1 Introduction

Let E be a linear space of functions on a set U and \(P(\partial ):{\mathcal {F}}(\varOmega )\rightarrow {\mathcal {F}}(\varOmega )\) be a linear partial differential operator with constant coefficients which acts continuously on a locally convex Hausdorff space of (generalized) differentiable scalar-valued functions \({\mathcal {F}}(\varOmega )\) on an open set \(\varOmega \subset \mathbb {R}^{n}\). We call the elements of U parameters and say that a family \((f_{\lambda })_{\lambda \in U}\) in \({\mathcal {F}}(\varOmega )\) depends on a parameter w.r.t. E if the map \(\lambda \mapsto f_{\lambda }(x)\) is an element of E for every \(x\in \varOmega \). The question of parameter dependence is whether for every family \((f_{\lambda })_{\lambda \in U}\) in \({\mathcal {F}}(\varOmega )\) depending on a parameter w.r.t. E there is a family \((u_{\lambda })_{\lambda \in U}\) in \({\mathcal {F}}(\varOmega )\) with the same kind of parameter dependence which solves the partial differential equation

$$\begin{aligned} P(\partial )u_{\lambda }=f_{\lambda },\quad \lambda \in U. \end{aligned}$$

In particular, it is the question of \({\mathcal {C}}^{k}\)-smooth (holomorphic, distributional, etc.) parameter dependence if E is the space \({\mathcal {C}}^{k}(U)\) of k-times continuously partially differentiable functions on an open set \(U\subset \mathbb {R}^{d}\) (the space \({\mathcal {O}}(U)\) of holomorphic functions on an open set \(U\subset {\mathbb {C}}\), the space of distributions \({\mathcal {D}}(V)'\) on an open set \(V\subset \mathbb {R}^{d}\) where \(U={\mathcal {D}}(V)\), etc.).

The question of parameter dependence has been subject of extensive research varying in the choice of the spaces E, \({\mathcal {F}}(\varOmega )\) and the properties of the partial differential operator \(P(\partial )\), e.g. being (hypo)elliptic, parabolic or hyperbolic. Even partial differential differential operators \(P_{\lambda }(\partial )\) where the coefficients also depend \({\mathcal {C}}^{k}([0,1])\)-smoothly [49], \({\mathcal {C}}^{\infty }\)-smoothly [61], holomorphically [50, 61] or differentiable resp. real analytic [13] on the parameter \(\lambda \) were considered. The case that the coefficients of the partial differential differential operator \(P(x,\partial )\) are non-constant functions in \(x\in \varOmega \) was treated for \({\mathcal {F}}(\varOmega )={\mathscr {A}}(\mathbb {R}^{n})\), the space of real analytic functions on \(\mathbb {R}^{n}\), as well [3].

The answer to the question of \({\mathcal {C}}^{k}\)-smooth (holomorphic, distributional, etc.) parameter dependence is obviously affirmative if \(P(\partial )\) has a linear continuous right inverse. The problem to determine those \(P(\partial )\) which have such a right inverse was posed by Schwartz in the early 1950s (see [21, p. 680]). In the case that \({\mathcal {F}}(\varOmega )\) is the space of \({\mathcal {C}}^{\infty }\)-smooth functions or distributions on an open set \(\varOmega \subset \mathbb {R}^{n}\) the problem was solved in [51, 52] and in the case of ultradifferentiable functions or ultradistributions in [53] by means of Phragmén-Lindelöf type conditions. The case that \({\mathcal {F}}(\varOmega )\) is a weighted space of \({\mathcal {C}}^{\infty }\)-smooth functions on \(\varOmega =\mathbb {R}^{n}\) or its dual was handled in [40], even for some \(P(x,\partial )\) with smooth coefficients, the case of tempered distributions in [38] and of Fourier (ultra-)hyperfunctions in [44, 45]. For Hörmander’s spaces \(B_{p,\kappa }^{loc}(\varOmega )\) as \({\mathcal {F}}(\varOmega )\) the problem was studied in [25].

The necessary condition of surjectivity of the partial differential operator \(P(\partial )\) was studied in many papers, e.g. in [1, 23, 28, 48, 67] on \({\mathcal {C}}^{\infty }\)-smooth functions and distributions, in [9, 26, 43] on real analytic functions, in [8, 14] on Gevrey classes, in [10, 12, 41, 42, 55] on ultradifferentiable functions of Roumieu type, in [22] on ultradistributions of Beurling type, in [7, 11] on ultradifferentiable functions and ultradistributions and in [47] on the multiplier space \({\mathcal {O}}_{M}\).

