain that every element of the completion +F ̂⊗εE of the injective tensor product F ⊗ε E has a series representation as well if +F is a complete space with an equicontinuous Schauder basis and E is complete. +Concerning series representation in F ̂⊗εE, little seems to be known whereas for +the completion F ̂⊗πE of the projective tensor product F ⊗π E of two metrisable +locally convex spaces F and E it is well-known that every f ∈ F ̂⊗πE has a series +representation +f = +∞ +∑ +n=1 +anfn ⊗ en +where (an) ∈ ℓ1, i.e. (an) is absolutely summable, and (fn) and (en) are null +sequences in F and E, respectively (see e.g. [83, Chap. I, §2 , n○1, Théorème 1, p. +51] or [89, 15.6.4 Corollary, p. 334]). If F and E are metrisable and one of them is +nuclear, then the isomorphism F ̂⊗πE ≅ F ̂⊗εE holds and we trivially have a series +representation of the elements of F ̂⊗εE as well. Other conditions on the existence +of series representations of the elements of F ̂⊗εE can be found in [151, Proposition +4.25, p. 88], where F and E are Banach spaces and both of them have a Schauder +basis, and in [91, Theorem 2, p. 283], where F and E are locally convex Hausdorff +spaces and both of them have an equicontinuous Schauder basis. + +120 +5. APPLICATIONS +5.6.1. Schauder decomposition. Let us start with our main theorem on +Schauder decompositions of ε-products. We recall from (3) that we consider the +tensor product F ⊗ E as a linear subspace of FεE for two locally convex Hausdorff +spaces F and E by means of the linear injection +Θ∶F ⊗ E → FεE, +k +∑ +n=1 +fn ⊗ en �→ [y ↦ +k +∑ +n=1 +y(fn)en]. +The next theorem is essentially due to José Bonet, improving a previous version +of us which became Corollary 5.6.5. +5.6.1. Theorem. Let F and E be lcHs, (fn)n∈N an equicontinuous Schauder +basis of F with associated coefficient functionals (λn)n∈N and set Qn∶F → F, +Qn(f) ∶= λn(f)fn for every n ∈ N. Then the following holds: +a) The sequence (Pk)k∈N given by Pk ∶= (∑k +n=1 Qn)εidE is a Schauder decom- +position of FεE. +b) Each u ∈ FεE has the series representation +u(f ′) = +∞ +∑ +n=1 +u(λn)f ′(fn), +f ′ ∈ F ′. +c) F ⊗ E is sequentially dense in FεE. +Proof. Since (fn) is a Schauder basis of F, the sequence (∑k +n=1 Qn) converges +to idF in Lσ(F,F). Thus we deduce from the equicontinuity of (fn) that (∑k +n=1 Qn) +converges to idF in Lκ(F,F) by [89, Theorem 8.5.1 (b), p. 156]. For f ′ ∈ F ′ and +f ∈ F it holds +(Qt +n ○ Qt +m)(f ′)(f) = Qt +m(f ′)(Qn(f)) = Qt +m(f ′)(λn(f)fn) = f ′(λm(λn(f)fn)fm) += λm(fn)λn(f)f ′(fm) = +⎧⎪⎪⎨⎪⎪⎩ +λn(f)f ′(fn) +, m = n, +0 +, m ≠ n, +due to the uniqueness of the coefficient functionals (λn) (see [89, 14.2.1 Proposition, +p. 292]) and it follows for k,j ∈ N that +( +j +∑ +n=1 +Qt +n ○ +k +∑ +m=1 +Qt +m)(f ′)(f) = +min(j,k) +∑ +n=1 +λn(f)f ′(fn) = +min(j,k) +∑ +n=1 +Qt +n(f ′)(f). +This implies that +(PkPj)(u) = u ○ +j +∑ +n=1 +Qt +n ○ +k +∑ +m=1 +Qt +m = u ○ +min(j,k) +∑ +n=1 +Qt +n = Pmin(j,k)(u) +for all u ∈ FεE. If k ≠ j, w.l.o.g. k > j, we choose x ∈ E, x ≠ 0,3 and consider fk ⊗ x +as an element of FεE via the map Θ. Then +(Pk − Pj)(fk ⊗ x) = +k +∑ +n=j+1 +(fk ⊗ x) ○ Qt +n = fk ⊗ x ≠ 0 +since +((fk⊗x)○Qt +n)(f ′) = (fk⊗x)(λn(⋅)f ′(fn)) = λn(fk)f ′(fn)x = +⎧⎪⎪⎨⎪⎪⎩ +(fk ⊗ x)(f ′) , n = k, +0 +, n ≠ k. +It remains to prove that for each u ∈ FεE +lim +k→∞Pk(u) = u +3The lcHs E is non-trivial by our assumptions in Chapter 2. + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +121 +in FεE. Let (qβ)β∈B denote the system of seminorms inducing the locally convex +topology of F. Let u ∈ FεE and α ∈ A. Due to the continuity of u there are an +absolutely convex compact set K = K(u,α) ⊂ F and C0 = C0(u,α) > 0 such that +for each f ′ ∈ F ′ we have +pα((Pk(u) − u)(f ′)) = pα(u(( +k +∑ +n=1 +Qt +n − idF ′)(f ′))) ≤ C0 sup +f∈K +∣( +k +∑ +n=1 +Qt +n − idF ′)(f ′)(f)∣ += C0 sup +f∈K +∣f ′( +k +∑ +n=1 +Qnf − f)∣. +Let V be an absolutely convex zero neighbourhood in F. As a consequence of the +equicontinuity of the polar V ○ there are C1 > 0 and β ∈ B such that +sup +f ′∈V ○ pα((Pk(u) − u)(f ′)) ≤ C0C1 sup +f∈K +qβ( +k +∑ +n=1 +Qnf − f). +In combination with the convergence of (∑k +n=1 Qn) to idF in Lκ(F,F) this yields +the convergence of (Pk(u)) to u in FεE and settles part a). +Let us turn to b) and c). Since +Pk(u)(f ′) = u( +k +∑ +n=1 +Qt +n(f ′)) = +k +∑ +n=1 +u(λn)f ′(fn) +for every f ′ ∈ F ′, we note that the range of Pk(u) is contained in span{u(λn) ∣ 1 ≤ +n ≤ k} for each u ∈ FεE and k ∈ N. Hence Pk(u) has finite rank and thus belongs +to F ⊗ E, implying the sequential density of F ⊗ E in FεE and the desired series +representation by part a). +□ +The index set of the equicontinuous Schauder basis of F in Theorem 5.6.1 need +not be N (or N0) but may be any other countable index set as long as the equicon- +tinuous Schauder basis is unconditional which is, for instance, always fulfilled if F +is nuclear by [89, 21.10.1 Dynin-Mitiagin Theorem, p. 510]. +5.6.2. Remark. If F and E are complete, we have under the assumption of +Theorem 5.6.1 that F ̂⊗εE ≅ FεE by c) since FεE is complete by [94, Satz 10.3, p. +234] and F ̂⊗εE is the closure of F ⊗ E in FεE. Thus each element of F ̂⊗εE has +a series representation. +Let us apply the preceding theorem to spaces of Lebesgue integrable functions. +We consider the measure space ([0,1],L ([0,1]),λ) of Lebesgue measurable sets +and use the notation Lp[0,1] for the space of (equivalence classes) of Lebesgue +p-integrable functions on [0,1]. The Haar system hn∶[0,1] → R, n ∈ N, given by +h1(x) ∶= 1 for all x ∈ [0,1] and +h2k+j(x) ∶= +⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩ +1 +,(2j − 2)/2k+1 ≤ x < (2j − 1)/2k+1, +−1 +,(2j − 1)/2k+1 ≤ x < 2j/2k+1, +0 +,else, +for k ∈ N0 and 1 ≤ j ≤ 2k forms a Schauder basis of Lp[0,1] for every 1 ≤ p < ∞ and +the associated coefficient functionals are +λn(f) ∶= ∫ +[0,1] +f(x)hn(x)dλ(x), +f ∈ Lp[0,1], n ∈ N, +(see [154, Satz I, p. 317]). Because Lp[0,1] is Banach space and thus barrelled, its +Schauder basis (hn) is equicontinuous and we directly obtain from Theorem 5.6.1 +the following corollary. + +122 +5. APPLICATIONS +5.6.3. Corollary. Let E be an lcHs and 1 ≤ p < ∞. (∑k +n=1 λn(⋅)hnεidE)k∈N is +a Schauder decomposition of Lp[0,1]εE and for each u ∈ Lp[0,1]εE it holds +u(f ′) = +∞ +∑ +n=1 +u(λn)f ′(hn), +f ′ ∈ Lp[0,1]′. +Defining Lp([0,1],E) ∶= Lp[0,1]εE, we can read the corollary above as a state- +ment on series representations in the vector-valued version of Lp[0,1]. However, +in many cases of spaces F(Ω) of scalar-valued functions there is a more natural +way to define the vector-valued version F(Ω,E) of F(Ω), namely, that F(Ω) and +F(Ω,E) are ε-compatible. +5.6.4. Remark. If F(Ω) and F(Ω,E) are ε-into-compatible, then we get by +identification of isomorphic subspaces +F(Ω) ⊗ε E ⊂ F(Ω)εE ⊂ F(Ω,E) +and the embedding F(Ω) ⊗ E ↪ F(Ω,E) is given by f ⊗ e �→ [x ↦ f(x)e]. +Proof. The inclusions obviously hold and F(Ω)εE and F(Ω,E) induce the +same topology on F(Ω) ⊗ E. Further, we have +f ⊗ e +Θ +�→ [y ↦ y(f)e] +S +�→ [x �→ [y ↦ y(f)e](δx)] = [x ↦ f(x)e]. +□ +5.6.5. Corollary. Let F(Ω) and F(Ω,E) be ε-compatible, (fn)n∈N an equi- +continuous Schauder basis of F(Ω) with associated coefficient functionals λK ∶= +(λK +n)n∈N. Let there be λE∶F(Ω,E) → EN such that (λE,λK) is a consistent family +for (F,E), and set QE +n ∶F(Ω,E) → F(Ω,E), QE +n (f) ∶= λE +n (f)fn for every n ∈ N. +Then the following holds: +a) The sequence (P E +k )k∈N given by P E +k ∶= ∑k +n=1 QE +n is a Schauder decomposi- +tion of F(Ω,E). +b) Each f ∈ F(Ω,E) has the series representation +f = +∞ +∑ +n=1 +λE +n (f)fn. +c) F(Ω) ⊗ E is sequentially dense in F(Ω,E). +Proof. For each u ∈ F(Ω)εE and x ∈ Ω we note that with Pk from Theorem +5.6.1 it holds +(S ○ Pk)(u)(x) = u( +k +∑ +n=1 +Qt +n(δx)) = u( +k +∑ +n=1 +λK +n(⋅)fn(x)) = +k +∑ +n=1 +u(λK +n)fn(x) += +k +∑ +n=1 +λE +n (S(u))fn(x) = (P E +k ○ S)(u)(x), +which means that S ○ Pk = P E +k ○ S. This implies part a) and b) by Theorem 5.6.1 +a) since S is an isomorphism. Part c) is a direct consequence of Theorem 5.6.1 c) +and the isomorphism F(Ω)εE ≅ F(Ω,E). +□ +In the preceding corollary we used the isomorphism S to obtain a Schauder +decomposition. On the other hand, if S is an isomorphism into, which is often the +case (see Theorem 3.1.12), we can use a Schauder decomposition of F(Ω,E) to +prove the surjectivity of S. +5.6.6. Proposition. Let F(Ω) and F(Ω,E) be ε-into-compatible. Let there be +(fn)n∈N in F(Ω) and for every f ∈ F(Ω,E) a sequence (λE +n (f))n∈N in E such that +f = +∞ +∑ +n=1 +λE +n (f)fn, +f ∈ F(Ω,E). +Then the following holds: + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +123 +a) F(Ω) ⊗ E is sequentially dense in F(Ω,E). +b) If F(Ω) and E are sequentially complete, then +F(Ω,E) ≅ F(Ω)εE. +c) If F(Ω) and E are complete, then +F(Ω,E) ≅ F(Ω)εE ≅ F(Ω)̂⊗εE. +Proof. Let f ∈ F(Ω,E) and observe that +P E +k (f) ∶= +k +∑ +n=1 +λE +n (f)fn = +k +∑ +n=1 +fn ⊗ λE +n (f) ∈ F(Ω) ⊗ E +for every k ∈ N by Remark 5.6.4. Due to our assumption we have the convergence +P E +k (f) → f in F(Ω,E). Thus F(Ω) ⊗ E is sequentially dense in F(Ω,E). +Let us turn to part b). If F(Ω) and E are sequentially complete, then F(Ω)εE +is sequentially complete by [94, Satz 10.3, p. 234]. Since S is an isomorphism into +and +S(Θ( +k +∑ +n=q +fn ⊗ λE +n (f))) = +k +∑ +n=q +λE +n (f)fn +for all k,q ∈ N, k > q, we get that (Θ(∑k +n=1 fn ⊗ λE +n (f)) is a Cauchy sequence in +F(Ω)εE and thus convergent. Hence we deduce that +S( lim +k→∞Θ( +k +∑ +n=1 +fn ⊗ λE +n (f))) = lim +k→∞ +k +∑ +n=1 +(S ○ Θ)(fn ⊗ λE +n (f)) = +∞ +∑ +n=1 +λE +n (f)fn = f, +which proves the surjectivity of S. +If F(Ω) and E are complete, then F(Ω)̂⊗εE is the closure of F(Ω)⊗ε E in the +complete space F(Ω)εE by [94, Satz 10.3, p. 234]. As limk→∞ Θ(∑k +n=1 fn ⊗ λE +n (f)) +is an element of the closure, we obtain part c). +□ +5.6.2. Examples of Schauder decompositions. +Sequence spaces. For our first application we recall the definition of some +sequence spaces. For an lcHs E and a Köthe matrix A ∶= (ak,j)k,j∈N we define the +topological subspace of λ∞(A,E) from Corollary 4.2.3 a) by +c0(A,E) ∶= {x = (xk) ∈ EN ∣ ∀ j ∈ N ∶ lim +k→∞xkak,j = 0}. +In particular, the space c0(N,E) of null-sequences in E is obtained as c0(N,E) = +c0(A,E) with ak,j ∶= 1 for all k,j ∈ N. The space of convergent sequences in E is +defined by +c(N,E) ∶= {x ∈ EN ∣ x = (xk) converges in E} +and equipped with the system of seminorms +∣x∣α ∶= sup +k∈N +pα(xk), +x ∈ c(N,E), +for α ∈ A. Further, we set c0(A) ∶= c0(A,K), c0(N) ∶= c0(N,K) and c(N) ∶= c(N,K). +Furthermore, we equip the space EN with the system of seminorms given by +∥x∥l,α ∶= sup +k∈N +pα(xk)χ{1,...,l}(k), +x = (xk) ∈ EN, +for l ∈ N and α ∈ A. For a non-empty set Ω we define for n ∈ Ω the n-th unit +function by +ϕn,Ω∶Ω → K, ϕn,Ω(k) ∶= +⎧⎪⎪⎨⎪⎪⎩ +1 +, k = n, +0 +, else, +and we simply write ϕn instead of ϕn,Ω if no confusion seems to be likely. Further, +we set ϕ∞∶N → K, ϕ∞(k) ∶= 1, and x∞ ∶= δ∞(x) ∶= limk→∞ xk for x ∈ c(N,E). +For series representations of the elements in these sequence spaces we do not need + +124 +5. APPLICATIONS +Corollary 5.6.5 due to the subsequent proposition but we can use the representation +to obtain the surjectivity of S for sequentially complete E. +5.6.7. Proposition. Let E be an lcHs and ℓ(Ω,E) one of the spaces c0(A,E), +EN, s(Nd,E), s(Nd +0,E) or s(Zd,E). +a) Then (∑n∈Ω,∣n∣≤k δnϕn)k∈N is a Schauder decomposition of ℓ(Ω,E) and +x = ∑ +n∈Ω +xnϕn, +x ∈ ℓ(Ω,E). +b) Then (δ∞ϕ∞+∑k +n=1(δn−δ∞)ϕn)k∈N is a Schauder decomposition of c(N,E) +and +x = x∞ϕ∞ + +∞ +∑ +n=1 +(xn − x∞)ϕn, +x ∈ c(N,E). +Proof. Let us begin with a). First, we note that (ϕn)n∈Ω is an unconditional +equicontinuous Schauder basis of s(Ω), Ω = Nd, Nd +0, Zd, since s(Ω) is a nuclear +Fréchet space. Now, for x = (xn) ∈ ℓ(Ω,E) let (P E +k ) be the sequence in ℓ(Ω,E) +given by P E +k (x) ∶= ∑∣n∣≤k xnϕn. It is easy to see that P E +k is a continuous projection +on ℓ(Ω,E), P E +k P E +j += P E +min(k,j) for all k,j ∈ N and P E +k ≠ P E +j +for k ≠ j. Let ε > 0, +α ∈ A and j ∈ N. For x ∈ c0(A,E) there is N0 ∈ N such that pα(xnan,j) < ε for all +n ≥ N0. Hence we have for x ∈ c0(A,E) +∣x − P E +k (x)∣j,α = sup +n>k +pα(xn)an,j ≤ sup +n≥N0 +pα(xn)an,j ≤ ε +for all k ≥ N0. For x ∈ EN and l ∈ N we have +∥x − P E +k (x)∥l,α = 0 < ε +for all k ≥ l. For x ∈ s(Ω,E), Ω = Nd, Nd +0, Zd, we notice that there is N1 ∈ N such +that for all n ∈ Ω with ∣n∣ ≥ N1 we have +(1 + ∣n∣2)j/2 +(1 + ∣n∣2)j += (1 + ∣n∣2)−j/2 < ε. +Thus we deduce for ∣n∣ ≥ N1 +pα(xn)(1 + ∣n∣2)j/2 < εpα(xn)(1 + ∣n∣2)j ≤ ε∣x∣2j,α +and hence +∣x − P E +k (x)∣j,α = sup +∣n∣>k +pα(xn)(1 + ∣n∣2)j/2 ≤ sup +∣n∣≥N1 +pα(xn)(1 + ∣n∣2)j/2 ≤ ε∣x∣2j,α +for all k ≥ N1. Therefore (P E +k (x)) converges to x in ℓ(Ω,E) and +x = lim +k→∞P E +k (x) = ∑ +n∈Ω +xnϕn. +Now, we turn to b). For x = (xn) ∈ c(N,E) let ( ̃P E +k (x)) be the sequence in +c(N,E) given by ̃P E +k (x) ∶= x∞ϕ∞ + ∑k +n=1(xn − x∞)ϕn. +Again, it is easy to see +that ̃P E +k is a continuous projection on c(N,E), ̃P E +k ̃P E +j += ̃P E +min(k,j) for all k,j ∈ N +and ̃P E +k ≠ ̃P E +j +for k ≠ j. Let ε > 0 and α ∈ A. Then there is N2 ∈ N such that +pα(xn − x∞) < ε for all n ≥ N2. Thus we obtain +∣x − ̃P E +k (x)∣α = sup +n>k +pα(xn − x∞) ≤ sup +n≥N2 +pα(xn − x∞) ≤ ε +for all k ≥ N2, implying that ( ̃P E +k (x)) converges to x in c(N,E) and +x = lim +k→∞ +̃P E +k (x) = x∞ϕ∞ + +∞ +∑ +n=1 +(xn − x∞)ϕn. +□ + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +125 +5.6.8. Theorem. Let E be a sequentially complete lcHs and ℓ(Ω,E) one of the +spaces c0(A,E), EN, s(Nd,E), s(Nd +0,E) or s(Zd,E). Then +(i) ℓ(Ω,E) ≅ ℓ(Ω)εE, +(ii) c(N,E) ≅ c(N)εE. +Proof. The map Sℓ(Ω) is an isomorphism into by Theorem 3.1.12 and, in +addition, by Proposition 4.1.9 (i) if ℓ(Ω,E) = c0(A,E). Considering c(N,E), we +observe that for x ∈ c(N) +δn(x) = xn → x∞ = δ∞(x), +which implies the convergence δn → δ∞ in c(N)′ +γ by the Banach–Steinhaus theorem +since c(N) is a Banach space. Hence we get +u(δ∞) = lim +n→∞u(δn) = lim +n→∞S(u)(n) = δ∞(S(u)) +for every u ∈ c(N)εE, which implies that Sc(N) is an isomorphism into by Theorem +3.1.12. From Proposition 5.6.7 and Proposition 5.6.6 we deduce our statement. +□ +More general, we note that Theorem 5.6.8 holds for any lcHs E if ℓ(Ω,E) = EN +by Example 4.2.1, for E with metric ccp if ℓ(Ω,E) = c0(A,E) by Example 4.2.11 (ii), +and for locally complete E if ℓ(Ω,E) = s(Ω,E) with Ω = Nd, Nd +0, Zd by Corollary +4.2.3 b). +Continuous and differentiable functions on a compact interval. We +start with continuous functions on compact sets. Let E be an lcHs and Ω ⊂ Rd +compact. We equip the space C(Ω,E) of continuous functions on Ω with values in +E with the system of seminorms given by +∣f∣α ∶= sup +x∈Ω +pα(f(x)), +f ∈ C(Ω,E), +for α ∈ A. We want to apply our preceding results to intervals. Let −∞ < a < b < ∞ +and T ∶= (tj)0≤j≤n be a partition of the interval [a,b], i.e. a = t0 < t1 < ... < tn = b. +The hat functions hT +tj∶[a,b] → R for the partition T are given by +hT +tj(x) ∶= +⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩ +x−tj +tj−tj−1 +,tj−1 ≤ x ≤ tj, +tj+1−x +tj+1−tj +,tj < x ≤ tj+1, +0 +,else, +for 2 ≤ j ≤ n − 1 and +hT +a (x) ∶= +⎧⎪⎪⎨⎪⎪⎩ +t1−x +t1−a +,a ≤ x ≤ t1, +0 +,else, +hT +b (x) ∶= +⎧⎪⎪⎨⎪⎪⎩ +x−tn−1 +b−tn−1 +,tn−1 ≤ x ≤ b, +0 +,else. +Let T ∶= (tn)n∈N0 be a dense sequence in [a,b] with t0 = a, t1 = b and tn ≠ tm for n ≠ +m. For T n ∶= {t0,...,tn} there is a (unique) enumeration σ∶{0,...,n} → {0,...,n} +of T n such that Tn ∶= (tσ(j))0≤j≤n is a partition of [a,b] with T n = {tσ(1),...,tσ(n)}. +The functions ϕT +0 ∶= hT1 +t0 , ϕT +1 ∶= hT1 +t1 and ϕT +n ∶= hTn +tσ(j) with j = σ−1(n) for n ≥ 2 +are called Schauder hat functions for the sequence T and form a Schauder basis of +C([a,b]) with associated coefficient functionals given by λK +0 (f) ∶= f(t0), λK +1 (f) ∶= +f(t1) and +λK +n+1(f) ∶= f(tn+1) − +n +∑ +k=0 +λK +k (f)ϕT +k (tn+1), +f ∈ C([a,b]), n ≥ 1, +by [166, 2.3.5 Proposition, p. 29]. Looking at the coefficient functionals, we see that +the right-hand sides even make sense if f ∈ C([a,b],E) and thus we define λE +n on +C([a,b],E) for n ∈ N0 accordingly. + +126 +5. APPLICATIONS +5.6.9. Theorem. Let E be an lcHs with metric ccp and T ∶= (tn)n∈N0 a dense +sequence in [a,b] with t0 = a, t1 = b and tn ≠ tm for n ≠ m. Then (∑k +n=0 λE +k ϕT +n )k∈N0 +is a Schauder decomposition of C([a,b],E) and +f = +∞ +∑ +n=0 +λE +n (f)ϕT +n , +f ∈ C([a,b],E). +Proof. The spaces C([a,b]) and C([a,b],E) are ε-compatible by Example +4.2.12 if E has metric ccp. C([a,b]) is a Banach space and thus barrelled, implying +that its Schauder basis (ϕT +n ) is equicontinuous. We note that for all u ∈ C([a,b])εE +and x ∈ [a,b] +λE +n (S(u))(x) = u(δtn) = u(λK +n), +n ∈ {0,1}, +and by induction +λE +n+1(S(u))(x) = u(δtn+1) − +n +∑ +k=0 +λE +k (S(u))ϕT +k (tn+1) = u(δtn+1) − +n +∑ +k=0 +u(λK +k )ϕT +k (tn+1) += u(λK +n+1), +n ≥ 1. +Thus (λE,λK) is consistent, proving our claim by Corollary 5.6.5. +□ +If a = 0, b = 1 and T is the sequence of dyadic numbers given in [166, 2.1.1 +Definitions, p. 21], then (ϕT +n ) is the so-called Faber–Schauder system. Using the +Schauder basis and coefficient functionals of the space C0(R) of continuous functions +vanishing at infinity given in [166, 2.7.1, p. 41–42] and [166, 2.7.4 Corollary, p. 43] +and that SC0(R) is an isomorphism by Example 4.2.11 (ii) if E has metric ccp, +the corresponding result for the E-valued counterpart C0(R,E) holds as well by a +similar reasoning. Another corresponding result holds for the space C[γ] +0,0 ([0,1],E), +0 < γ < 1, of γ-Hölder continuous functions on [0,1] with values in E that vanish +at zero and at infinity if one uses the Schauder basis and coefficient functionals of +C[γ] +0,0 ([0,1]) from [44, Theorem 2, p. 220] and [43, Theorem 3, p. 230]. This result is +a bit weaker since Example 4.2.9 only guarantees that SC[γ] +0,0([0,1]) is an isomorphism +if E is quasi-complete. +Now, we turn to the spaces Ck([a,b],E) of continuously differentiable functions +on an interval (a,b) with values in an lcHs E such that all derivatives can be +continuously extended to the boundary from Example 4.2.28. We set f (k)(x) ∶= +(∂k)Kf(x) for x ∈ (a,b) and f ∈ Ck([a,b]). From the Schauder hat functions (ϕT +n ) +for a dense sequence T ∶= (tn)n∈N0 in [a,b] with t0 = a, t1 = b and tn ≠ tm for n ≠ m +and the associated coefficient functionals λK +n we can easily get a Schauder basis for +the space Ck([a,b]), k ∈ N, by applying ∫ +(⋅) +a +k-times to the series representation +f (k) = +∞ +∑ +n=0 +λK +n(f (k))ϕT +n , +f ∈ Ck([a,b]), +where we identified f (k) with its continuous extension. The resulting Schauder basis +f T +n ∶[a,b] → R and associated coefficient functionals µK +n∶Ck([a,b]) → K, n ∈ N0, are +f T +n (x) = 1 +n!(x − a)n, +µK +n(f) = f (n)(a), +0 ≤ n ≤ k − 1, +f T +n (x) = +x +∫ +a +sk−1 +∫ +a +⋯ +s2 +∫ +a +s1 +∫ +a +ϕT +n−kdsds1 ...dsk−1, +µK +n(f) = λK +n−k(f (k)), +n ≥ k, +for x ∈ [a,b] and f ∈ Ck([a,b]) (see e.g. [157, p. 586–587], [166, 2.3.7, p. 29]). Again, +the mapping rule for the coefficient functionals still makes sense if f ∈ Ck([a,b],E) +and so we define µE +n on Ck([a,b],E) for n ∈ N0 accordingly. + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +127 +5.6.10. Theorem. Let E be an lcHs with metric ccp, k ∈ N, T ∶= (tn)n∈N0 +a dense sequence in [a,b] with t0 = a, t1 = b and tn ≠ tm for n ≠ m. +Then +(∑l +n=0 µE +n f T +n )l∈N0 is a Schauder decomposition of Ck([a,b],E) and +f = +∞ +∑ +n=0 +µE +n (f)f T +n , +f ∈ Ck([a,b],E). +Proof. The spaces Ck([a,b]) and Ck([a,b],E) are ε-compatible by Example +4.2.28 if E has metric ccp. The Banach space Ck([a,b]) is barrelled giving the +equicontinuity of its Schauder basis. Due to Proposition 3.1.11 c) we have for all +u ∈ Ck([a,b])εE, β ∈ N0, β ≤ k, and x ∈ (a,b) +(∂β)ES(u)(x) = u(δx ○ (∂β)K). +Further, for every sequence (xn) in (a,b) converging to t ∈ {a,b} we obtain by +Proposition 4.1.7 in combination with Lemma 4.1.8 applied to T ∶= (∂β)K +lim +n→∞(∂β)ES(u)(xn) = u( lim +n→∞δxn ○ (∂β)K). +From these observations we deduce that µE +n (S(u)) = u(µK +n) for all n ∈ N0, i.e. +(µE,µK) is consistent. Therefore our statement is a consequence of Corollary 5.6.5. +□ +Holomorphic functions. In this short subsection we show how to get the +result on power series expansion of holomorphic functions from the introduction. +Let E be an lcHs over C, z0 ∈ C, r ∈ (0,∞] and equip O(Dr(z0),E) with the +topology τc of compact convergence. +5.6.11. Theorem. Let E be a locally complete lcHs over C, z0 ∈ C and r ∈ +(0,∞]. Then (f ↦ ∑k +n=0 +(∂n +C )Ef(z0) +n! +(⋅ − z0)n)k∈N0 is a Schauder decomposition of +O(Dr(z0),E) and +f = +∞ +∑ +n=0 +(∂n +C)Ef(z0) +n! +(⋅ − z0)n, +f ∈ O(Dr(z0),E). +Proof. The spaces O(Dr(z0)) and O(Dr(z0),E) are ε-compatible by Propo- +sition 4.2.17 and (23) (cf. [30, Theorem 9, p. 232]) if E is locally complete. Further, +the Schauder basis ((⋅−z0)n) of O(Dr(z0)) is equicontinuous since the Fréchet space +O(Dr(z0)) is barrelled. Due to Proposition 5.2.32 we have for all u ∈ O(Dr(z0))εE +(∂n +C)ES(u)(z) = u(δz ○ (∂n +C)C), +n ∈ N0, z ∈ Dr(z0), +which yields that (λE,λC) is consistent where λE∶O(Dr(z0),E) → EN0 is given +by λE +n (f) ∶= (∂n +C )Ef(z0) +n! +for n ∈ N0 (and analogously for E replaced by C). Hence +Corollary 5.6.5 implies our statement. +□ +Theorem 5.6.11 holds for holomorphic functions in several variables as well (see +[113, Theorem 5.7, p. 264]). +Fourier expansions. In this subsection we turn our attention to Fourier ex- +pansions in the Schwartz space S(Rd,E) and in the space C∞ +2π(Rd,E) of smooth +functions that are 2π-periodic in each variable. +We recall the definition of the Hermite functions. For n ∈ N0 we set +hn∶R → R, hn(x) ∶= (2nn!√π)−1/2(x − d +dx) +n +e−x2/2 = (2nn!√π)−1/2Hn(x)e−x2/2, +with the Hermite polynomials Hn of degree n which can be computed recursively +by +H0(x) = 1, Hn+1(x) = 2xHn(x) − H′ +n(x) and H′ +n(x) = 2nHn−1(x), +x ∈ R, n ∈ N0. + +128 +5. APPLICATIONS +For n = (nk) ∈ Nd +0 we define the n-th Hermite function by +hn∶Rd → R, hn(x) ∶= +d +∏ +k=1 +hnk(xk), +and +Hn∶Rd → R, Hn(x) ∶= +d +∏ +k=1 +Hnk(xk). +5.6.12. Proposition. Let E be a locally complete lcHs, f ∈ S(Rd,E) and n ∈ +Nd +0. Then fhn is Pettis-integrable on Rd. +Proof. First, we set ψ∶Rd → R, ψ(x) ∶= e−∣x∣2/2, as well as g∶Rd → [0,∞), +g(x) ∶= e∣x∣2/2. Then ψ ∈ L1(Rd,λ) and ψg = 1. Moreover, let u∶Rd → E, u(x) ∶= +f(x)hn(x)g(x), and note that +(∂ej)Eu(x) = (∂ej)Ef(x)hn(x)g(x) + f(x)g(x)∂ejhn(x) + f(x)hn(x)g(x)xj +where +∂ejhn(x) = (2njnj!√π)−1/2(H′ +nj(xj)e−x2 +j/2 − Hnj(xj)xje−x2 +j/2) +d +∏ +k=1,k≠j +hnk(xk) += (2njnj!√π)−1/2(2njHnj−1(xj) − xjHnj(xj))e−x2 +j/2 +d +∏ +k=1,k≠j +hnk(xk) +for all x = (xk) ∈ Rd and 1 ≤ j ≤ d. We set Cn ∶= (∏d +i=1 2nini!√π)−1/2 and observe +that +g(x)∂ejhn(x) = e∣x∣2/2∂ejhn(x) = Cn(2njHnj−1(xj) − xjHnj(xj)) +d +∏ +k=1,k≠j +Hnk(xk) +is a polynomial in d variables. The functions given by +hn(x)g(x) = e∣x∣2/2hn(x) = CnHn(x) +and +hn(x)g(x)xj = CnHn(x)xj +are polynomials in d variables as well. Thus there are m ∈ N and C > 0 such that +max(∣hn(x)g(x)∣,∣g(x)∂ejhn(x)∣,∣hn(x)g(x)xj∣) ≤ C(1 + ∣x∣2)m/2 +for all x ∈ Rd and 1 ≤ j ≤ d, which implies +pα((∂ej)Eu(x)) ≤ C(pα((∂ej)Ef(x))(1 + ∣x∣2)m/2 + 2pα(f(x))(1 + ∣x∣2)m/2) +for all α ∈ A and hence +sup +x∈Rd +β∈Nd +0,∣β∣≤1 +pα((∂β)Eu(x)) ≤ 3C∣f∣S(Rd),m,α. +Therefore u = fhng is (weakly) C1 +b , which yields u ∈ C[1] +b (Rd,E) by Proposition +A.1.5. Further, we set h∶Rd → (0,∞), h(x) ∶= 1 + ∣x∣2, and observe that +sup +x∈Rd pα(u(x)h(x)) ≤ sup +x∈Rd pα(f(x))∣hn(x)g(x)h(x)∣ ≤ C∣f∣m+2,α < ∞ +for all α ∈ A. In addition, we remark that for every ε > 0 there is r > 0 such that +1 ≤ εh(x) for all x ∉ Br(0) =∶ K. We deduce from Proposition A.2.7 (iii) that fhn +is Pettis-integrable on Rd. +□ +Due to the previous proposition we can define the n-th Fourier coefficient of +f ∈ S(Rd,E) by +̂f(n) ∶= F E +n (f) ∶= ∫ +Rd +f(x)hn(x)dx = ∫ +Rd +f(x)hn(x)dx, +n ∈ Nd +0, +if E is locally complete. We know that the map +F K∶S(Rd) → s(Nd +0), F K(f) ∶= ( ̂f(n))n∈Nd +0, + +5.6. SERIES REPRESENTATION OF VECTOR-VALUED FUNCTIONS +129 +is an isomorphism (see e.g. [94, Satz 3.7, p. 66]). We improve this result to locally +complete E and derive a Schauder decomposition of S(Rd,E) as well. +5.6.13. Theorem. Let E be a locally complete lcHs. Then the following holds: +a) (∑n∈Nd +0,∣n∣≤k F E +n hn)k∈N is a Schauder decomposition of S(Rd,E) and +f = ∑ +n∈Nd +0 +̂f(n)hn, +f ∈ S(Rd,E). +b) The map +F E∶S(Rd,E) → s(Nd +0,E), F E(f) ∶= ( ̂f(n))n∈Nd +0, +is an isomorphism and +F E = Ss(Nd +0) ○ (F KεidE) ○ S−1 +S(Rd). +Proof. Let us begin with part a). Due to Corollary 3.2.10 the spaces S(Rd) +and S(Rd,E) are ε-compatible and the inverse of the isomorphism S∶S(Rd)εE → +S(Rd,E) is given by the map Rt∶S(Rd,E) → S(Rd)εE, f ↦ J −1 ○ Rt +f, according +to Theorem 3.2.4. Moreover, S(Rd) is a nuclear Fréchet space, thus barrelled, and +hence its Schauder basis (hn) is equicontinuous and unconditional. From the Pettis- +integrability of fhn by Proposition 5.6.12 and Proposition 4.3.3 with (T E +0 ,T K +0 ) ∶= +(hn idERd ,hn idKRd ) we obtain that (F E,F K) is consistent. Hence we conclude our +statement from Corollary 5.6.5. +Let us turn to part b). First, we show that the map F E is well-defined. Let +f ∈ S(Rd,E). Then e′ ○ f ∈ S(Rd) and +⟨e′,F E(f)n⟩ = ⟨e′, ̂f(n)⟩ = ̂ +e′ ○ f(n) = F K(e′ ○ f)n +for every n ∈ Nd +0 and e′ ∈ E′. Thus we have F K(e′ ○ f) ∈ s(Nd +0) for every e′ ∈ E′, +which implies by [131, Mackey’s theorem 23.15, p. 268] that F E(f) ∈ s(Nd +0,E) and +that F E is well-defined. Due to Corollary 3.2.10 and Corollary 4.2.3 the maps +SS(Rd) and Ss(Nd +0) are isomorphisms, which implies that F E is also an isomorphism +with F E = Ss(Nd +0) ○ (F KεidE) ○ S−1 +S(Rd) by Theorem 5.1.2 b). +□ +Our last example of this subsection is devoted to Fourier expansions in the +space C∞ +2π(Rd,E). We recall that C∞ +2π(Rd,E) denotes the topological subspace of +CW∞(Rd,E) consisting of the functions which are 2π-periodic in each variable. Due +to Lemma A.2.2 we are able to define the n-th Fourier coefficient of f ∈ C∞ +2π(Rd,E) +by +̂f(n) ∶= FE +n (f) ∶= (2π)−d +∫ +[−π,π]d +f(x)e−i⟨n,x⟩dx, +n ∈ Zd, +where ⟨⋅,⋅⟩ is the usual scalar product on Rd, if E is locally complete. We know +that the map +FC∶C∞ +2π(Rd) → s(Zd), FC(f) ∶= ( ̂f(n))n∈Zd, +is an isomorphism (see e.g. [94, Satz 1.7, p. 18]), which we lift to the E-valued case. +5.6.14. Theorem. Let E be a locally complete lcHs over C. +a) Then (∑n∈Zd,∣n∣≤k FE +n ei⟨n,⋅⟩)k∈N is a Schauder decomposition of C∞ +2π(Rd,E) +and +f = ∑ +n∈Zd +̂f(n)ei⟨n,⋅⟩, +f ∈ C∞ +2π(Rd,E). + +130 +5. APPLICATIONS +b) The map +FE∶C∞ +2π(Rd,E) → s(Zd,E), FE(f) ∶= ( ̂f(n))n∈Zd, +is an isomorphism and +FE = Ss(Zd) ○ (FCεidE) ○ S−1 +C∞ +2π(Rd). +Proof. The spaces C∞ +2π(Rd) and C∞ +2π(Rd,E) are ε-compatible by Example +4.2.27. +The space C∞ +2π(Rd) is barrelled since it is a nuclear Fréchet space and thus +its Schauder basis (ei⟨n,⋅⟩) is equicontinuous and unconditional. By Theorem 3.2.4 +the inverse of SC∞ +2π(Rd) is given by Rt∶C∞ +2π(Rd,E) → C∞ +2π(Rd)εE, f ↦ J −1 ○ Rt +f. +From the Pettis-integrability of fe−i⟨n,⋅⟩ and Proposition 4.3.3 with (T E +0 ,T K +0 ) ∶= +(e−i⟨n,⋅⟩ idERd ,e−i⟨n,⋅⟩ idCRd ) we obtain that (FE,FC) is consistent. Hence we con- +clude part a) from Corollary 5.6.5. +Let us turn to part b). As in Theorem 5.6.13 it follows from [131, Mackey’s +theorem 23.15, p. 268] that the map FE is well-defined. Due to Corollary 4.2.3 and +Example 4.2.27 the maps Ss(Zd) and SC∞ +2π(Rd) are isomorphisms, which implies that +FE is an isomorphism as well with FE = Ss(Zd) ○ (F CεidE) ○ S−1 +C∞ +2π(Rd) by Theorem +5.1.2 b). +□ +For quasi-complete E Theorem 5.6.14 is already known by [94, Satz 10.8, p. +239]. +5.7. Representation by sequence spaces +Our last section is dedicated to the representation of weighted spaces of E- +valued functions by weighted spaces of E-valued sequences if there is a counterpart +of this representation in the scalar-valued case involving the coefficient functionals +associated to a Schauder basis (see Remark 5.2.3 b)). We only touched upon this +problem in Section 5.6 for special cases like S(Rd,E) and C∞ +2π(Rd,E) in Theorem +5.6.13 b) and Theorem 5.6.14 b). We solve this problem in a different way by an +application of our extension results from Section 5.2. As an example we treat the +space O(DR(0),E) of holomorphic functions and the multiplier space OM(R,E) of +the Schwartz space (see Corollary 5.7.3). +5.7.1. Theorem. Let E be a locally complete lcHs, G ⊂ E′ determine bound- +edness and F(Ω) and F(Ω,E) resp. ℓ(N) and ℓ(N,E) be ε-into-compatible with +e′ ○ g ∈ ℓ(N) for all e′ ∈ E′ and g ∈ ℓ(N,E). Let (fn)n∈N be an equicontinuous +Schauder basis of F(Ω) with associated coefficient functionals (T K +n )n∈N such that +T K∶F(Ω) → ℓ(N), T K(f) ∶= (T K +n (f))n∈N, +is an isomorphism and let there be T E∶F(Ω,E) → EN such that (T E,T K) is a +strong, consistent family for (F,E). If +(i) F(Ω) is a Fréchet–Schwartz space, or +(ii) E is sequentially complete, G = E′ and F(Ω) is a semi-Montel BC-space, +then the following holds: +a) FG(N,E) = ℓ(N,E). +b) ℓ(N) and ℓ(N,E) are ε-compatible, in particular, ℓ(N)εE ≅ ℓ(N,E). +c) The map +T E∶F(Ω,E) → ℓ(N,E), T E(f) ∶= (T E +n (f))n∈N, +is a well-defined isomorphism, F(Ω) and F(Ω,E) are ε-compatible, in +particular, F(Ω)εE ≅ F(Ω,E), and T E = Sℓ(N) ○ (T KεidE) ○ S−1 +F(Ω). + +5.7. REPRESENTATION BY SEQUENCE SPACES +131 +Proof. a)(1) First, we remark that N is a set of uniqueness for (T K,F). Let +u ∈ F(Ω)εE and n ∈ N. Then +RN,G(SF(Ω)(u))(n) = (T E ○ SF(Ω))(u)(n) = T E +n (SF(Ω)(u)) = u(T K +n ) = u(δn ○ T K) += (u ○ (T K)t)(δn) = (T KεidE)(u)(δn) += (Sℓ(N) ○ (T KεidE))(u)(n) +(61) +by consistency and the ε-into-compatibility, yielding FG(N,E) ⊂ ℓ(N,E) once we +have shown that RN,G is surjective, which we postpone to part b). +a)(2) Let g ∈ ℓ(N,E). Then e′ ○g ∈ ℓ(N) for all e′ ∈ E′ and ge′ ∶= (T K)−1(e′ ○g) ∈ +F(Ω). We note that T K +n (ge′) = (e′ ○ g)(n) for all n ∈ N, which implies ℓ(N,E) ⊂ +FG(N,E). +b) We only need to show that Sℓ(N) is surjective. Let g ∈ ℓ(N,E), which implies +g ∈ FG(N,E) by part a)(2). +We claim that RN,G is surjective. In case (i) this follows directly from Theorem +5.2.20. Let us turn to case (ii) and denote by (fn)n∈N the equicontinuous Schauder +basis of F(Ω) associated to (T K +n )n∈N. +We check that condition (ii) of Theorem +5.2.15 is fulfilled. Let f ′ ∈ F(Ω)′ and set +f ′ +k∶F(Ω) → K, f ′ +k(f) ∶= +k +∑ +n=1 +T K +n (f)f ′(fn), +for k ∈ N. Then f ′ +k ∈ F(Ω)′ for every k ∈ N and (f ′ +k) converges to f ′ in F(Ω)′ +σ since +(∑k +n=1 T K +n (f)fn) converges to f in F(Ω). From the equicontinuity of the Schauder +basis we deduce that (f ′ +k) converges to f ′ in F(Ω)′ +κ by [89, 8.5.1 Theorem (b), p. +156]. Let f ∈ FE′(N,E). For each e′ ∈ E′ and k ∈ N we have +Rt +f(f ′ +k)(e′) = f ′ +k(fe′) = +k +∑ +n=1 +T K +n (fe′)f ′(fn) = e′( +k +∑ +n=1 +f(n)f ′(fn)) +since f ∈ FE′(N,E), implying Rt +f(f ′ +k) ∈ J (E). Hence we can apply Theorem 5.2.15 +(ii) and obtain that RN,E′ is surjective, finishing the proof of part a)(1). +Thus there is u ∈ F(Ω)εE such that RN,E′(SF(Ω)(u)) = g in both cases. Then +(T KεidE)(u) ∈ ℓ(N)εE and from (61) we derive +Sℓ(N)((T KεidE)(u)) = RN,G(SF(Ω)(u)) = g, +proving the surjectivity of Sℓ(N). +c) First, we note that the map T E is well-defined. Indeed, we have (e′○T E)(f) = +T K(e′ ○f) ∈ ℓ(N) for all f ∈ F(Ω,E) and e′ ∈ E′ by the strength of the family. Part +a) implies that T E(f) ∈ FG(N,E) = ℓ(N,E) and thus the map T E is well-defined +and its linearity follows from the linearity of the T E +n for n ∈ N. Next, we prove that +T E is surjective. Let g ∈ ℓ(N,E). Since T KεidE is an isomorphism and Sℓ(N) by +part b) as well, we obtain that u ∶= ((T KεidE)−1 ○ S−1 +ℓ(N))(g) ∈ F(Ω)εE. Therefore +SF(Ω)(u) ∈ F(Ω,E) and from (61) we get +T E(SF(Ω)(u)) = (T E ○ SF(Ω))(u) = (Sℓ(N) ○ (T KεidE))(u) = g, +which means that T E is surjective. The injectivity of T E by Proposition 5.2.8, +implies that +SF(Ω) = (T E)−1 ○ (Sℓ(N) ○ (T KεidE)), +yielding the surjectivity of SF(Ω) and thus the ε-compatibility of F(Ω) and F(Ω,E). +Furthermore, we have T E = Sℓ(N) ○ (T KεidE) ○ S−1 +F(Ω), resulting in T E being an +isomorphism. +□ + +132 +5. APPLICATIONS +We note that one should not confuse the coefficient space ℓ(N) of the Schauder +series expansion of functions from F(Ω) in the theorem above with the space +ℓ1 = ℓ1(N) of absolutely summable sequences. +We remark again (see Theorem +5.6.1) that the index set of the equicontinuous Schauder basis of F(Ω) in Theorem +5.7.1 need not be N (or N0) but may be any other countable index set as long as +the equicontinuous Schauder basis is unconditional which is, for instance, always +fulfilled if F(Ω) is nuclear by [89, 21.10.1 Dynin-Mitiagin Theorem, p. 510]. +Theorem 5.7.1 (i) gives another proof of Theorem 5.6.13 b) and Theorem 5.6.14 +b). +Let us demonstrate an application of the preceding theorem which relates +the space of O(DR(0),E), 0 < R ≤ ∞, of holomorphic functions on DR(0) with +values in a complex locally complete lcHs E (see Theorem 5.6.11) and the Köthe +space λ∞(AR,E) with Köthe matrix AR ∶= (rk +j )k∈N0,j∈N for some strictly increasing +sequence (rj)j∈N in (0,R) converging to R (see Corollary 4.2.3), using the sequence +of Taylor coefficients of a holomorphic function. +5.7.2. Corollary. Let E be a locally complete lcHs over C, 0 < R ≤ ∞ and +define the Köthe matrix AR ∶= (rk +j )k∈N0,j∈N for some strictly increasing sequence +(rj)j∈N in (0,R) converging to R. Then λ∞(AR)εE ≅ λ∞(AR,E) and +λE∶O(DR(0),E) → λ∞(AR,E), λE(f) ∶= ((∂k +C)Ef(0) +k! +) +k∈N0, +is an isomorphism with λE = Sλ∞(AR) ○ (λCεidE) ○ S−1 +O(DR(0)). +Proof. By Proposition 4.2.17 and (23) the spaces O(DR(0)) and O(DR(0),E) +are ε-compatible. Moreover, λ∞(AR) and λ∞(AR,E) are ε-compatible by Corollary +4.2.3 as limk→∞( rj +rj+1 )k = 0 for any j ∈ N. Clearly, we have e′ ○ x ∈ λ∞(AR) for all +e′ ∈ E′ and x ∈ λ∞(AR,E). The space O(DR(0)) with the topology τc of compact +convergence is a nuclear Fréchet space and thus a Fréchet–Schwartz space. +In +particular, this space is barrelled and its Schauder basis of monomials (z ↦ zk)k∈N0 +is equicontinuous. The corresponding coefficient functionals are given by λC +k and +the map λC is an isomorphism by [131, Example 27.27, p. 341–342]. By the proof +of Theorem 5.6.11 the family (λE,λC) is consistent for (O,E) and its strength +follows from Proposition 5.2.32. Now, we can apply Theorem 5.7.1 (i), yielding our +statement. +□ +Let us present another application of Theorem 5.7.1 to the space OM(Rd,E) of +multipliers for the Schwartz space from Example 3.1.9 d). For simplicity we restrict +to the case d = 1. Fix a compactly supported test function ϕ ∈ C∞ +c (R) with ϕ(x) = 1 +for x ∈ [0, 1 +4] and ϕ(x) = 0 for x ≥ 1 +2. For f ∈ C∞(R,E) we set +fj(x) ∶= f(x + j) − +∞ +∑ +k=0 +akϕ(−2k(x − 1))f(−2k(x − 1 + j) + 1), x ∈ [0,1], j ∈ Z, +where +ak ∶= +∞ +∏ +j=0,j≠k +1 + 2j +2j − 2k , k ∈ N0. +Fixing x ∈ [0,1), we observe that fj(x) is well-defined for each j ∈ Z since there are +only finitely many summands due to the compact support of ϕ and −2k(x−1) → ∞ +for k → ∞. For x = 1 we have fj(1) = 0 for each j and the convergence of the series +in E follows from the uniform continuity of f on [0,1], f(0) = 0 and ∑∞ +k=0 ak = 1 by +the case n = 0 in [160, Lemma (iii), p. 625]. For each e′ ∈ E′ and j ∈ Z we note that +e′(fj(x)) = (e′ ○f)(x+j)− +∞ +∑ +k=0 +akϕ(−2k(x−1))(e′ ○f)(−2k(x−1+j)+1), x ∈ [0,1], + +5.7. REPRESENTATION BY SEQUENCE SPACES +133 +which implies that e′ ○fj ∈ E0 by [11, Proposition 3.2, p. 15]. Using the weak-strong +principle Corollary 5.2.24, we obtain that fj ∈ E0(E) for all j ∈ Z if E is locally +complete. Setting +ρ∶R → [0,1], ρ(x) ∶= 1 − cos(arctan(x)) = 1 − +1 +√ +1 + x2 , +we deduce from the proof and with the notation of [12, Proposition 2.2, p. 1494] +that e′○fj ○ρ = (Φ−1 +2 ○Φ1)(e′○fj) is an element of the Schwartz space S(R) for each +e′ ∈ E′. The weak-strong principle Corollary 5.2.21 c) yields that fj ○ ρ ∈ S(R,E) +if E is locally complete. Hence (fj ○ ρ) ⋅ h2n is Pettis-integrable on R for every +j ∈ Z and n ∈ N0 by Proposition 5.6.12 if E is locally complete where hn is the n-th +Hermite function. Therefore the Pettis-integral +bn,j(f) ∶= ⟨fj ○ ρ,h2n⟩L2 ∶= ∫ +R +fj(ρ(x))h2n(x)dx, j ∈ Z, n ∈ N0, +is a well-defined element of E by Proposition 5.6.12 if E is locally complete. By +[12, Theorem 2.1, p. 1496–1497] (cf. [172, Theorem 3, p. 478]) the map +ΦK∶OM(R) → s(N)′ +b̂⊗πs(N), ΦK(f) ∶= (bσ(n,j)(f))(n,j)∈N2, +is an isomorphism where σ∶N2 → N0 × Z is the enumeration given by σ(n,j) ∶= +(n − 1,(j − 1)/2) if j is odd, and σ(n,j) ∶= (n − 1,−j/2) if j is even. Here, we have +to interpret ΦK(f) as an element of s(N)′ +b̂⊗πs(N) by identification of isomorphic +spaces. Namely, +s(N)′ +b̂⊗πs(N) ≅ s(N)̂⊗πs(N)′ +b ≅ s(N)εs(N)′ +b ≅ s(N,s(N)′ +b) +holds where the first isomorphism is due to the commutativity of ̂⊗π, the second +due to the nuclearity of s(N) and the last due to Corollary 4.2.3 b) via Ss(N). Then +we interpret ΦK(f) as an element of s(N,s(N)′ +b) by means of +j ∈ N �→ [a ∈ s(N) ↦ ∑ +n∈N +anbσ(n,j)] +(see also (62) below). +5.7.3. Corollary. If E is a sequentially complete lcHs, then the map +ΦE∶OM(R,E) → s(N,Lb(s(N),E)), ΦE(f) ∶= (bσ(n,j)(f))(n,j)∈N2, +is an isomorphism where we interpret ΦE(f) as an element of s(N,Lb(s(N),E)). +Proof. The spaces OM(R) and OM(R,E) are ε-compatible by Corollary +3.2.10 with the inverse of SOM(R) given by the map Rt∶OM(R,E) → OM(R)εE, +f ↦ J −1 ○ Rt +f, according to Theorem 3.2.4. The barrelled nuclear space OM(R) +has the equicontinuous unconditional Schauder basis (ψσ(n,j))(n,j)∈N2 with asso- +ciated coefficient functionals δn,j ○ ΦK = bσ(n,j) given in [12, Proposition 3.2, p. +1499]. Next, we show that (ΦE,ΦK) is a strong, consistent family for (OM,E). Let +f ∈ OM(R,E). For each e′ ∈ E′ and (n,j) ∈ N2 we have +δn,j ○ ΦK(e′ ○ f) = bσ(n,j)(e′ ○ f) = ∫ +R +(e′ ○ f)(j−1)/2(ρ(x))h2(n−1)(x)dx += ⟨e′,∫ +R +f(j−1)/2(ρ(x))h2(n−1)(x)dx⟩ = ⟨e′,δn,j ○ ΦE(f)⟩ += e′(bσ(n,j)(f)) +if j is odd since (f(j−1)/2 ○ ρ) ⋅ h2(n−1) is Pettis-integrable on