However, if \(P(\partial ):{\mathcal {C}}^{\infty }(\varOmega )\rightarrow {\mathcal {C}}^{\infty }(\varOmega )\), \(\varOmega \subset \mathbb {R}^{n}\) open, is elliptic, then \(P(\partial )\) has a linear right inverse (by means of a Hamel basis of \({\mathcal {C}}^{\infty }(\varOmega )\)) and it has a continuous right inverse due to Michael’s selection theorem [56, Theorem 3.2”, p. 367] and [29, Satz 9.28, p. 217], but \(P(\partial )\) has no linear continuous right inverse if \(n\ge 2\) by a result of Grothendieck [62, Theorem C.1, p. 109]. Nevertheless, the question of parameter dependence w.r.t. E has a positive answer for several locally convex Hausdorff spaces E due to tensor product techniques. In this case the question of parameter dependence obviously has a positive answer if the topology of E is stronger than the topology of pointwise convergence on U and

$$\begin{aligned} P(\partial )^{E}:{\mathcal {C}}^{\infty }(\varOmega ,E) \rightarrow {\mathcal {C}}^{\infty }(\varOmega ,E) \end{aligned}$$

is surjective where \({\mathcal {C}}^{\infty }(\varOmega ,E)\) is the space of \({\mathcal {C}}^{\infty }\)-smooth E-valued functions on \(\varOmega \) and \(P(\partial )^{E}\) the version of \(P(\partial )\) for E-valued functions. From Grothendieck’s classical theory of tensor products [24] and the surjectivity of \(P(\partial )\) it follows that \(P(\partial )^{E}\) is also surjective if E is a Fréchet space. In general, Grothendieck’s theory of tensor products can be applied if \(P(\partial )\) is surjective and \({\mathcal {F}}(\varOmega )\) a nuclear Fréchet space. However, the surjectivity of \(P(\partial )^{E}\), \(n\ge 2\), can even be extended beyond the class of Fréchet spaces E due to the splitting theory of Vogt for Fréchet spaces [64, 65] and of Bonet and Domański for PLS-spaces [4, 6] if, in addition, \({\text {ker}}P(\partial )\) has the property \((\varOmega )\) and E is the dual of a Fréchet space with the property (DN) or an ultrabornological PLS-space with the property (PA). The splitting theory of Bonet and Domański can also be applied if \({\mathcal {F}}(\varOmega )\) is a non-Fréchet PLS-space and for PLH-spaces \({\mathcal {F}}(\varOmega )\), e.g. \({\mathcal {D}}_{L^{2}}\) and \(B_{2,\kappa }^{loc}(\varOmega )\) which are non-PLS-spaces, the splitting theory of Dierolf and Sieg [15, 16] is available. For applications we refer the reader to the already mentioned papers [4, 6, 15, 16, 64, 65] as well as [5, 18] where \({\mathcal {F}}(\varOmega )\) is the space of ultradistributions of Beurling type or of ultradifferentiable functions of Roumieu type and E, amongst others, the space of real analytic functions and to [30] where \({\mathcal {F}}(\varOmega )\) is the space of \({\mathcal {C}}^{\infty }\)-smooth functions or distributions.

Notably, the preceding results imply that the inhomogeneous Cauchy–Riemann equation with a right-hand side \(f\in {\mathcal {E}}(\varOmega ,E):={\mathcal {C}}^{\infty }(\varOmega ,E)\), where \(\varOmega \subset \mathbb {R}^{2}\) is open and E a locally convex Hausdorff space over \({\mathbb {C}}\) whose topology is induced by a system of seminorms \((p_{\alpha })_{\alpha \in {\mathfrak {A}}}\), given by

$$\begin{aligned} {\overline{\partial }}^{E}u:=(1/2)(\partial _{1}^{E}+i\partial _{2}^{E})u=f \end{aligned}$$
(1)