R. The analogous +result holds for even j as well. This implies the strength of the family. Due to + +134 +5. APPLICATIONS +Proposition 4.3.3 with (T E +0 ,T K +0 ) given by T E +0 (f) ∶= (fj ○ ρ)h2n, f ∈ OM(R,E), and +T K +0 (f) ∶= (fj ○ ρ)h2n, f ∈ OM(R), the family (ΦE,ΦK) is consistent. +In order to apply Theorem 5.7.1 we need spaces ℓV(N2) and ℓV(N2,E) of +sequences with values in K and E, respectively. +In addition, the space ℓV(N2) +has to be isomorphic to s(N,s(N)′ +b) so that ΦK∶OM(R) → s(N,s(N)′ +b) ≅ ℓV(N2) +becomes the isomorphism we need for Theorem 5.7.1. We set +ℓV(N2,E) ∶= {x = (xn,j) ∈ EN2 ∣ ∀ k ∈ N, B ⊂ s(N) bounded, α ∈ A ∶ ∥x∥k,B,α < ∞} +where +∥x∥k,B,α ∶= +sup +(j,a)∈ωB +pα(T E(x)(j,a))νk,B(j,a) +with ωB ∶= N × B and νk,B∶ωB → [0,∞), νk,B(j,a) ∶= (1 + j2)k/2, and +T E(x)(j,a) ∶= ∑ +n∈N +anxn,j. +We claim that the map +T E∶ℓV(N2,E) → s(N,Lb(s(N),E)), x ↦ (T E(x)(j,⋅))j∈N, +(62) +is an isomorphism. We remark for each k ∈ N, bounded B ⊂ s(N) and α ∈ A that +∣T E(x)∣s(N),k,(B,α) = sup +j∈N +sup +a∈B +pα(T E(x)(j,a))(1 + j2)k/2 = ∥x∥k,B,α +for all x ∈ ℓV(N2,E), implying that T E is an isomorphism into. Let y ∶= (yj) ∈ +s(N,Lb(s(N),E)). Then yj ∈ Lb(s(N),E) for j ∈ N and we set xn,j ∶= yj(en) for n ∈ +N where en is the n-th unit sequence in s(N). We note that with x ∶= (xn,j)(n,j)∈N2 +T E(x)(j,a) = ∑ +n∈N +anxn,j = ∑ +n∈N +anyj(en) = yj(∑ +n∈N +anen) = yj(a) +holds for all j ∈ N and a ∶= (an) ∈ s(N) since (en) is a Schauder basis of s(N) with +associated coefficient functionals a ↦ an. It follows that x ∈ ℓV(N2,E) and the +surjectivity of T E. +The next step is to prove that ℓV(N2) and ℓV(N2,E) are ε-into-compatible. Due +to Theorem 3.1.12 we only need to show that (T E,T K) is a consistent generator +for (ℓV,E). Let u ∈ ℓV(N2)εE. Then +m +∑ +n=1 +anSℓV(N2)(u)(j,n) = +m +∑ +n=1 +anu(δj,n) = u( +m +∑ +n=1 +anδj,n) +(63) +for all m ∈ N and a ∶= (an) ∈ s(N). Since +( +m +∑ +n=1 +anδj,n)(x) = +m +∑ +n=1 +anxj,n → T K(x)(j,a) = T K +(j,a)(x), +m → ∞, +for all x ∈ ℓV(N2), we deduce that (∑m +n=1 anδj,n)m converges to T K +(j,a)(x) in ℓV(N2)′ +κ +by the Banach–Steinhaus theorem, which is applicable as ℓV(N2) ≅ s(N,s(N)′ +b) ≅ +OM(R) is barrelled. We conclude that +u(T K +(j,a)) = lim +m→∞u( +m +∑ +n=1 +anδj,n) = +(63) +∞ +∑ +n=1 +anSℓV(N2)(u)(j,n) = T ESℓV(N2)(u)(j,a) +and thus the consistency of (T E,T K) for (ℓV,E). +Furthermore, we clearly have e′ ○ x ∈ ℓV(N2) for all x ∈ ℓV(N2,E) and the map +Φ∶OM(R) → s(N)′ +b̂⊗πs(N) ≅ ℓV(N2) is an isomorphism by [12, Theorem 2.1, p. +1496–1497] and (62). Due to [83, Chap. II, §4, n○4, Théorème 16, p. 131] the dual +OM(R)′ +b is an LF-space and thus OM(R) ≅ (OM(R)′ +b)′ +b is the strong dual of an +LF-space by reflexivity and therefore webbed by [94, Satz 7.25, p. 165]. Finally, we +can apply Theorem 5.7.1 (ii), yielding our statement. +□ + +5.7. REPRESENTATION BY SEQUENCE SPACES +135 +5.7.4. Remark. The actual isomorphism in Corollary 5.7.3 (without the inter- +pretation) is given by ̃ΦE ∶= T E ○ ΦE with T E from (62) and we have +̃ΦE = T E ○ ΦE = T E ○ SℓV(N2) ○ (ΦKεidE) ○ S−1 +OM(R). +Furthermore, Corollary 5.7.3 is valid for locally complete E as well. Indeed, similar +to Example 4.2.2 we may show that ℓV(N2,E) ≅ ℓV(N2)εE for locally complete E. +In combination with Corollary 3.2.10 and Theorem 5.1.2 b) this proves Corollary +5.7.3 for locally complete E as in Theorem 5.6.13 b). + + +Appendices +137 + + +APPENDIX A +Compactness of closed absolutely convex hulls and +Pettis-integrals +A.1. Compactness of closed absolutely convex hulls +In this section of the appendix we treat the question for which functions f∶Ω → +E, subsets K ⊂ Ω and lcHs E sets like acx(f(K)) are compact or sets like +Nj,m(f) ∶= {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm}, +j ∈ J, m ∈ M, +for f ∈ FV(Ω,E) are contained in an absolutely convex compact set. This is useful +in connection with ε-compatibility due to Corollary 3.2.5 (iv) and also relevant +in connection with the Pettis-integrability of a vector-valued function due to the +Mackey–Arens theorem. +We recall that the space of càdlàg functions on a set Ω ⊂ R with values in an +lcHs E is defined by +D(Ω,E) ∶= {f ∈ EΩ ∣ ∀ x ∈ Ω ∶ +lim +w→x+f(w) = f(x) and f(x−) ∶= lim +w→x−f(w) exists}.4 +A.1.1. Proposition. Let Ω ⊂ R, K ⊂ Ω be compact and E an lcHs. Then f(K) +is precompact for every f ∈ D(Ω,E). If E is quasi-complete, then acx(f(K)) is +compact. +Proof. Let f ∈ D(Ω,E), α ∈ A and ε > 0. We recall and define +Br(x) = {w ∈ R ∣ ∣w − x∣ < r} +and +Bε,α(y) ∶= {w ∈ E ∣ pα(w − y) < ε} +for every x ∈ Ω, y ∈ E and r > 0. Let x ∈ Ω. Then there is rx− > 0 such that +pα(f(w) − f(x−)) < ε for all w ∈ Brx−(x) ∩ (−∞,x) ∩ Ω if x is an accumulation +point of (−∞,x] ∩ Ω. Further, there is rx+ > 0 such that pα(f(w) − f(x)) < ε for +all w ∈ Brx+(x) ∩ [x,∞) ∩ Ω if x is an accumulation point of [x,∞) ∩ Ω. If x is an +accumulation point of (−∞,x]∩Ω and [x,∞)∩Ω, we choose rx ∶= min(rx−,rx+). If x +is an accumulation point of (−∞,x]∩Ω but not of [x,∞)∩Ω, we choose rx ∶= rx−. If +x is an accumulation point of [x,∞)∩Ω but not of (−∞,x]∩Ω, we choose rx ∶= rx+. +If x is neither an accumulation point of (−∞,x] ∩ Ω nor of [x,∞) ∩ Ω, then there is +rx > 0 such that Brx(x) ∩ Ω = {x}. +Setting Vx ∶= Brx(x)∩Ω, we note that the sets Vx are open in Ω with respect to +the topology induced by R and K ⊂ ⋃x∈K Vx. Since K is compact, there are n ∈ N +and x1,...,xn ∈ K such that K ⊂ ⋃n +i=1 Vxi. W.l.o.g. each xi is an accumulation point +of (−∞,xi]∩Ω and [xi,∞)∩Ω. Then we have f(w) ∈ (Bε,α(f(xi−))∪Bε,α(f(xi))) +for all w ∈ Vxi and get +f(K) ⊂ +n +⋃ +i=1 +f(Vxi) ⊂ +n +⋃ +i=1 +(Bε,α(f(xi−)) ∪ Bε,α(f(xi))), +which means that f(K) is precompact. +4We recall that for x ∈ Ω we only demand limw→x+ f(w) = f(x) if x is an accumulation +point of [x, ∞) ∩ Ω, and the existence of the limit limw→x− f(w) if x is an accumulation point of +(−∞, x] ∩ Ω. +139 + +140 +A. COMPACTNESS OF CLOSED ABS. CONVEX HULLS & PETTIS-INTEGRALS +If E is quasi-complete, then the precompact set f(K) is relatively compact +by [89, 3.5.3 Proposition, p. 65]. Hence acx(f(K)) is compact as quasi-complete +spaces have ccp. +□ +For f ∈ D(Ω,E) we define the jump function ∆∗f(x) ∶= f(x) − f(x−), x ∈ Ω, +where we set f(x−) ∶= 0 if x is not an accumulation point of (−∞,x] ∩ Ω. +A.1.2. Proposition. Let Ω ⊂ R, K ⊂ Ω be compact and E an lcHs. Then +∆∗f(K) is precompact for every f ∈ D(Ω,E). If E is quasi-complete, then the set +acx(∆∗f(K)) is compact. +Proof. If K is a finite set, then ∆∗f(K) is finite, thus compact, and we are +done. So let us assume that K is not finite. Let α ∈ A and ε > 0. We define +∆ε,α ∶= {x ∈ K ∣ pα(∆∗f(x)) ≥ ε} and claim that ∆α,ε is a finite set. +Let us +assume the contrary. Then there is an infinite sequence (xn) in ∆ε,α ⊂ K. Due +to the compactness of K there is a subsequence of (xn) which converges to some +x ∈ K. W.l.o.g. this subsequence is strictly increasing and we call this subsequence +again (xn). Since f has left limits (in left-accumulation points), for every n ∈ N, +n ≥ 2, there is wn ∈ (xn−1,xn) such that pα(f(xn−) − f(wn)) ≤ ε/2 (if xn is not an +accumulation point of (−∞,xn] ∩ Ω, then there is wn ∈ (xn−1,xn) with wn ∉ Ω and +we set f(wn) ∶= 0). Hence we have +pα(f(xn) − f(wn)) ≥ pα(f(xn) − f(xn−)) − pα(f(xn−) − f(wn)) += pα(∆∗f(xn)) − pα(f(xn−) − f(wn)) ≥ ε/2 +for all n ≥ 2. But this is a contradiction because +lim +n→∞f(xn) = lim +n→∞f(wn) = f(x−), +which proves our claim. +Next, we note that +∆∗f(K) ⊂ (Bε,α(0) ∪ ∆∗f(∆ε,α)) ⊂ +⋃ +z∈{0}∪∆∗f(∆ε,α) +z + Bε,α(0), +which implies that ∆∗f(K) is precompact as {0} ∪ ∆∗f(∆ε,α) is finite. +If E is quasi-complete, then the precompact set ∆∗f(K) is relatively compact +by [89, 3.5.3 Proposition, p. 65]. Hence acx(∆∗f(K)) is compact as quasi-complete +spaces have ccp. +□ +Proposition A.1.1 and Proposition A.1.2 are known in the case that Ω = [0,1] +and E = K (see the comments after [19, Chap. 3, Sect. 14, Lemma 1, p. 110]) since +precompactness is equivalent to boundedness if E = K. +A.1.3. Proposition. Let Ω be a locally compact topological Hausdorff space +and f ∈ C0(Ω,E). If +(i) E is an lcHs with ccp, or +(ii) E is an lcHs with metric ccp and Ω second-countable, +then acx(f(Ω)) is compact. +Proof. Let Ω be compact, then f(Ω) is compact in E as f is continuous. If +Ω is even second-countable, then Ω is metrisable by [58, Chap. XI, 4.1 Theorem, p. +233] and thus f(Ω) as well by [34, Chap. IX, §2.10, Proposition 17, p. 159]. This +yields that acx(f(Ω)) is compact in both cases. +Let Ω be non-compact and Ω∗ denote the one-point compactification of Ω. +Since f ∈ C0(Ω,E), it has a unique continuous extension ̂f to Ω∗ with ̂f(∞) = 0. +Hence K ∶= ̂f(Ω∗) is a compact set in E as Ω∗ is compact and ̂f continuous. If +Ω is even second-countable, then Ω∗ is metrisable by [58, Chap. XI, 8.6 Theorem, + +A.1. COMPACTNESS OF CLOSED ABSOLUTELY CONVEX HULLS +141 +p. 247] and thus K as well by [34, Chap. IX, §2.10, Proposition 17, p. 159]. This +yields that acx(K) is compact in both cases and thus the closed subset acx(f(Ω)), +too. +□ +We note that C0(Ω,E) = C(Ω,E) if Ω is compact. For our next proposition we +define the space of bounded γ-Hölder continuous functions, 0 < γ ≤ 1, from a metric +space (Ω,d) to an lcHs E by +C[γ] +b +(Ω,E) ∶= {f ∈ EΩ ∣ ∀α ∈ A ∶ sup +x∈Ω +pα(f(x)) < ∞ and sup +x,y∈Ω +x≠y +pα(f(x) − f(y)) +d(x,y)γ +< ∞}. +A.1.4. Proposition. Let (Ω,d) be a metric space, E a locally complete lcHs +and f ∈ C[γ] +b +(Ω,E) for some 0 < γ ≤ 1. If there is h∶Ω → (0,∞) such that fh is +bounded on Ω and with N ∶= {x ∈ Ω ∣ f(x) = 0} it holds that +∀ ε > 0 ∃ K ⊂ Ω compact ∀ x ∈ Ω ∖ (K ∪ N) ∶ 1 ≤ εh(x), +then acx(f(Ω)) is compact. +Proof. Since f ∈ C[γ] +b +(Ω,E), the sets f(Ω) and +B1 ∶= {f(z) − f(t) +d(z,t)γ +∣ z,t ∈ Ω, z ≠ t} +are bounded in E. Further, the range (fh)(Ω) is bounded in E by assumption. +Thus B ∶= acx(B1 ∪ f(Ω) ∪ (fh)(Ω)) is a closed disk and EB a Banach space with +the norm ∥x∥B ∶= inf{r > 0 ∣ x ∈ rB}, x ∈ EB, as E is locally complete. Next, +we show that f(Ω) is precompact in EB. Let V be a zero neighbourhood in EB. +Then there is ε > 0 such that Uε ∶= {x ∈ EB ∣ ∥x∥B ≤ ε} ⊂ V . Moreover, there is a +compact set K ⊂ Ω such that 1 ≤ εh(x) for all x ∈ Ω∖(K ∪N). The map f∶Ω → EB +is well-defined and uniformly continuous because ∥f(z) − f(t)∥B ≤ d(z,t)γ for all +z,t ∈ Ω, which follows from B1 ⊂ B. We deduce that f(K) is compact in EB. We +note that +f(x) = f(x)h(x) +1 +h(x), +x ∈ Ω ∖ N, +which implies that ∥f(x)∥B ≤ +1 +h(x) as (fh)(Ω) ⊂ B. Hence we have +∥f(x)∥B ≤ +1 +h(x) ≤ ε, +x ∈ Ω ∖ (K ∪ N), +and the estimate 0 = ∥f(x)∥B ≤ ε is still valid for x ∈ N, yielding f(Ω ∖ K) ⊂ Uε. +Since f(K) is compact in EB, it is also precompact and so there is a finite set +P ⊂ EB such that f(K) ⊂ P + V . We derive that +f(Ω) = (f(K) ∪ f(Ω ∖ K)) ⊂ ((P + V ) ∪ Uε) ⊂ ((P ∪ {0}) + V ), +which means that f(Ω) is precompact in EB and thus acx(f(Ω)) as well by [89, +6.7.1 Proposition, p. 112]. Therefore the set acx(f(Ω)) is compact in the Banach +space EB and also compact in the weaker topology of E. +□ +The underlying idea of Proposition A.1.4 is taken from [29, Lemma 1, Propo- +sition 2, p. 354]. +A.1.5. Proposition. Let Ω ⊂ Rd be an open convex set, E an lcHs over K and +f∶Ω → E weakly C1 +b , i.e. e′ ○ f ∈ C1 +b (Ω) for each e′ ∈ E′. Then f ∈ C[1] +b (Ω,E). +Proof. Let z,t ∈ Ω, z ≠ t. By the mean value theorem we have +∣(e′ ○ f)(z) − (e′ ○ f)(t)∣ +∣z − t∣ +≤ Cd max +1≤n≤dsup +x∈Ω +∣(∂en)K(e′ ○ f)(x)∣ ≤ Cd∣e′ ○ f∣C1 +b (Ω) < ∞ + +142 +A. COMPACTNESS OF CLOSED ABS. CONVEX HULLS & PETTIS-INTEGRALS +for all e′ ∈ E′ where Cd ∶= +√ +d if K = R, and Cd ∶= 2 +√ +d if K = C. It follows from +[131, Mackey’s theorem 23.15, p. 268] that f is Lipschitz continuous and bounded +as well, thus f ∈ C[1] +b (Ω,E). +□ +A.1.6. Proposition. Let FV(Ω,E) be a dom-space, let there be a set X, a +family K of sets and a map π∶⋃m∈M ωm → X such that ⋃K∈K K ⊂ X. +If f ∈ +FV(Ω,E) fulfils +∀ ε > 0, j ∈ J, m ∈ M, α ∈ A ∃ K ∈ K ∶ +(i) +sup +x∈ωm, +π(x)∉K +pα(T E +m(f)(x))νj,m(x) < ε, +(ii) Nπ⊂K,j,m(f) ∶= {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm, π(x) ∈ K} is precompact in E, +then the set Nj,m(f) is precompact in E for every j ∈ J and m ∈ M. +If E is +quasi-complete, then acx(Nj,m(f)) is compact. +Proof. Let V be a zero neighbourhood in E. Then there are α ∈ A and ε > 0 +such that Bε,α ⊂ V where Bε,α ∶= {x ∈ E ∣ pα(x) < ε}. Let j ∈ J and m ∈ M. Due to +(i) there is K ∈ K such that the set +Nπ⊄K,j,m(f) ∶= {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm, π(x) ∉ K} +is contained in Bε,α. Further, the precompactness of Nπ⊂K,j,m(f) by (ii) implies +that there exists a finite set P ⊂ E such that Nπ⊂K,j,m(f) ⊂ P + V . Hence we +conclude +Nj,m(f) = (Nπ⊄K,j,m(f) ∪ Nπ⊂K,j,m(f)) +⊂ (Bε,α ∪ (P + V )) ⊂ (V ∪ (P + V )) = (P ∪ {0}) + V, +which means that Nj,m(f) is precompact. +The second part of the statement follows from the fact that a precompact set +in a quasi-complete space is relatively compact by [89, 3.5.3 Proposition, p. 65] and +that quasi-complete spaces have ccp. +□ +The most common case is that K consists of the compact subsets of Ω and π +is a projection on X ∶= Ω (see e.g. Example 4.2.11, Example 4.2.16 and Example +4.2.22). +A.2. The Pettis-integral +We start with the definition of the Pettis-integral which we use to define Fourier +transformations of vector-valued functions (see Proposition 4.2.25, Theorem 5.6.13 +and Theorem 5.6.14) and for Riesz–Markov–Kakutani theorems in Section 4.3. +Let Σ be a σ-algebra on a set X. +A function µ∶Σ → K is called K-valued +measure if µ(∅) = 0 and µ is countably additive, i.e. for any sequence (An)n∈N of +pairwise disjoint sets in Σ it holds that +µ( ⋃ +n∈N +An) = ∑ +n∈N +µ(An) ∈ K. +If K = R, µ is also called a signed measure, and if K = C a complex measure. If K is +replaced by [0,∞], we say that µ is a positive measure. For a K-valued measure µ +its total variation ∣µ∣ given by +∣µ∣(A) ∶= sup{∑ +n∈N +∣µ(An)∣ ∣ An ∈ Σ,Am ∩ An = ∅ if m ≠ n,A = ⋃ +n∈N +An}, +A ∈ Σ, + +A.2. THE PETTIS-INTEGRAL +143 +is a well-defined positive measure by [149, 6.2 Theorem, p. 117] and it is finite by +[149, 6.4 Theorem, p. 118], i.e. ∣µ∣(X) < ∞. Obviously, a K-valued measure µ is +positive if and only if ∣µ∣ = µ. For a positive measure µ on X and 1 ≤ p < ∞ let +Lp(X,µ) ∶= {f∶X → K measurable ∣ qp(f) ∶= ∫ +X +∣f(x)∣pdµ(x) < ∞} +and define the quotient space of p-integrable functions by Lp(X,µ) ∶= Lp(X,µ)/{f ∈ +Lp(X,µ) ∣ qp(f) = 0}, which becomes a Banach space if it is equipped with the norm +∥f∥p ∶= ∥f∥Lp ∶= qp(F)1/p, f = [F] ∈ Lp(X,µ). From now on we do not distinguish +between equivalence classes and their representatives anymore. +For a K-valued measure µ there is a unique h ∈ L1(X,∣µ∣) with dµ = hd∣µ∣ by +the Radon–Nikodým theorem (see [149, 6.12 Theorem, p. 124]) and h can be chosen +such that ∣h∣ = 1, i.e. has a representative with modulus equal to 1. Now, we say +that f ∈ Lp(X,µ) if f ⋅ h ∈ Lp(X,∣µ∣). For f ∈ L1(X,µ) we define the integral of f +on X w.r.t. µ by +∫ +X +f(x)dµ(x) ∶= ∫ +X +f(x)h(x)d∣µ∣(x). +For a measure space (X,Σ,µ) and f∶X → K we say that f is integrable on +Λ ∈ Σ and write f ∈ L1(Λ,µ) if χΛf ∈ L1(X,µ). Then we set +∫ +Λ +f(x)dµ(x) ∶= ∫ +X +χΛ(x)f(x)dµ(x). +A.2.1. Definition (Pettis-integral). Let (X,Σ,µ) be a measure space and E +an lcHs. A function f∶X → E is called weakly measurable if the function e′ ○f∶X → +K, (e′ ○ f)(x) ∶= ⟨e′,f(x)⟩ ∶= e′(f(x)), is measurable for all e′ ∈ E′. +A weakly +measurable function is said to be weakly integrable if e′ ○ f ∈ L1(X,µ). A function +f∶X → E is called Pettis-integrable on Λ ∈ Σ if it is weakly integrable on Λ and +∃ eΛ ∈ E ∀e′ ∈ E′ ∶ ⟨e′,eΛ⟩ = ∫ +Λ +⟨e′,f(x)⟩dµ(x). +In this case eΛ is unique due to E being Hausdorff and we set the Pettis-integral +∫ +Λ +f(x)dµ(x) ∶= eΛ. +If we consider the measure space (X,L (X),λ) of Lebesgue measurable sets +for X ⊂ Rd, we just write dx ∶= dλ(x). +A.2.2. Lemma. Let E be a locally complete lcHs, Ω ⊂ Rd open and f∶Ω → E. +If f is weakly C1, i.e. e′ ○ f ∈ C1(Ω) for every e′ ∈ E′, then f is Pettis-integrable on +every compact subset K ⊂ Ω with respect to any locally finite positive measure µ on +Ω and +pα(∫ +K +f(x)dµ(x)) ≤ µ(K)sup +x∈K +pα(f(x)), +α ∈ A. +Proof. Let K ⊂ Ω be compact and (Ω,Σ,µ) a measure space with locally +finite measure µ, i.e. Σ contains the Borel σ-algebra B(Ω) on Ω and for every +x ∈ Ω there is a neighbourhood Ux ⊂ Ω of x such that µ(Ux) < ∞. Since the map +e′ ○ f is differentiable for every e′ ∈ E′, thus Borel-measurable, and B(Ω) ⊂ Σ, it is +measurable. We deduce that e′ ○f ∈ L1(K,µ) for every e′ ∈ E′ because locally finite +measures are finite on compact sets. Therefore the map +I∶E′ → K, I(e′) ∶= ∫ +K +⟨e′,f(x)⟩dµ(x) + +144 +A. COMPACTNESS OF CLOSED ABS. CONVEX HULLS & PETTIS-INTEGRALS +is well-defined and linear. We estimate +∣I(e′)∣ ≤ ∣µ(K)∣ sup +x∈f(K) +∣e′(x)∣ ≤ µ(K) +sup +x∈acx(f(K)) +∣e′(x)∣, +e′ ∈ E′. +Due to f being weakly C1 and [29, Proposition 2, p. 354] the absolutely convex set +acx(f(K)) is compact, yielding I ∈ (E′ +κ)′ ≅ E by the theorem of Mackey–Arens, +which means that there is eK ∈ E such that +⟨e′,eK⟩ = I(e′) = ∫ +K +⟨e′,f(x)⟩dµ(x), +e′ ∈ E′. +Hence f is Pettis-integrable on K w.r.t. µ. For α ∈ A we set Bα ∶= {x ∈ E ∣ pα(x) < 1} +and observe that +pα(∫ +K +f(x)dµ(x)) = sup +e′∈B○α +∣⟨e′,∫ +K +f(x)dµ(x)⟩∣ = sup +e′∈B○α +∣∫ +K +e′(f(x))dµ(x)∣ +≤ µ(K) sup +e′∈B○α +sup +x∈K +∣e′(f(x))∣ = µ(K)sup +x∈K +pα(f(x)) +where we used [131, Proposition 22.14, p. 256] in the first and last equation to get +from pα to supe′∈B○α and back. +□ +A.2.3. Lemma. Let E be a sequentially complete lcHs, Ω ⊂ Rd open, (Ω,Σ,µ) +a measure space with locally finite positive measure µ and f∶Ω → E. If f is weakly +C1 and there are ψ ∈ L1(Ω,µ) and g∶Ω → [0,∞) measurable such that ψg ≥ 1 and +fg is bounded on Ω, then f is Pettis-integrable on Ω and +pα(∫ +Ω +f(x)dµ(x)) ≤ ∥ψ∥1 sup +x∈Ω +pα(f(x)g(x)), +α ∈ A. +Proof. Let (Kn)n∈N be a compact exhaustion of Ω. Due to Lemma A.2.2 the +Pettis-integral +en ∶= ∫ +Kn +f(x)dµ(x) +is a well-defined element of E for every n ∈ N. +Next, we show that (en) is a +Cauchy sequence in E. +Let α ∈ A, m ∈ N0 and k,n ∈ N with k > n. +We set +Bα ∶= {x ∈ E ∣ pα(x) < 1} and Qk,n ∶= Kk ∖ Kn and note that +pα(ek − en) = sup +e′∈B○α +∣e′(ek − en)∣ = sup +e′∈B○α +∣ ∫ +Qk,n +e′(f(x))dµ(x)∣ +≤ ∫ +Qk,n +∣ψ(x)∣dµ(x) sup +e′∈B○α +sup +x∈Ω +∣e′(f(x)g(x))∣ += ∫ +Qk,n +∣ψ(x)∣dµ(x)sup +x∈Ω +pα(f(x)g(x)) +(64) +where we used [131, Proposition 22.14, p. 256] to switch from pα to supe′∈B○α and +back. Since ψ ∈ L1(Ω,µ), we have that (en) is a Cauchy sequence in the sequentially +complete space E. Thus eΩ ∶= limn→∞ en exists in E and the dominated convergence +theorem implies +e′(eΩ) = lim +n→∞e′(en) = lim +n→∞∫ +Kn +e′(f(x))dµ(x) = ∫ +Ω +e′(f(x))dµ(x), +e′ ∈ E′. +Hence f is Pettis-integrable on Ω with ∫Ω f(x)dµ(x) = eΩ. As in (64) we have +pα(en) ≤ ∫ +Kn +∣ψ(x)∣dµ(x)sup +x∈Ω +pα(f(x)g(x)) ≤ ∥ψ∥1 sup +x∈Ω +pα(f(x)g(x)) + +A.2. THE PETTIS-INTEGRAL +145 +for every n ∈ N. Letting n → ∞, we derive the estimate in our statement. +□ +A.2.4. Remark. Let µ be a K-valued measure and Σ contain B(Ω). +Then +Lemma A.2.2 is still valid with µ(K) replaced by ∣µ∣(K) due to the definition of +the integral w.r.t. a K-valued measure and as ∣µ∣(K) ≤ ∣µ∣(Ω) < ∞. Thus Lemma +A.2.3 holds in this case as well. +The following definition is analogous to the definition of the Pettis-integral. +A.2.5. Definition (Pettis-summable). Let I be a non-empty set and E an +lcHs. A family (fi)i∈I in E is called weakly summable if (⟨e′,fi⟩)i∈I ∈ ℓ1(I,K) for +all e′ ∈ E′. A family (fi)i∈I in E is called Pettis-summable if it is weakly summable +and +∃ eI ∈ E ∀e′ ∈ E′ ∶ ⟨e′,eI⟩ = ∑ +i∈I +⟨e′,fi⟩. +In this case eI is unique due to E being Hausdorff and we set +∑ +i∈I +fi ∶= eI. +For the elements f of the space D([0,1],E) of E-valued càdlàg functions on +[0,1] and their jump functions ∆∗f we have the following result. +A.2.6. Proposition. Let E be a quasi-complete lcHs, µ a K-valued Borel mea- +sure on [0,1] and ψ ∈ ℓ1([0,1],K). Then f ∈ D([0,1],E) is Pettis-integrable on +[0,1] and +pα( ∫ +[0,1] +f(x)dµ(x)) ≤ ∣µ∣([0,1]) sup +x∈[0,1] +pα(f(x)), +α ∈ A, +and (∆∗f)ψ is Pettis-summable on [0,1] and +pα( ∑ +x∈[0,1] +(∆∗f)(x)ψ(x)) ≤ ∥ψ∥ℓ1 sup +x∈[0,1] +pα(∆∗f(x)), +α ∈ A. +Proof. By [19, Chap. 3, Sect. 14, Lemma 1, p. 110] e′ ○ f is Borel measurable +for every e′ ∈ E′ and integrable due to its boundedness on [0,1]. Thus the map +I∶E′ → K, I(e′) ∶= ∫ +[0,1] +e′(f(x))dµ(x), +is well-defined and linear. It follows from Proposition A.1.1 that acx(f([0,1])) is +absolutely convex and compact in E. In combination with the estimate +∣I(e′)∣ ≤ ∣µ∣([0,1]) +sup +x∈f([0,1]) +∣e′(x)∣ ≤ ∣µ∣([0,1]) +sup +x∈acx(f([0,1])) +∣e′(x)∣ +for every e′ ∈ E′ we deduce that I ∈ (E′ +κ)′ ≅ E by the theorem of Mackey–Arens, +which implies that there is e[0,1] ∈ E such that +⟨e′,e[0,1]⟩ = I(e′) = ∫ +[0,1] +e′(f(x))dµ(x), +e′ ∈ E′. +Thus f is Pettis-integrable on [0,1]. +Since ψ ∈ ℓ1([0,1]) and e′ ○ ∆∗f bounded on [0,1] for every e′ ∈ E′, the map +I0∶E′ → K, I0(e′) ∶= +∑ +x∈[0,1] +e′((∆∗f)(x)ψ(x)), +is well-defined and linear. Moreover, the set acx(∆∗f([0,1])) is absolutely convex +and compact by Proposition A.1.2. Again, the estimate +∣I0(e′)∣ ≤ +∑ +x∈[0,1] +∣ψ(x)∣ +sup +x∈∆∗f([0,1]) +∣e′(x)∣ ≤ ∥ψ∥ℓ1 +sup +x∈acx(∆∗f([0,1])) +∣e′(x)∣ + +146 +A. COMPACTNESS OF CLOSED ABS. CONVEX HULLS & PETTIS-INTEGRALS +for every e′ ∈ E′, implies our statement. +The remaining estimates are deduced +analogously to Lemma A.2.2. +□ +A.2.7. Proposition. Let E be an lcHs, Ω a topological Hausdorff space and +(Ω,Σ,µ) a measure space. +If f∶Ω → E is weakly integrable and there are ψ ∈ +L1(Ω,µ) and g∶Ω → [0,∞) measurable such that ψg ≥ 1 and +(i) E has ccp, Ω is locally compact and fg ∈ C0(Ω,E), or +(ii) E has metric ccp, Ω is locally compact and second-countable, and fg ∈ +C0(Ω,E), or +(iii) E is locally complete, Ω a metric space, fg ∈ C[γ] +b +(Ω,E) for some 0 < γ ≤ 1 +and there is h∶Ω → (0,∞) such that fgh is bounded on Ω and with N ∶= +{x ∈ Ω ∣ f(x)g(x) = 0} it holds that +∀ ε > 0 ∃ K ⊂ Ω compact ∀ x ∈ Ω ∖ (K ∪ N) ∶ 1 ≤ εh(x), +then f is Pettis-integrable on Ω and +pα(∫ +Ω +f(x)dµ(x)) ≤ ∥ψ∥1 sup +x∈Ω +pα(f(x)g(x)), +α ∈ A. +Proof. Since f is weakly integrable, the map +I∶E′ → K, I(e′) ∶= ∫ +Ω +e′(f(x))dµ(x), +is well-defined and linear. It follows from Proposition A.1.3 in case (i)-(ii) and from +Proposition A.1.4 in case (iii) that acx(fg(Ω)) is absolutely convex and compact +in E. If µ is a positive measure, i.e. [0,∞]-valued, we observe that +∣I(e′)∣ ≤ ∫ +Ω +∣ψ(x)∣dµ(x) +sup +x∈fg(Ω) +∣e′(x)∣ ≤ ∥ψ∥1 +sup +x∈acx(fg(Ω)) +∣e′(x)∣ +for every e′ ∈ E′. If µ is a K-valued measure, then the same estimate holds with µ +replaced by ∣µ∣. We deduce from this estimate that I ∈ (E′ +κ)′ ≅ E by the theorem +of Mackey–Arens, which implies that there is eΩ ∈ E such that +⟨e′,eΩ⟩ = I(e′) = ∫ +Ω +e′(f(x))dµ(x), +e′ ∈ E′. +Thus f is Pettis-integrable on Ω. The remaining estimate is deduced analogously +to Lemma A.2.2. +□ +The idea how to prove Proposition A.2.7 (ii) for Ω = Rd is due to an anonymous +reviewer of [110] but did not make it into [110] because of page limits. + +List of Symbols +Sets and systems of sets +acx(M) +absolutely convex hull of the set M 17 +acx(M) +closure of the absolutely convex hull of the set M 17 +B○F (Ω)′ +Fν(Ω) +the polar {y′ ∈ F(Ω)′ ∣ ∀ f ∈ BFν(Ω) ∶ ∣y′(f)∣ ≤ 1} 84 +Br(x) +ball {w ∈ Rd ∣ ∣w − x∣ < r} around x ∈ Rd with radius r > 0 17 +ch(M) +circled hull of the set M 17 +cx(M) +convex hull of the set M 17 +Dr(z) +disc {w ∈ C ∣ ∣w − z∣ < r} around z ∈ C with radius r > 0 17 +D +open unit disc D1(0) 19 +G○ +the polar set of G 17 +Mm +the set {β ∈ Nd +0 ∣ ∣β∣ ≤ min(m,k)} for m ∈ N0 and k ∈ N∞ 24 +M +closure of the set M 17 +M +t +closure of the set M w.r.t. the topology t 17 +M +X +closure of the set M in the topological space X 17 +∂M +boundary of the set M 17 +N∞ +the set N ∪ {∞} 19 +σ(E,G) +weak topology induced on E by a separating subspace G ⊂ E′ 78 +τc +topology of compact convergence 21 +Locally convex Hausdorff spaces & spaces of continuous linear operators +E′⋆ +algebraic dual of the dual E′ 29 +E +locally convex Hausdorff space 17 +ED +space ⋃n∈N nD for a disk D ⊂ E 18 +F ′ +dual space of F 17 +t(F ′,F) +topology on F ′ where t = b, γ, κ, σ or τ 17 +t(F) +bornology on F that induces the topology t(F ′,F) 17 +F ≅ E +locally convex Hausdorff spaces F and E are isomorphic 17 +FεE +space Le(F ′ +κ,E) where L(F ′ +κ,E) is equipped with the topology of +uniform convergence on the equicontinuous subsets of F ′ 17 +F ⊗ E +tensor product of F and E 18 +F ⊗ε E +F ⊗ E equipped with the topology induced by FεE +18 +F ̂⊗εE +completion of F ⊗ε E 18 +F ⊗π E +F ⊗ E equipped with the projective topology 70 +F ̂⊗πE +completion of F ⊗π E 70 +L(F,E) +space of continuous linear operators from F to E 17 +Lt(F,E) +space L(F,E) equipped with the topology t 17 +b +topology t = b on L(F,E) of uniform convergence on +the bounded subsets of F 17 +147 + +148 +List of Symbols +γ +topology t = γ on L(F,E) of uniform convergence on +the precompact (totally bounded) subsets of F 17 +κ +topology t = κ on L(F,E) of uniform convergence on +the absolutely convex, compact subsets of F 17 +σ +topology t = σ on L(F,E) of uniform convergence on +the finite subsets of F 17 +τ +topology t = τ on L(F,E) of uniform convergence on +the absolutely convex, σ(F,F ′)-compact subsets of F +17 +(pα)α∈A +directed system