has a solution \(u\in {\mathcal {E}}(\varOmega ,E)\) if E is a Fréchet space or \(E:=F_{b}'\) where F is a Fréchet space satisfying the condition (DN) or if E is an ultrabornological PLS-space having the property (PA). Among these spaces E are several spaces of distributions like \({\mathcal {D}}(V)'\), the space of tempered distributions, the space of ultradistributions of Beurling type etc. In the present paper we study this problem under the constraint that the right-hand side f fulfils additional growth conditions given by an increasing family of positive continuous functions \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) on an increasing sequence of open subsets \((\varOmega _{n})_{n\in \mathbb {N}}\) of \(\varOmega \) with \(\varOmega =\bigcup _{n\in \mathbb {N}}\varOmega _{n}\), namely,

$$\begin{aligned} |f|_{n,m,\alpha }:=\sup _{\begin{array}{c} x\in \varOmega _{n}\\ \beta \in \mathbb {N}^{2}_{0},\,|\beta |\le m \end{array}} p_{\alpha }\bigl ((\partial ^{\beta })^{E}f(x)\bigr )\nu _{n}(x)<\infty \end{aligned}$$

for every \(n\in \mathbb {N}\), \(m\in \mathbb {N}_{0}\) and \(\alpha \in {\mathfrak {A}}\). Let us call the space of such functions \(\mathcal {EV}(\varOmega ,E)\). Our interest is in conditions on \({\mathcal {V}}\) and \((\varOmega _{n})_{n\in \mathbb {N}}\) such that there is a solution \(u\in \mathcal {EV}(\varOmega ,E)\) of (1), i.e. we search for conditions that guarantee the surjectivity of

$$\begin{aligned} {\overline{\partial }}^{E}:\mathcal {EV}(\varOmega ,E)\rightarrow \mathcal {EV}(\varOmega ,E). \end{aligned}$$

Using Grothendieck’s theory of tensor products, this was already done in [33] in the case that E is a Fréchet space. In the present paper we want to extend this result beyond the class of Fréchet spaces E. Concerning the sequence \((\varOmega _{n})_{n\in \mathbb {N}}\), we concentrate on the case that it is a sequence of strips along the real axis, i.e. \(\varOmega _{n}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<n\}\). The case that this sequence has holes along the real axis is treated in [35].

Let us briefly outline the content of our paper. In Sect. 2 we summarise the necessary definitions and preliminaries which are needed in the subsequent sections. In Sect. 3 we recall the definitions of the topological invariants \((\varOmega )\), (DN) and (PA) and provide some examples of spaces E having these invariants. Then we prove our main result on the surjectivity of Cauchy–Riemann operator on \(\mathcal {EV}(\varOmega ,E)\) which depends on these invariants (see Theorem 5). To apply our main result, the kernel \({\text {ker}}{\overline{\partial }}\) needs to have \((\varOmega )\) and in Sect. 4 we provide sufficient conditions on the weights and the sequence \((\varOmega _{n})_{n\in \mathbb {N}}\) which guarantee \((\varOmega )\) (see Theorem 10 and Corollary 13). We close this section with a special case of our main theorem where \((\varOmega _{n})_{n\in \mathbb {N}}\) is a sequence of strips along the real axis (see Corollary 17) and for example \(\nu _{n}(z):=\exp (a_{n}|{{\,\mathrm{Re}\,}}(z)|^{\gamma })\) for some \(0<\gamma \le 1\) and \(a_{n}\nearrow 0\) (see Corollary 18).

2 Notation and preliminaries

The notation and preliminaries are essentially the same as in [33, 36, Sect. 2]. We define the distance of two subsets \(M_{0}, M_{1} \subset \mathbb {R}^{2}\) w.r.t. a norm \(\Vert \cdot \Vert \) on \(\mathbb {R}^{2}\) via

$$\begin{aligned} \mathrm {d}^{\Vert \cdot \Vert }(M_{0},M_{1}) :={\left\{ \begin{array}{ll} \inf _{x\in M_{0},\,y\in M_{1}}\Vert x-y\Vert , &{} M_{0},\,M_{1} \ne \emptyset , \\ \infty , &{} M_{0}= \emptyset \;\text {or}\; M_{1}=\emptyset . \end{array}\right. } \end{aligned}$$

Moreover, we denote by \(\Vert \cdot \Vert _{\infty }\) the sup-norm, by \(|\cdot |\) the Euclidean norm on \(\mathbb {R}^{2}\), by \(\mathbb {B}_{r}(x):=\{w\in \mathbb {R}^{2}\;|\;|w-x|<r\}\) the Euclidean ball around \(x\in \mathbb {R}^{2}\) with radius \(r>0\) and identify \(\mathbb {R}^{2}\) and \({\mathbb {C}}\) as (normed) vector spaces. We denote the complement of a subset \(M\subset \mathbb {R}^{2}\) by \(M^{C}:= \mathbb {R}^{2}{\setminus } M\), the closure of M by \({\overline{M}}\) and the boundary of M by \(\partial M\). For a function \(f:M\rightarrow {\mathbb {C}}\) and \(K\subset M\) we denote by \(f_{\mid K}\) the restriction of f to K and by

$$\begin{aligned} \Vert f\Vert _{K}:=\sup _{x\in K}|f(x)| \end{aligned}$$

the sup-norm on K. By \(L^{1}(\varOmega )\) we denote the space of (equivalence classes of) \({\mathbb {C}}\)-valued Lebesgue integrable functions on a measurable set \(\varOmega \subset \mathbb {R}^{2}\) and by \(L^{q}(\varOmega )\), \(q\in \mathbb {N}\), the space of functions f such that \(f^{q}\in L^{1}(\varOmega )\). If \((a_{n})_{n\in \mathbb {N}}\) is a strictly increasing real sequence, we write \(a_{n}\nearrow 0\) resp. \(a_{n}\nearrow \infty \) if \(a_{n}<0\) for all \(n\in \mathbb {N}\) and \(\lim _{n\rightarrow \infty }a_{n}=0\) resp. \(a_{n}\ge 0\) for all \(n\in \mathbb {N}\) and \(\lim _{n\rightarrow \infty }a_{n}=\infty \).

By E we always denote a non-trivial locally convex Hausdorff space over the field \({\mathbb {C}}\) equipped with a directed fundamental system of seminorms \((p_{\alpha })_{\alpha \in {\mathfrak {A}}}\). If \(E={\mathbb {C}}\), then we set \((p_{\alpha })_{\alpha \in {\mathfrak {A}}}:=\{|\cdot |\}\). Further, we denote by L(FE) the space of continuous linear maps from a locally convex Hausdorff space F to E and sometimes write \(\langle T,f\rangle :=T(f)\), \(f\in F\), for \(T\in L(F,E)\). If \(E={\mathbb {C}}\), we write \(F':=L(F,{\mathbb {C}})\) for the dual space of F. If F and E are (linearly topologically) isomorphic, we write \(F\cong E\). We denote by \(L_{t}(F,E)\) the space L(FE) equipped with the locally convex topology of uniform convergence on the finite subsets of F if \(t=\sigma \), on the precompact subsets of F if \(t=\gamma \), on the absolutely convex, compact subsets of F if \(t=\kappa \) and on the bounded subsets of F if \(t=b\).

The so-called \(\varepsilon \)-product of Schwartz is defined by

$$\begin{aligned} F\varepsilon E:=L_{e}(F_{\kappa }',E) \end{aligned}$$
(2)

where \(L(F_{\kappa }',E)\) is equipped with the topology of uniform convergence on equicontinuous subsets of \(F'\). This definition of the \(\varepsilon \)-product coincides with the original one by Schwartz [59, Chap. I, Sect. 1, Définition, p. 18].

We recall the following well-known definitions concerning continuous partial differentiability of vector-valued functions (c.f. [34, p. 237]). A function \(f:\varOmega \rightarrow E\) on an open set \(\varOmega \subset \mathbb {R}^{2}\) to E is called continuously partially differentiable (f is \({\mathcal {C}}^{1}\)) if for the n-th unit vector \(e_{n}\in \mathbb {R}^{2}\) the limit

$$\begin{aligned} (\partial ^{e_{n}})^{E}f(x):=(\partial _{n})^{E}f(x) :=\lim _{\begin{array}{c} h\rightarrow 0\\ h\in \mathbb {R}, h\ne 0 \end{array}}\frac{f(x+he_{n})-f(x)}{h} \end{aligned}$$