of seminorms inducing the locally convex Hausdorff +topology on E 17 +Spaces of functions +Aτ +∂(C,E) +space of holomorphic functions f∶C → E of exponential type τ 51 +Aτ +∆(Rd,E) +space of harmonic functions f∶Rd → E of exponential type τ 51 +A(Ω,E) +space of continuous functions f∶Ω → E such that f is holomorphic +on Ω 47 +Bν(D,E) +Bloch type space 87 +C(Ω,X) +space of continuous functions f∶Ω → X 17 +C0(Ω,X) +space of continuous functions f∶Ω → X that vanish at infinity 17 +c0(A,E) +space of elements (xk) in the Köthe space λ∞(A,E) such that +(xkak,j) converges to 0 in E for all j ∈ N 123 +c(N,E) +space of convergent sequences in E 123 +Cb(Ω,E) +space of bounded continuous functions f∶Ω → E 46 +C[γ] +b +(Ω,E) +space of bounded γ-Hölder continuous functions f∶Ω → E 140 +C1 +b (Ω,E) +space of continuously partially differentiable functions f∶Ω → E +such that (∂β)Ef is bounded on Ω for all ∣β∣ ≤ 1 20 +C∞ +P (∂)(Ω,E) +kernel of the linear partial differential operator P(∂)E in C∞(Ω,E) +48 +C∞ +P (∂),b(Ω,E) +space of bounded functions f in C∞ +P (∂)(Ω,E) 53 +Cu(Ω,E) +space of uniformly continuous functions f∶Ω → E 39 +Cbu(Ω,E) +space of bounded uniformly continuous functions f∶Ω → E 45 +CC(Ω,E) +space of Cauchy continuous functions f∶Ω → E 38 +Cext(Ω,E) +space of continuous functions f∶Ω → E which have a continuous +extension to Ω 39 +C[γ] +z (Ω,E) +space of γ-Hölder continuous functions f∶Ω → E such that f(z) = 0 +45 +C[γ] +z,0(Ω,E) +space of functions f in C[γ] +z (Ω,E) that vanish at infinity 45 +Ck(Ω,E) +space of k-times continuously partially differentiable functions +f∶Ω → E 19 +C∞ +2π(Rd,E) +space of functions f in C∞(Rd,E) which are 2π-periodic in each +variable 58 +Ck(Ω,E) +space of functions f in Ck(Ω,E) such that all partial derivatives +(∂β)Ef up to order k are continuously extendable on Ω 58 +Ck,γ(Ω,E) +space of functions f in Ck(Ω,E) such that all partial derivatives +(∂β)Ef of order k are γ-Hölder continuous 105 +Ck,γ +loc (Ω,E) +space of functions f in Ck(Ω,E) such that all partial derivatives +(∂β)Ef of order k are locally γ-Hölder continuous 107 +CVk(Ω,E) +space of functions f in Ck(Ω,E) s.t. (β,x) ↦ (∂β)Ef(x)νj,m(β,x) +is bounded on ωm for all j ∈ J and m ∈ N0 24 + +List of Symbols +149 +CWk(Ω,E) +space Ck(Ω,E) equipped with the topology of uniform convergence +of partial derivatives up to order k on compact subsets of Ω 25 +CVk +P (∂)(Ω,E) +kernel of the linear partial differential operator P(∂)E in CVk(Ω,E) +26 +CVk +0(Ω,E) +space of functions f in CVk(Ω,E) that vanish with all their deriva- +tives when weighted at infinity 51 +CVk +0,P (∂)(Ω,E) kernel of the linear partial differential operator P(∂)E in CVk +0(Ω,E) +52 +CV(Ω,E) +space of continuous functions f∶Ω → E such that fν is bounded on +Ω for all ν ∈ V 46 +CV0(Ω,E) +space of functions f in CV(Ω,E) such that fν vanishes at infinity +for all ν ∈ V 46 +CVP (∂)(Ω,E) +kernel of the linear partial differential operator P(∂)E in CV(Ω,E) +48 +CV0,P (∂)(Ω,E) kernel of the linear partial differential operator P(∂)E in CV0(Ω,E) +48 +D(Ω,E) +space of càdlàg functions f∶Ω → E 44 +E0(E) +space of functions f in C∞([0,1],E) such that (∂k)Ef(1) = 0 59 +E{Mp}(Ω,E) +space of ultradifferentiable functions of class {Mp} of Roumieu-type +25 +E(Mp)(Ω,E) +space of ultradifferentiable functions of class (Mp) of Beurling-type +25 +FG(U,E) +space of functions f∶U → E such that for every e′ ∈ G there is +fe′ ∈ F(Ω) with T K +m(fe′)(x) = (e′ ○ f)(m,x) for all (m,x) ∈ U 76 +FVG(U,E)lb +lb-restriction space 92 +FVG(U,E)sb +sb-restriction space 89 +FV(Ω,E)σ +space of functions f∶Ω → E such that e′ ○ f ∈ FV(Ω) for all e′ ∈ E′ +28 +FV(Ω,E)κ +space of functions f in FV(Ω,E)σ such that Rf(B○ +α) is relatively +compact in FV(Ω) for all α ∈ A 29 +FV(Ω,E) +space of functions in F(Ω,E) with a weighted graph topology in- +duced by the family of weights V 22 +AP(Ω,E) +subspace of functions with additional properties in +EΩ 22 +APFV(Ω,E) the space AP(Ω,E) with an emphasis on the depen- +dence on FV(Ω) 22 +∣f∣j,m,α +seminorms applied to f inducing the weighted graph +topology on FV(Ω,E) 22 +∣f∣FV(Ω),j,m,αthe seminorm ∣f∣j,m,α applied to f with an emphasis +on the dependence on FV(Ω) 22 +F(Ω,E) +the intersection AP(Ω,E) ∩ (⋂m∈M domT E +m) 22 +F(Ω) +the space F(Ω,K) 22 +FV(Ω) +the space FV(Ω,K) 22 +Nj,m(f) +the set {T E +m(f)(x)νj,m(x) ∣ x ∈ ωm} 22 +Fν(Ω,E) +the space FV(Ω,E) with V = (ν) 84 +Fεν(Ω,E) +the space of all functions S(u) s.t. u ∈ F(Ω)εE and u(B○F (Ω)′ +Fν(Ω) ) is +bounded in E 84 +H∞(Ω,E) +space of bounded holomorphic functions f∶Ω → E 84 +Lp(X,µ) +space of equivalence classes of p-integrable functions f∶X → K w.r.t. +the measure µ 142 +λ∞(A,E) +Köthe space 42 + +150 +List of Symbols +ℓV(Ω,E) +space of functions f in EΩ such that fν is bounded on Ω for all +ν ∈ V 42 +M(Ω,E) +space of meromorphic functions f∶Ω → E 81 +OM(Rd,E) +multiplier space for the Schwartz space 25 +O(Ω,E) +space of holomorphic functions f∶Ω → E 20 +s(Ω,E) +space of sequences (xk) in E such that the sequence (xk(1+∣k∣2)j/2) +is bounded in E for all j ∈ N 43 +Sµ(Rd,E) +Beurling–Björck space 55 +(γ) +property of µ 55 +S(Rd,E) +Schwartz space 25 +XΩ +space of maps f∶Ω → X 17 +Maps +χK +characteristic function of a set K ⊂ Ω 17 +∂ +Ef +Cauchy–Riemann operator applied to an E-valued function f 20 +∆∗f +the jump function of a càdlàg function f 139 +δx +point evaluation functional f ↦ f(x) 21 +Hn +n-th Hermite polynomial 128 +hn +n-th Hermite function 128 +J +the canonical injection E → E′⋆, x �→ [e′ ↦ e′(x)] 30 +∣µ∣ +total variation of a K-valued measure µ 142 +(∂β)Ef +β-th partial derivative of an E-valued function f 19 +∂βf +β-th partial derivative (∂β)Kf of a K-valued function +f 20 +(∂n +C)Ef +n-th complex derivative (∂n +C)Ef of an E-valued function f 20 +f (n) +n-th complex derivative (∂n +C)Cf of a C-valued func- +tion f 20 +Rf +the map E′ → F(Ω), e′ ↦ fe′, for given f ∈ FG(U,E) 76 +Rf +the map E′ → FV(Ω), e′ ↦ e′ ○ f, for given f ∈ FV(Ω,E)σ 29 +Rt +f +the map FV(Ω)′ → E′⋆, f ′ �→ [e′ ↦ f ′(Rf(e′))], for given f ∈ +FV(Ω,E)σ 29 +S +the map F(Ω)εE → F(Ω,E), u �→ [x ↦ u(δx)] 21 +SF(Ω) +the map S with an emphasis on the dependence on +F(Ω) 21 +T1εT2 +ε-product of the continuous linear operators T1 and T2 18 +T E +m,x +the map f ↦ T E +m(f)(x) 23 +Θ +the linear injection F ⊗ E → FεE 18 +V +family of weight functions 22 +(V∞) +vanishing at infinity condition on the family of weight +functions 34 +Miscellaneous +(DN) +property of a Fréchet space 67 +(Ω) +property of a Fréchet space 67 +(PA) +property of a PLS-space 67 + +Index +almost norming 89 +B-complete 78 +BC-space 78 +Beurling–Björck space 55 +Blaschke’s convergence theorem 110 +Bloch type space 87 +Borel–Ritt theorem 73 +Br-complete 78 +càdlàg function 44 +Cauchy continuous 38 +Cauchy–Riemann operator 20 +coefficient functional 118 +completely regular 37 +consistent family 60 +continuously partially differentiable 19 +convex compactness property (ccp) 18 +metric (metric ccp) 18 +determine boundedness 78 +disk 18 +dom-space 23 +ε-compatible 21 +ε-into-compatible 21 +ε-product 17 +equicontinuous basis 118 +E-valued weak FV-function 28 +expansion operator 118 +exponential type 51 +family of weight functions 22 +C1-controlled 34 +directed 23 +locally bounded 52 +locally bounded away from zero 26, 48 +Favard space 43 +fix the topology 89, 92 +Fourier expansion 127 +Fourier transformation 55, 74 +generalised Gelfand space 109 +generalised Schwartz space 93 +generator 23 +consistent 23 +strong 23 +hat function 125 +Hermite function 128 +Hermite polynomial 127 +Hölder continuous 45 +holomorphic 20 +infra-exponential type 51 +injective tensor product 18 +jump function 139 +Köthe matrix 42 +Köthe space 42 +kR-space 37 +k-space 37 +K-valued measure 142 +lb-restriction space 92 +lcHs 17 +local closure 18 +local limit point 18 +locally closed 18 +locally complete 18 +multiplier space 25 +Nachbin-family 46 +Pettis-integrable 143 +Pettis-integral 143 +Pettis-summable 144 +PLS-space 67 +positive measure 142 +projective tensor product 70 +restriction space 76 +Riesz–Markov–Kakutani theorem 59 +sb-restriction space 89 +Schauder basis 118 +Schauder decomposition 118 +151 + +152 +Index +Schauder hat function 125 +Schwartz space 25 +semiflow 43 +separating subspace 17 +set of uniqueness 76 +strict topology 53, 114 +strong family 60 +topological basis 118 +total variation 142 +ultradifferentiable 25 +uniformly discrete metric space 42 +vanish at infinity in the weighted topol- +ogy w.r.t. (π,K) 40 +weakly integrable 143 +weakly measurable 143 +weakly summable 144 +weak-strong principle 75 +Wolff’s theorem 114 + +Bibliography +[1] A. V. Abanin and L. H. Khoi. Dual of the function algebra A−∞(D) and +representation of functions in Dirichlet series. Proc. Amer. Math. Soc., 138 +(10):3623–3635, 2010. doi:10.1090/S0002-9939-10-10383-9. +[2] A. V. Abanin and V. A. Varziev. +Sufficient sets in weighted Fréchet +spaces of entire functions. +Siberian Math. J., +54(4):575–587, +2013. +doi:10.1134/S0037446613040010. +[3] A. V. Abanin, L. H. Khoi, and Yu. S. Nalbandyan. Minimal absolutely rep- +resenting systems of exponentials for A−∞(Ω). J. Approx. Theory, 163(10): +1534–1545, 2011. doi:10.1016/j.jat.2011.05.011. +[4] H. W. Alt. Lineare Funktionalanalysis. Springer, Berlin, 6th edition, 2012. +doi:10.1007/978-3-642-22261-0. +[5] J. Alvarez, M. S. Eydenberg, and H. Obiedat. The action of operator semi- +groups on the topological dual of the Beurling–Björck space. J. Math. Anal. +Appl., 339(1):405–418, 2008. doi:10.1016/j.jmaa.2007.06.065. +[6] W. Arendt. Vector-valued holomorphic and harmonic functions. Concrete +Operators, 3(1):68–76, 2016. doi:10.1515/conop-2016-0007. +[7] W. Arendt and N. Nikolski. Vector-valued holomorphic functions revisited. +Math. Z., 234(4):777–805, 2000. doi:10.1007/s002090050008. +[8] W. Arendt and N. Nikolski. Addendum: Vector-valued holomorphic functions +revisited. Math. Z., 252(3):687–689, 2006. doi:10.1007/s00209-005-0858-x. +[9] W. Arendt, C. J. K. Batty, M. Hieber, and F. Neubrander. Vector-valued +Laplace transforms and Cauchy problems. +Monogr. Math. 96. Birkhäuser, +Basel, 2nd edition, 2011. doi:10.1007/978-3-0348-0087-7. +[10] D. H. Armitage. Uniqueness theorems for harmonic functions which vanish +at lattice points. J. Approx. Theory, 26(3):259–268, 1979. doi:10.1016/0021- +9045(79)90063-7. +[11] C. +Bargetz. +Commutativity +of +the +Valdivia–Vogt +table +of +repre- +sentations +of +function +spaces. +Math. +Nachr., +287(1):10–22, +2014. +doi:10.1002/mana.201200258. +[12] C. Bargetz. +Explicit representations of spaces of smooth functions +and +distributions. +J. +Math. +Anal. +Appl., +424(2):1491–1505, +2015. +doi:10.1016/j.jmaa.2014.12.009. +[13] G. Beer and S. Levi. Strong uniform continuity. J. Math. Anal. Appl., 350 +(2):568–589, 2009. doi:10.1016/j.jmaa.2008.03.058. +[14] A. Beurling. The collected works of Arne Beurling, Vol. 2 Harmonic analysis. +Contemporary Mathematicians. Birkhäuser, Boston, 1989. +[15] K.-D. Bierstedt. Gewichtete Räume stetiger vektorwertiger Funktionen und +das injektive Tensorprodukt. +PhD thesis, Johannes-Gutenberg Universität +Mainz, Mainz, 1971. urn: urn:nbn:de:hbz:466:2-6756. +[16] K.-D. Bierstedt. Gewichtete Räume stetiger vektorwertiger Funktionen und +das injektive Tensorprodukt. I. J. Reine Angew. Math., 259:186–210, 1973. +doi:10.1515/crll.1973.259.186. +153 + +154 +BIBLIOGRAPHY +[17] K.-D. Bierstedt. Gewichtete Räume stetiger vektorwertiger Funktionen und +das injektive Tensorprodukt. II. J. Reine Angew. Math., 260:133–146, 1973. +doi:10.1515/crll.1973.260.133. +[18] K.-D. Bierstedt and S. Holtmanns. Weak holomorphy and other weak prop- +erties. Bull. Soc. Roy. Sci. Liège, 72(6):377–381, 2003. +[19] P. Billingsley. +Convergence of probability measures. +John Wiley & Sons, +Chichester, 1968. +[20] G. Björck. Linear partial differential operators and generalized distributions. +Ark. Mat., 6(4-5):351–407, 1966. doi:10.1007/BF02590963. +[21] R. P. Boas, Jr. Entire functions. Academic Press, New York, 1954. +[22] R. P. Boas, Jr. A uniqueness theorem for harmonic functions. J. Approx. +Theory, 50(4):425–427, 1972. doi:10.1016/0021-9045(72)90009-3. +[23] V. I. Bogachev. Gaussian measures. Math. Surveys Monogr. 