exists in E for every \(x\in \varOmega \) and \((\partial ^{e_{n}})^{E}f\) is continuous on \(\varOmega \) (\((\partial ^{e_{n}})^{E}f\) is \({\mathcal {C}}^{0}\)) for every \(n\in \{1,2\}\). For \(k\in \mathbb {N}\) a function f is said to be k-times continuously partially differentiable (f is \({\mathcal {C}}^{k}\)) if f is \({\mathcal {C}}^{1}\) and all its first partial derivatives are \({\mathcal {C}}^{k-1}\). A function f is called infinitely continuously partially differentiable (f is \({\mathcal {C}}^{\infty }\)) if f is \({\mathcal {C}}^{k}\) for every \(k\in \mathbb {N}\). The linear space of all functions \(f:\varOmega \rightarrow E\) which are \({\mathcal {C}}^{\infty }\) is denoted by \({\mathcal {C}}^{\infty }(\varOmega ,E)\). Let \(f\in {\mathcal {C}}^{\infty }(\varOmega ,E)\). For \(\beta =(\beta _{n})\in \mathbb {N}_{0}^{2}\) we set \((\partial ^{\beta _{n}})^{E}f:=f\) if \(\beta _{n}=0\), and

$$\begin{aligned} (\partial ^{\beta _{n}})^{E}f :=\underbrace{(\partial ^{e_{n}})^{E}\cdots (\partial ^{e_{n}})^{E}}_{\beta _{n} \text {-times}}f \end{aligned}$$

if \(\beta _{n}\ne 0\) as well as

$$\begin{aligned} (\partial ^{\beta })^{E}f :=(\partial ^{\beta _{1}})^{E}(\partial ^{\beta _{2}})^{E}f. \end{aligned}$$

Due to the vector-valued version of Schwarz’ theorem \((\partial ^{\beta })^{E}f\) is independent of the order of the partial derivatives on the right-hand side, we call \(|\beta |:=\beta _{1}+\beta _{2}\) the order of differentiation and write \(\partial ^{\beta }f:=(\partial ^{\beta })^{{\mathbb {C}}}f\).

A function \(f:\varOmega \rightarrow E\) on an open set \(\varOmega \subset \mathbb {C}\) to E is called holomorphic if the limit

$$\begin{aligned} \biggl (\frac{\partial }{\partial z}\biggr )^{E}f(z_{0}) :=\lim _{\begin{array}{c} h\rightarrow 0\\ h\in \mathbb {C}, h\ne 0 \end{array}}\frac{f(z_{0}+h)-f(z_{0})}{h} \end{aligned}$$

exists in E for every \(z_{0}\in \varOmega \) and the space of such functions is denoted by \({\mathcal {O}}(\varOmega ,E)\). The exact definition of the spaces from the introduction is as follows.

Definition 1

[34, Definition 3.2, p. 238] Let \(\varOmega \subset \mathbb {R}^{2}\) be open and \((\varOmega _{n})_{n\in \mathbb {N}}\) a family of non-empty open sets such that \(\varOmega _{n}\subset \varOmega _{n+1}\) and \(\varOmega =\bigcup _{n\in \mathbb {N}} \varOmega _{n}\). Let \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) be a countable family of positive continuous functions \(\nu _{n}:\varOmega \rightarrow (0,\infty )\) such that \(\nu _{n}\le \nu _{n+1}\) for all \(n\in \mathbb {N}\). We call \({\mathcal {V}}\) a directed family of continuous weights on \(\varOmega \) and set for \(n\in \mathbb {N}\)

  1. (a)
    $$\begin{aligned} {\mathcal {E}}\nu _{n}(\varOmega _{n}, E):= \left\{ f \in {\mathcal {C}}^{\infty }(\varOmega _{n}, E)\; | \; \forall \;\alpha \in {\mathfrak {A}},\,m \in \mathbb {N}_{0}^{2}:\; |f|_{n,m,\alpha } < \infty \right\} \end{aligned}$$

    and

    $$\begin{aligned} \mathcal {EV}(\varOmega , E):= \left\{ f\in {\mathcal {C}}^{\infty }(\varOmega , E)\; | \;\forall \; n \in \mathbb {N}: \; f_{\mid \varOmega _{n}}\in {\mathcal {E}}\nu _{n}(\varOmega _{n}, E)\right\} \end{aligned}$$