62. AMS, Prov- +idence, RI, 1998. doi:10.1090/surv/062. +[24] V. I. Bogachev and O. G. Smolyanov. +Topological vector spaces and +their applications. +Springer Monogr. Math. Springer, New York, 2017. +doi:10.1007/978-3-319-57117-1. +[25] W. M. Bogdanowicz. Analytic continuation of holomorphic functions with +values in a locally convex space. +Proc. Amer. Math. Soc., 22(3):660–666, +1969. doi:10.2307/2037454. +[26] J. Bonet and P. Domański. The splitting of exact sequences of PLS-spaces +and smooth dependence of solutions of linear partial differential equations. +Adv. Math., 217:561–585, 2008. doi:10.1016/j.aim.2007.07.010. +[27] J. Bonet and W. J. Ricker. Schauder decompositions and the Grothendieck +and Dunford-Pettis properties in Köthe echelon spaces of infinite order. Pos- +itivity, 11(1):77–93, 2007. doi:10.1007/s11117-006-2014-1. +[28] J. Bonet, P. Domański, and M. Lindström. +Weakly compact composition +operators on analytic vector-valued function spaces. Ann. Acad. Sci. Fenn. +Math., 26:233–248, 2001. doi:10.5186/aasfm.00. +[29] J. Bonet, E. Jordá, and M. Maestre. Vector-valued meromorphic functions. +Arch. Math. (Basel), 79(5):353–359, 2002. doi:10.1007/PL00012457. +[30] J. Bonet, L. Frerick, and E. Jordá. +Extension of vector-valued holo- +morphic and harmonic functions. +Studia Math., 183(3):225–248, 2007. +doi:10.4064/sm183-3-2. +[31] J. Bonet, C. Fernández, A. Galbis, and J. M. Ribera. Frames and representing +systems in Fréchet spaces and their duals. Banach J. Math. Anal., 11(1):1–20, +2017. doi:10.1215/17358787-3721183. +[32] F. F. Bonsall. Domination of the supremum of a bounded harmonic function +by its supremum over a countable subset. Proc. Edinb. Math. Soc. (2), 30(3): +471–477, 1987. doi:10.1017/S0013091500026869. +[33] A. Borichev, R. Dhuez, and K. Kellay. +Sampling and interpolation in +large Bergman and Fock spaces. +J. Funct. Anal., 242(2):563–606, 2007. +doi:10.1016/j.jfa.2006.09.002. +[34] N. Bourbaki. General topology, Part 2. Elem. Math. Addison-Wesley, Read- +ing, 1966. +[35] N. Bourbaki. +Integration I. +Elem. Math. Springer, +Berlin, +2004. +doi:10.1007/978-3-642-59312-3. +[36] L. Brown, A. Shields, and K. Zeller. On absolutely convergent exponential +sums. Trans. Amer. Math. Soc., 96(1):162–183, 1960. doi:10.2307/1993491. +[37] H. Buchwalter. Topologies et compactologies. Publ. Dép. Math., Lyon, 6(2): +1–74, 1969. + +BIBLIOGRAPHY +155 +[38] R. B. Burckel. +An introduction to classical complex analysis. +Pure Appl. +Math. (Amst.) 82. Academic Press, New York, 1979. +doi:10.1016/S0079- +8169(08)61287-8. +[39] P. L. Butzer and H. Berens. Semi-groups of operators and approximation. +Grundlehren Math. Wiss. 145. Springer, Berlin, 1967. +[40] R. W. Carroll. Some singular Cauchy problems. Ann. Mat. Pura Appl. (4), +56(1):1–31, 1961. doi:10.1007/BF02414262. +[41] H. Chen. Boundedness from below of composition operators on the Bloch +spaces. Sci. China Ser. A, 46(6):838–846, 2003. doi:10.1360/02ys0212. +[42] H. Chen and P. Gauthier. +Boundedness from below of composition op- +erators on α-Bloch spaces. +Canad. Math. Bull., 51(2):195–204, 2008. +doi:10.4153/CMB-2008-021-2. +[43] Z. Ciesielski. On Haar functions and on the Schauder basis of the space C⟨0,1⟩. +Bull. Acad. Pol. Serie des Sc. Math., Ast. et Ph., 7(4):227–232, 1959. +[44] Z. Ciesielski. On the isomorphisms of the spaces Hα and m. Bull. Acad. Pol. +Serie des Sc. Math., Ast. et Ph., 8(4):217–222, 1960. +[45] J. B. Cooper. The strict topology and spaces with mixed topologies. Proc. +Amer. Math. Soc., 30(3):583–592, 1971. doi:10.2307/2037739. +[46] J. B. Cooper. Saks spaces and applications to functional analysis. North- +Holland Math. Stud. 28. North-Holland, Amsterdam, 1978. +[47] A. Debrouwere and T. Kalmes. Linear topological invariants for kernels of +convolution and differential operators, 2022. arXiv preprint https://arxiv. +org/abs/2204.11733v1. +[48] A. Debrouwere and T. Kalmes. Linear topological invariants for kernels of +differential operators by shifted fundamental solutions, 2023. arXiv preprint +https://arxiv.org/abs/2301.02617v1. +[49] A. Defant and K. Floret. Tensor norms and operator ideals. North-Holland +Math. Stud. 176. North-Holland, Amsterdam, 1993. +[50] F. Deng, L. Jiang, and C. Ouyang. Closed range composition operators on +the Bloch space in the unit ball of Cn. Complex Var. Elliptic Equ., 52(10-11): +1029–1037, 2007. doi:10.1080/17476930701579846. +[51] J. Diestel and J. J. Uhl. Vector measures. Math. Surveys 15. AMS, Provi- +dence, RI, 1977. doi:10.1090/surv/015. +[52] S. Dineen. Complex analysis in locally convex spaces. North-Holland Math. +Stud. 57. North-Holland, Amsterdam, 1981. +[53] P. Domański. +Classical PLS-spaces: +Spaces of distributions, real ana- +lytic functions and their relatives. +In Z. Ciesielski, A. Pełczyński, and +L. Skrzypczak, editors, Orlicz Centenary Volume (Proc., Poznań, 2003), vol- +ume 64 of Banach Center Publ., pages 51–70, Warszawa, 2004. Inst. Math., +Polish Acad. Sci. doi:10.4064/bc64-0-5. +[54] P. Domański and M. Langenbruch. Vector valued hyperfunctions and bound- +ary values of vector valued harmonic and holomorphic functions. Publ. RIMS, +Kyoto Univ., 44(4):1097–1142, 2008. doi:10.2977/prims/1231263781. +[55] P. Domański and M. Lindström. +Sets of interpolation and sampling for +weighted Banach spaces of holomorphic functions. Ann. Polon. Math., 79 +(3):233–264, 2002. doi:10.4064/ap79-3-3. +[56] W. F. Donoghue. Distributions and Fourier transforms. Academic Press, +New York, 1969. +[57] B. K. Driver. Analysis tools with examples, 2004. e-book http://www.math. +ucsd.edu/~bdriver/DRIVER/Book/anal.pdf (January 30, 2023). +[58] J. Dugundji. Topology. Allyn and Bacon Series in Advanced Mathematics. +Allyn and Bacon, Boston, 1966. + +156 +BIBLIOGRAPHY +[59] N. Dunford. Uniformity in linear spaces. Trans. Amer. Math. Soc., 44:305– +356, 1938. doi:10.1090/S0002-9947-1938-1501971-X. +[60] N. Dunford and J. T. Schwartz. Linear operators, Part 1: General theory. +Pure Appl. Math. (N.Y.) 7. Wiley-Intersci., New York, 1958. +[61] L. Ehrenpreis. +Fourier analysis in several complex variables. +Pure Appl. +Math. (N. Y.) 17. Wiley-Interscience, New York, 1970. +[62] K. J. Engel and R. Nagel. +One-parameter semigroups for linear evolu- +tion equations. +Grad. Texts in Math. 194. Springer, New York, 2000. +doi:10.1007/b97696. +[63] R. Engelking. General topology. Sigma Series Pure Math. 6. Heldermann, +Berlin, 1989. +[64] S. Esmaeili, Y. Estaremi, and A. Ebadian. Harmonic Bloch function spaces +and their composition operators. Kragujevac J. Math., 48(4):535–546, 2024. +[65] M. Fabian, P. Habala, P. Hájek, V. Montesinos, and V. Zizler. +Banach +space theory: The basis for linear and nonlinear analysis. CMS Books Math. +Springer, New York, 2011. doi:10.1007/978-1-4419-7515-7. +[66] H. Federer. +Geometric measure theory. +Grundlehren Math. Wiss. 153. +Springer, Berlin, 1969. +[67] J. Fernández, S. Hui, and H. Shapiro. Unimodular functions and uniform +boundedness. Publ. Mat., 33(1):139–146, 1989. doi:10.2307/43737122. +[68] K. Floret and J. Wloka. Einführung in die Theorie der lokalkonvexen Räume. +Lecture Notes in Math. 56. Springer, Berlin, 1968. doi:10.1007/BFb0098549. +[69] L. Frerick and E. Jordá. Extension of vector-valued functions. Bull. Belg. +Math. Soc. Simon Stevin, 14:499–507, 2007. doi:10.36045/bbms/1190994211. +[70] L. Frerick, E. Jordá, and J. Wengenroth. Extension of bounded vector-valued +functions. Math. Nachr., 282(5):690–696, 2009. doi:10.1002/mana.200610764. +[71] D. García, M. Maestre, and P. Rueda. Weighted spaces of holomorphic func- +tions on Banach spaces. Studia Math., 138(1):1–24, 2000. doi:10.4064/sm- +138-1-1-24. +[72] P. Ghatage, J. Yan, and D. Zheng. +Composition operators with closed +range on the Bloch space. Proc. Amer. Math. Soc., 129(7):2039–2044, 2001. +doi:10.2307/2669002. +[73] P. Ghatage, D. Zheng, and N. Zorboska. +Sampling sets and closed range +composition operators on the Bloch space. Proc. Amer. Math. Soc., 133(5): +1371–1377, 2005. doi:10.2307/4097789. +[74] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second +order. Classics Math. Springer, Berlin, 2001. doi:10.1007/978-3-642-61798-0. +[75] J. Giménez, R. Malavé, and J. C. Ramos-Fernández. Composition operators +on µ-Bloch type spaces. Rend. Circ. Mat. Palermo (2), 59(1):107–119, 2010. +doi:10.1007/s12215-010-0007-1. +[76] J. +Globevnik. +Boundaries +for +polydisc +algebras +in +infinite +dimen- +sions. +Math. +Proc. +Cambridge +Philos. +Soc., +85(2):291–303, +1979. +doi:10.1017/S0305004100055705. +[77] B. Gramsch. +Ein Schwach-Stark-Prinzip der Dualitätstheorie lokalkon- +vexer Räume als Fortsetzungsmethode. +Math. Z., 156(3):217–230, 1977. +doi:10.1007/BF01214410. +[78] K.-G. Grosse-Erdmann. Lebesgue’s theorem of differentiation in Fréchet lat- +tices. Proc. Amer. Math. Soc., 112(2):371–379, 1991. doi:10.2307/2048729. +[79] K.-G. Grosse-Erdmann. +The Borel-Okada theorem revisited. +Habilitation. +Fernuniversität Hagen, Hagen, 1992. +[80] K.-G. Grosse-Erdmann. +The locally convex topology on the space +of meromorphic functions. +J. Aust. Math. Soc., 59(3):287–303, 1995. + +BIBLIOGRAPHY +157 +doi:10.1017/S1446788700037198. +[81] K.-G. +Grosse-Erdmann. +A +weak +criterion +for +vector-valued +holo- +morphy. +Math. +Proc. +Camb. +Phil. +Soc., +136(2):399–411, +2004. +doi:10.1017/S0305004103007254. +[82] A. Grothendieck. Sur certains espaces de fonctions holomorphes. I. J. Reine +Angew. Math., 192:35–64, 1953. doi:10.1515/crll.1953.192.35. +[83] A. Grothendieck. +Produits tensoriels topologiques et espaces nucléaires. +Mem. Amer. Math. Soc. 16. AMS, Providence, RI, 4th edition, 1966. +doi:10.1090/memo/0016. +[84] A. Grothendieck. Topological vector spaces. Notes on mathematics and its +applications. Gordon and Breach, New York, 1973. +[85] W. Gustin. A bilinear integral identity for harmonic functions. Amer. J. +Math., 70(1):212–220, 1948. doi:10.2307/2371948. +[86] L. Hörmander. The analysis of linear partial differential operators II. Classics +Math. Springer, Berlin, 2nd edition, 2005. doi:10.1007/b138375. +[87] J. Horváth. +Topological vector spaces and distributions. +Addison-Wesley, +Reading, Mass., 1966. +[88] I. M. James. +Topologies and uniformities. +Springer Undergr. Math. Ser. +Springer, London, 1999. doi:10.1007/978-1-4471-3994-2. +[89] H. Jarchow. +Locally convex spaces. +Math. Leitfäden. Teubner, Stuttgart, +1981. doi:10.1007/978-3-322-90559-8. +[90] A. Jiménez-Vargas. The approximation property for spaces of Lipschitz func- +tions with the bounded weak∗ topology. Rev. Mat. Iberoamericana, 34(2): +637–654, 2018. doi:10.4171/RMI/999. +[91] H. F. Joiner II. +Tensor product of Schauder bases. +Math. Ann., 185(4): +279–284, 1970. doi:710.1007/BF01349950. +[92] E. Jordá. +Extension of vector-valued holomorphic and meromorphic +functions. +Bull. Belg. Math. Soc. Simon Stevin, +12(1):5–21, +2005. +doi:10.36045/bbms/1113318125. +[93] E. Jordá. Weighted vector-valued holomorphic functions on Banach spaces. +Abstract and Applied Analysis, 2013:1–9, 2013. doi:10.1155/2013/501592. +[94] W. Kaballo. Aufbaukurs Funktionalanalysis und Operatortheorie. Springer, +Berlin, 2014. doi:10.1007/978-3-642-37794-5. +[95] S. Kakutani. Concrete representation of abstract M-spaces (A characteriza- +tion of the space of continuous functions). Ann. of Math. (2), 42(4):994–1024, +1941. doi:10.2307/1968778. +[96] N. +J. +Kalton. +Schauder +decompositions +in +locally +convex +spaces. +Math. +Proc. +Cambridge +Philos. +Soc., +68(2):377–392, +1970. +doi:10.1017/S0305004100046193. +[97] K. Knopp. Theory and application of infinite series. Blackie & Son, London +and Glasgow, 2nd edition, 1951. +[98] M. Yu. Kokurin. +Sets of uniqueness for harmonic and analytic functions +and inverse problems for wave equations. Math. Notes, 97(3):376–383, 2015. +doi:10.1134/S0001434615030086. +[99] H. Komatsu. Ultradistributions, I, Structure theorems and a characterization. +J. Fac. Sci. Univ. Tokyo, Sect. IA Math., 20(1):25–105, 1973. +[100] H. Komatsu. Ultradistributions, II, The kernel theorem and ultradistributions +with support in a submanifold. J. Fac. Sci. Univ. Tokyo, Sect. IA Math., 24 +(3):607–628, 1977. doi:10.15083/00039689. +[101] H. Komatsu. Ultradistributions, III, Vector valued ultradistributions and the +theory of kernels. J. Fac. Sci. Univ. Tokyo, Sect. IA Math., 29(3):653–718, +1982. doi:10.15083/00039571. + +158 +BIBLIOGRAPHY +[102] Yu. F. Korobe˘ınik. +Inductive and projective topologies. Sufficient sets +and +representing +systems. +Math. +USSR +Izv., +28(3):529–554, +1987. +doi:10.1070/im1987v028n03abeh000896. +[103] G. Köthe. +Topological vector spaces II. +Grundlehren Math. Wiss. 237. +Springer, Berlin, 1979. doi:10.1007/978-1-4684-9409-9. +[104] A. Kriegl and P. W. Michor. The convenient setting of global analysis. Math. +Surveys Monogr. 