    where

    $$\begin{aligned} |f|_{n,m,\alpha }:=\sup _{\begin{array}{c} x \in \varOmega _{n}\\ \beta \in \mathbb {N}_{0}^{2}, \, |\beta | \le m \end{array}} p_{\alpha }\bigl ((\partial ^{\beta })^{E}f(x)\bigr )\nu _{n}(x). \end{aligned}$$
  2. (b)
    $$\begin{aligned} {\mathcal {E}}\nu _{n,{\overline{\partial }}}(\varOmega _{n},E):= \left\{ f \in {\mathcal {E}}\nu _{n}(\varOmega _{n}, E)\; | \; f\in {\text {ker}}{\overline{\partial }}^{E} \right\} \end{aligned}$$

    and

    $$\begin{aligned} \mathcal {EV}_{{\overline{\partial }}}(\varOmega , E):=\{f\in \mathcal {EV}(\varOmega , E)\;|\; f\in {\text {ker}}{\overline{\partial }}^{E}\}. \end{aligned}$$
  3. (c)
    $$\begin{aligned} {\mathcal {O}}\nu _{n}(\varOmega _{n},E):= \left\{ f \in {\mathcal {O}}(\varOmega _{n}, E)\; | \; \forall \;\alpha \in {\mathfrak {A}}:\;|f|_{n,\alpha } < \infty \right\} \end{aligned}$$

    and

    $$\begin{aligned} \mathcal {OV}(\varOmega , E):=\{ f\in {\mathcal {O}}(\varOmega , E)\; | \;\forall \; n \in \mathbb {N}: \; f_{\mid \varOmega _{n}}\in {\mathcal {O}}\nu _{n}(\varOmega _{n}, E)\} \end{aligned}$$

    where

    $$\begin{aligned} |f|_{n,\alpha }:=\sup _{\begin{array}{c} x \in \varOmega _{n} \end{array}} p_{\alpha }(f(x))\nu _{n}(x). \end{aligned}$$

The subscript \(\alpha \) in the notation of the seminorms is omitted in the \({\mathbb {C}}\)-valued case. The letter E is omitted in the case \(E={\mathbb {C}}\) as well, e.g. we write \({\mathcal {E}}\nu _{n}(\varOmega _{n}):={\mathcal {E}}\nu _{n}(\varOmega _{n},{\mathbb {C}})\) and \(\mathcal {EV}(\varOmega ):=\mathcal {EV}(\varOmega ,{\mathbb {C}})\) .

A projective limit F of a sequence of locally convex Hausdorff spaces \((F_{n})_{n\in \mathbb {N}}\) is called weakly reduced if for every \(n\in \mathbb {N}\) there is \(m\in \mathbb {N}\) such that \(\pi _{n}(F)\) is dense in \(F_{m}\) w.r.t. the topology of \(F_{n}\) where \(\pi _{n}:F\rightarrow F_{n}\) is the canonical projection. The spaces \(\mathcal {FV}(\varOmega ,E)\), \({\mathcal {F}}={\mathcal {E}}\), \({\mathcal {O}}\), are projective limits, namely, we have

$$\begin{aligned} \mathcal {FV}(\varOmega , E)\cong \lim _{\begin{array}{c} \longleftarrow \\ n\in \mathbb {N} \end{array}}{\mathcal {F}}\nu _{n}(\varOmega _{n}, E) \end{aligned}$$

where the spectral maps are given by the restrictions

$$\begin{aligned} \pi _{k,n}:{\mathcal {F}}\nu _{k}(\varOmega _{k}, E)\rightarrow {\mathcal {F}}\nu _{n}(\varOmega _{n}, E),\; f\mapsto f_{\mid \varOmega _{n}},\;k\ge n. \end{aligned}$$

3 Main result

In this section we prove our main result that the surjectivity of the vector-valued Cauchy–Riemann operator on \(\mathcal {EV}(\varOmega ,E)\) is inherited from the surjectivity on \(\mathcal {EV}(\varOmega )\) if the kernel \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) in the scalar-valued case has \((\varOmega )\), and \(E:=F_{b}'\) where F is a Fréchet space satisfying the condition (DN) or E is an ultrabornological PLS-space having the property (PA). Therefore we recall the definitions of the topological invariants \((\varOmega )\), (DN) and (PA) and give some examples.

A Fréchet space F with an increasing fundamental system of seminorms satisfies \((\varOmega )\) if

$$\begin{aligned} \forall \; p\in \mathbb {N}\; \exists \; q\in \mathbb {N}\;\forall \; k\in \mathbb {N}\;\exists \; n\in \mathbb {N},\,C>0\;\forall \; r>0:\; U_{q}\subset Cr^n U_k + \frac{1}{r} U_p \end{aligned}$$
(3)

where (see [54, Chap. 29, Definition, p. 367]).

A Fréchet space satisfies (DN) by [54, Chap. 29, Definition, p. 359] if

A PLS-space is a projective limit \(X=\lim \limits _{\begin{array}{c} \longleftarrow \\ N\in \mathbb {N} \end{array}}X_{N}\), where the \(X_{N}\) given by inductive limits are DFS-spaces (which are also called LS-spaces), and it satisfies (PA) if

where denotes the dual norm of and \(i^{M}_{N}\), \(i^{K}_{N}\) the linking maps (see [6, Sect. 4, Eq. (24), p. 577]).

Due to [63, 1.4 Lemma, p. 110] and [6, Proposition 4.2, p. 577] we have the following relation between the properties (DN) and (PA).

Remark 2

Let F be a Fréchet-Schwartz space. Then F satisfies (DN) if and only if the DFS-space \(E:=F_{b}'\) satisfies (PA).

Let us summarise some examples of ultrabornological PLS-spaces satisfying (PA) and spaces of the form \(E:=F_{b}'\) where F is a Fréchet space satisfying (DN). The majority of them is already contained in [6, 19] and [64].

Example 3

(a) The following spaces are ultrabornological PLS-spaces with property (PA) and also strong duals of a Fréchet space satisfying (DN):

  • the strong dual of a power series space of inifinite type \(\varLambda _{\infty }(\alpha )_{b}'\),

  • the strong dual of any space of holomorphic functions \({\mathcal {O}}(U)_{b}'\) where U is a Stein manifold with the strong Liouville property (for instance, for \(U={\mathbb {C}}^{d}\)),

  • the space of germs of holomorphic functions \({\mathcal {O}}(K)\) where K is a completely pluripolar compact subset of a Stein manifold (for instance K consists of one point),

  • the space of tempered distributions \({\mathcal {S}}(\mathbb {R}^{d})_{b}'\) and the space of Fourier ultra-hyperfunctions \({\mathcal {P}}'_{**}\) (with the strong topology),

  • the weighted distribution spaces \((K\{pM\})_{b}'\) of Gelfand and Shilov if the weight M satisfies

    $$\begin{aligned} \sup _{|y|\le 1}M(x+y)\le C\inf _{|y|\le 1}M(x+y),\quad x\in \mathbb {R}^{d}, \end{aligned}$$
  • \({\mathcal {D}}(K)_{b}'\) for any compact set \(K\subset \mathbb {R}^{d}\) with non-empty interior,

  • \({\mathcal {C}}^{\infty }({\overline{U}})_{b}'\) for any non-empty open bounded set \(U\subset \mathbb {R}^{d}\) with \({\mathcal {C}}^{1}\)-boundary.

(b) The following spaces are ultrabornological PLS-spaces with property (PA):

  • an arbitrary Fréchet-Schwartz space,

  • a PLS-type power series space \(\varLambda _{r,s}(\alpha ,\beta )\) whenever \(s=\infty \) or \(\varLambda _{r,s}(\alpha ,\beta )\) is a Fréchet space,

  • the spaces of distributions \({\mathcal {D}}(U)_{b}'\) and ultradistributions of Beurling type \({\mathcal {D}}_{(\omega )}(U)_{b}'\) for any open set \(U\subset \mathbb {R}^{d}\),

  • the kernel of any linear partial differential operator with constant coefficients in \({\mathcal {D}}(U)_{b}'\) or in \({\mathcal {D}}_{(\omega )}(U)_{b}'\) when \(U\subset \mathbb {R}^{d}\) is open and convex,

  • the space \(L_{b}(X,Y)\) where X has (DN), Y has \((\varOmega )\) and both are nuclear Fréchet spaces. In particular, \(L_{b}(\varLambda _{\infty }(\alpha ),\varLambda _{\infty }(\beta ))\) if both spaces are nuclear.