53. AMS, Providence, RI, 1997. doi:10.1090/surv/053. +[105] K. Kruse. +Vector-valued Fourier hyperfunctions. +PhD thesis, Universität +Oldenburg, Oldenburg, 2014. urn: urn:nbn:de:gbv:715-oops-19095. +[106] K. Kruse. Weighted vector-valued functions and the ε-product, 2019. arXiv +preprint https://arxiv.org/abs/1712.01613v6 (extended version of [110]). +[107] K. Kruse. The approximation property for weighted spaces of differentiable +functions. In M. Kosek, editor, Function Spaces XII (Proc., Kraków, 2018), +volume 119 of Banach Center Publ., pages 233–258, Warszawa, 2019. Inst. +Math., Polish Acad. Sci. doi:10.4064/bc119-14. +[108] K. Kruse. Parameter dependence of solutions of the Cauchy-Riemann equa- +tion on spaces of weighted smooth functions, 2019. arXiv preprint https: +//arxiv.org/abs/1901.01235v2 (extended version of [112]). +[109] K. Kruse. The Cauchy-Riemann operator on smooth Fréchet-valued func- +tions with exponential growth on rotated strips. PAMM, 19(1):1–2, 2019. +doi:10.1002/pamm.201900141. +[110] K. Kruse. Weighted spaces of vector-valued functions and the ε-product. Ba- +nach J. Math. Anal., 14(4):1509–1531, 2020. doi:10.1007/s43037-020-00072-z. +[111] K. Kruse. On the nuclearity of weighted spaces of smooth functions. Ann. +Polon. Math., 124(2):173–196, 2020. doi:10.4064/ap190728-17-11. +[112] K. Kruse. Parameter dependence of solutions of the Cauchy–Riemann equa- +tion on weighted spaces of smooth functions. RACSAM Rev. R. Acad. Cienc. +Exactas Fís. Nat. Ser. A Mat., 114(3):1–24, 2020. doi:10.1007/s13398-020- +00863-x. +[113] K. Kruse. Vector-valued holomorphic functions in several variables. Funct. +Approx. Comment. Math., 63(2):247–275, 2020. doi:10.7169/facm/1861. +[114] K. Kruse. +Series representations in spaces of vector-valued functions +via +Schauder +decompositions. +Math. +Nachr., +294(2):354–376, +2021. +doi:10.1002/mana.201900172. +[115] K. Kruse. +Extension of weighted vector-valued functions and sequence +space representation, 2021. arXiv preprint https://arxiv.org/abs/1808. +05182v5, to appear in a shortened version in Bull. Belg. Math. Soc. Simon +Stevin, 29(3). doi:10.36045/j.bbms.211009. +[116] K. Kruse. Surjectivity of the ∂-operator between weighted spaces of smooth +vector-valued functions. Complex Var. Elliptic Equ., 67(11):2676–2707, 2022. +doi:10.1080/17476933.2021.1945587. +[117] K. Kruse. Extension of weighted vector-valued functions and weak–strong +principles for differentiable functions of finite order. Annals of Functional +Analysis, 13(1):1–26, 2022. doi:10.1007/s43034-021-00154-5. +[118] K. Kruse. Linearisation of weak vector-valued functions, 2022. arXiv preprint +https://arxiv.org/abs/2207.04681. +[119] K. Kruse. +The inhomogeneous Cauchy-Riemann equation for weighted +smooth vector-valued functions on strips with holes. Collect. Math., 74(1): +81–112, 2023. doi:10.1007/s13348-021-00337-2. +[120] K. Kruse. Extension of weighted vector-valued functions and weak-strong +principles for differentiable functions of finite order, 2023. +arXiv preprint +https://arxiv.org/abs/1910.01952v5 (extended version of [117]). + +BIBLIOGRAPHY +159 +[121] J. Laitila and H.-O. Tylli. Composition operators on vector-valued harmonic +functions and Cauchy transforms. Indiana Univ. Math. J., 55(2):719–746, +2006. doi:10.1512/iumj.2006.55.2785. +[122] G. M. Leibowitz. Lectures on complex function algebras. Scott, Foresman and +Company, Glenview, Ill, 1970. +[123] H. P. Lotz. +Uniform convergence of operators on L∞ and similar spaces. +Math. Z., 190(2):207–220, 1985. doi:10.1007/BF01160459. +[124] Yu. I. Lyubarski˘ı and K. Seip. Sampling and interpolation of entire functions +and exponential systems in convex domains. Ark. Mat., 32(1):157–193, 1994. +doi:10.1007/BF02559527. +[125] P. Mankiewicz. +On the differentiability of Lipschitz mappings in Fréchet +spaces. Studia Math., 45(1):15–29, 1973. doi:10.4064/sm-45-1-15-29. +[126] N. Marco, X. Massaneda, and J. Ortega-Cerdà. Interpolating and sampling +sequences for entire functions. +Geom. Funct. Anal., 13(4):862–914, 2003. +doi:10.1007/s00039-003-0434-7. +[127] A. Markov. On mean values and exterior densities. Rec. Math. [Mat. Sbornik] +N.S., 4(46)(1):165–190, 1938. +[128] X. Massaneda and P. J. Thomas. Sampling sequences for Hardy spaces of the +ball. Proc. Amer. Math. Soc., 128(3):837–843, 2000. doi:10.2307/119748. +[129] R. Meise. Spaces of differentiable functions and the approximation theory. In +J. B. Prolla, editor, Approximation Theory and Functional Analysis (Proc., +Brazil, 1977), North-Holland Math. Stud. 35, pages 263–307, Amsterdam, +1979. North-Holland. +[130] R. Meise and B. A. Taylor. Sequence space representations for (FN)-algebras +of entire functions modulo closed ideals. Studia Math., 85(3):203–227, 1987. +doi:10.4064/sm-85-3-203-227. +[131] R. Meise and D. Vogt. Introduction to functional analysis. Oxf. Grad. Texts +Math. 2. Clarendon Press, Oxford, 1997. +[132] S. +N. +Melikhov. +(DFS)-spaces +of +holomorphic +functions +invariant +under +differentiation. +J. +Math. +Anal. +Appl., +297(2):577–586, +2004. +doi:10.1016/j.jmaa.2004.03.030. +[133] J. R. Munkres. Topology. Prentice Hall, Upper Saddle River, NY, 2nd edition, +2000. +[134] O. Nygaard. Thick sets in Banach spaces and their properties. Quaest. Math., +29(1):59–72, 2006. doi:10.2989/16073600609486149. +[135] J. Ortega-Cerdà and K. Seip. Beurling-type density theorems for weighted +Lp spaces of entire functions. +J. Anal. Math., +75(1):247–266, +1998. +doi:10.1007/BF02788702. +[136] J. Ortega-Cerdà, A. Schuster, and D. Varolin. Interpolation and sampling +hypersurfaces for the Bargmann-Fock space in higher dimensions. +Math. +Ann., 335(1):79–107, 2006. doi:10.1007/s00208-005-0726-3. +[137] V. +P. +Palamodov. +Homological +methods +in +the +theory +of +lo- +cally +convex +spaces. +Russian +Math. +Surveys, +26:1–64, +1971. +doi:10.1070/RM1971v026n01ABEH003815. +[138] P. Pérez Carreras and J. Bonet. Barrelled locally convex spaces. North-Holland +Math. Stud. 131. North-Holland, Amsterdam, 1987. +[139] W. R. Pestman. +Measurability of linear operators in the Skorokhod +topology. +Bull. Belg. Math. Soc. Simon Stevin, +2(4):381–388, +1995. +doi:10.36045/bbms/1103408695. +[140] H.-J. Petzsche. Some results of Mittag–Leffler–type for vector valued func- +tions and spaces of class A. In K.-D. Bierstedt and B. Fuchssteiner, editors, +Functional analysis: Surveys and recent results II (Proc., Paderborn, 1979), + +160 +BIBLIOGRAPHY +volume 38 of North-Holland Math. Stud., pages 183–204, Amsterdam, 1980. +North-Holland. doi:10.1016/s0304-0208(08)70808-9. +[141] M. M. Pirasteh, N. Eghbali, and A. H. Sanatpour. +Closed range proper- +ties of Li–Stević integral-type operators between Bloch-type spaces and their +essential norms. Turkish J. Math., 42(6):3101–3116, 2018. doi:10.3906/mat- +1805-148. +[142] M. H. Powell. On K¯omura’s closed-graph theorem. Trans. Am. Math. Soc., +211:391–426, 1975. doi:10.1090/S0002-9947-1975-0380339-9. +[143] V. Pták. Completeness and the open mapping theorem. Bull. Soc. Math. +France, 86:41–74, 1958. doi:10.24033/bsmf.1498. +[144] J. C. Ramos-Fernández. Composition operators between µ-Bloch spaces. Ex- +tracta Math., 26(1):75–88, 2011. +[145] N. V. Rao. Carlson theorem for harmonic functions in Rn. J. Approx. Theory, +12(4):309–314, 1974. doi:10.1016/0021-9045(74)90074-4. +[146] F. Riesz. Sur les opérations fonctionnelles linéaires. C. R. Acad. Sci. Paris, +149:974–977, 1909. +[147] L. A. Rubel. Bounded convergence of analytic functions. Bull. Amer. Math. +Soc., 77(1):13–24, 1971. doi:10.1090/S0002-9904-1971-12600-9. +[148] L. A. Rubel and A. L. Shields. +The space of bounded analytic func- +tions on a region. +Ann. Inst. Fourier (Grenoble), 16(1):235–277, 1966. +doi:10.5802/aif.231. +[149] W. Rudin. Real and complex analysis. McGraw-Hill, New York, 1970. +[150] W. Ruess. On the locally convex structure of strict topologies. Math. Z., 153 +(2):179–192, 1977. doi:10.1007/BF01179791. +[151] R. A. Ryan. +Introduction to tensor products of Banach spaces. +Springer +Monogr. Math. Springer, Berlin, 2002. doi:10.1007/978-1-4471-3903-4. +[152] S. Saks. Integration in abstract metric spaces. Duke Math. J., 4(2):408–411, +1938. doi:10.1215/S0012-7094-38-00433-8. +[153] H. H. Schaefer. Topological vector spaces. Grad. Texts in Math. Springer, +Berlin, 1971. doi:10.1007/978-1-4684-9928-5. +[154] J. Schauder. Eine Eigenschaft des Haarschen Orthogonalsystems. Math. Z., +28(1), 1928. doi:10.1007/BF01181164. +[155] H.-J. Schmeisser and H. Triebel. +Topics in Fourier analysis and function +spaces. John Wiley & Sons, Chichester, 1987. +[156] D. M. Schneider. Sufficient sets for some spaces of entire functions. Trans. +Amer. Math. Soc., 197:161–180, 1974. doi:10.2307/1996933. +[157] S. Schonefeld. Schauder bases in spaces of differentiable functions. Bull. Amer. +Math. Soc., 75(3):586–590, 1969. doi:10.1090/S0002-9904-1969-12249-4. +[158] L. Schwartz. Espaces de fonctions différentiables à valeurs vectorielles. J. +Analyse Math., 4:88–148, 1955. doi:10.1007/BF02787718. +[159] L. Schwartz. Théorie des distributions à valeurs vectorielles. I. Ann. Inst. +Fourier (Grenoble), 7:1–142, 1957. doi:10.5802/aif.68. +[160] R. T. Seeley. Extension of C∞ functions defined in a half space. Proc. Amer. +Math. Soc., 15(4):625–626, 1964. doi:10.2307/2034761. +[161] K. Seip. Density theorems for sampling and interpolation in the Bargmann- +Fock space. Bull. Amer. Math. Soc., 26(2):322–328, 1992. doi:10.1090/S0273- +0979-1992-00290-2. +[162] K. +Seip. +Density +theorems +for +sampling +and +interpolation +in +the +Bargmann-Fock space I. +J. Reine Angew. Math., +429:91–106, +1992. +doi:10.1515/crll.1992.429.91. +[163] K. Seip. Beurling type density theorems in the unit disk. Invent. Math., 113 +(1):21–39, 1993. doi:10.1007/BF01244300. + +BIBLIOGRAPHY +161 +[164] K. Seip. Developments from nonharmonic Fourier series. Doc. Math., pages +713–722, 1998. +[165] K. Seip and R. Wallstén. Density theorems for sampling and interpolation +in the Bargmann-Fock space II. J. Reine Angew. Math., 429:107–113, 1992. +doi:10.1515/crll.1992.429.107. +[166] Z. Semadeni. Schauder bases in Banach spaces of continuous functions. Lec- +ture Notes in Math. 918. Springer, Berlin, 1982. doi:10.1007/BFb0094629. +[167] N. Sibony. Prolongement des fonctions holomorphes bornées et métrique de +Carathéodory. Invent. Math., 29(3):205–230, 1975. doi:10.1007/BF01389850. +[168] E. M. Stein and R. Shakarchi. Real analysis: Measure theory, integration, +and Hilbert spaces. Princeton Lectures in Analysis III. Princeton University +Press, Princeton, NJ, 2005. +[169] A. H. Stone. Metrisability of unions of spaces. Proc. Amer. Math. Soc., 10 +(3):361–366, 1959. doi:10.2307/2032847. +[170] E. L. Stout. The theory of uniform algebras. Bogden and Quigley, Tarrytown- +on-Hudson, NY, 1971. +[171] F. Trèves. Topological vector spaces, distributions and kernels. Dover, Mine- +ola, NY, 2006. +[172] M. Valdivia. A representation of the space OM. Math. Z., 177(4):463–478, +1981. doi:10.1007/BF01219081. +[173] D. Vogt. On the solvability of P(D)f = g for vector valued functions. In +H. Komatsu, editor, Generalized functions and linear differential equations 8 +(Proc., Kyoto, 1982), volume 508 of RIMS Kôkyûroku, pages 168–181, Kyoto, +1983. RIMS. +[174] D. Vogt. On the functors Ext1(E,F) for Fréchet spaces. Studia Math., 85 +(2):163–197, 1987. doi:10.4064/sm-85-2-163-197. +[175] J. Voigt. On the convex compactness property for the strong operator topol- +ogy. Note Math., XII:259–269, 1992. doi:10.1285/i15900932v12p259. +[176] L. Waelbroeck. Some theorems about bounded structures. J. Funct. Anal., 1 +(4):392–408, 1967. doi:10.1016/0022-1236(67)90009-2. +[177] L. Waelbroeck. Differentiable mappings into b-spaces. J. Funct. Anal., 1(4): +409–418, 1967. doi:10.1016/0022-1236(67)90010-9. +[178] S. Warner. +The topology of compact convergence on continuous function +spaces. +Duke Math. J., 25(2):265–282, 1958. +doi:10.1215/s0012-7094-58- +02523-7. +[179] N. Weaver. Lipschitz algebras. World Sci. Publ., Singapore, 1st edition, 1999. +doi:10.1142/4100. +[180] J. Wengenroth. Derived functors in functional analysis. Lecture Notes in +Math. 1810. Springer, Berlin, 2003. doi:10.1007/b80165. +[181] A. Wilansky. Modern methods in topological vector spaces. McGraw-Hill, New +York, 1978. +[182] E. Wolf. Weighted composition operators on weighted Bergman spaces of +infinite order with the closed range property. Mat. Vesnik, 63(1):33–39, 2011. +[183] M. J. Wolff. Sur les séries ∑ +Ak +z−αk . C. R. Acad. Sci. Paris, 173:1056–1058, +1327–1328, 1921. +[184] R. Yoneda. +Composition operators on the weighted Bloch space and the +weighted Dirichlet spaces, and BMOA with closed range. Complex Var. El- +liptic Equ., 63(5):730–747, 2018. doi:10.1080/17476933.2017.1345887. +[185] D. Zeilberger. Uniqueness theorems for harmonic functions of exponential +growth. Proc. Amer. Math. Soc., 61(2):335–340, 1976. doi:10.1090/S0002- +9939-1976-0425144-6. +