(c) The following spaces are strong duals of a Fréchet space satisfying (DN):

  • the strong dual \(F_{b}'\) of any Banach space F,

  • the strong dual \(\lambda ^{2}(A)_{b}'\) of the Köthe space \(\lambda ^{2}(A)\) with a Köthe matrix \(A=(a_{j,k})_{j,k\in \mathbb {N}_{0}}\) satisfying

    $$\begin{aligned} \exists \;p\in \mathbb {N}_{0}\;\forall \;k\in \mathbb {N}_{0}\;\exists \;n\in \mathbb {N}_{0},C>0:\;a_{j,k}^{2}\le Ca_{j,p}a_{j,n}. \end{aligned}$$

Proof

The statement for the spaces in (a) and (b) follows from [19, Corollary 4.8, p. 1116], [54, Proposition 31.12, p. 401], [54, Proposition 31.16, p. 402] and Remark 2. The first part of statement (c) is obvious since Banach spaces clearly satisfy the property (DN). The second part on the Köthe space \(\lambda ^{2}(A)\) follows from [29, Satz 12.11 a), p. 305]. \(\square \)

Since we will use the \(\varepsilon \)-product \(\mathcal {EV}(\varOmega )\varepsilon E\) to pass the surjectivity from \({\overline{\partial }}\) to \({\overline{\partial }}^{E}\), we remark the following which is not hard to prove (see [31, Sect. 39]).

Proposition 4

  1. (a)

    Let X be a semi-reflexive locally convex Hausdorff space and Y a Fréchet space. Then \(L_{b}(X_{b}',Y_{b}')\cong L_{b}(Y,(X_{b}')_{b}')\) via taking adjoints.

  2. (b)

    Let X be a Montel space and E a locally convex Hausdorff space. Then \(L_{b}(X_{b}',E)\cong X\varepsilon E\) where the topological isomorphism is the identity map.

Theorem 5

Let \(\mathcal {EV}(\varOmega )\) be a Schwartz space and \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) a nuclear subspace satisfying property \((\varOmega )\). Assume that the scalar-valued operator \({\overline{\partial }}:\mathcal {EV}(\varOmega )\rightarrow \mathcal {EV}(\varOmega )\) is surjective. Moreover, if

  1. (a)

    \(E:=F_{b}'\) where F is a Fréchet space over \({\mathbb {C}}\) satisfying (DN), or

  2. (b)

    E is an ultrabornological PLS-space over \({\mathbb {C}}\) satisfying (PA),

then

$$\begin{aligned} {\overline{\partial }}^{E}:\mathcal {EV}(\varOmega ,E)\rightarrow \mathcal {EV}(\varOmega ,E) \end{aligned}$$

is surjective.

Proof

Throughout this proof we use the notation \(X'':=(X_{b}')_{b}'\) for a locally convex Hausdorff space X. In both cases, (a) and (b), the space E is a complete locally convex Hausdorff space. The space \(\mathcal {EV}(\varOmega )\) is a Fréchet space by [34, Proposition 3.7, p. 240] and so its closed subspace \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) as well. Further, \(\mathcal {EV}(\varOmega )\) is a Schwartz space and \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) nuclear, thus both spaces are reflexive. As the Fréchet-Schwartz space \(\mathcal {EV}(\varOmega )\) is a Montel space,

$$\begin{aligned} S:\mathcal {EV}(\varOmega )\varepsilon E\rightarrow \mathcal {EV}(\varOmega ,E),\; u\longmapsto [z\mapsto u(\delta _{z})], \end{aligned}$$

is a topological isomorphism by [36, 3.21 Example b), p. 14] where \(\delta _{z}\) is the point-evaluation at \(z\in \varOmega \). We denote by \({\mathcal {J}}:E\rightarrow E'^{*}\) the canonical injection in the algebraic dual \(E'^{*}\) of the topological dual \(E'\) and for \(f\in \mathcal {EV}(\varOmega ,E)\) we set

$$\begin{aligned} R_{f}^{t}:\mathcal {EV}(\varOmega )'\rightarrow E'^{\star },\; y\longmapsto \bigl [e'\mapsto y(e'\circ f) \bigr ]. \end{aligned}$$

Then the map \(f\mapsto {\mathcal {J}}^{-1}\circ R_{f}^{t}\) is the inverse of S by [36, 3.17 Theorem, p. 12]. The sequence

$$\begin{aligned} 0\rightarrow \mathcal {EV}_{{\overline{\partial }}}(\varOmega )\overset{i}{\rightarrow }\mathcal {EV}(\varOmega ) \overset{{\overline{\partial }}}{\rightarrow }\mathcal {EV}(\varOmega )\rightarrow 0, \end{aligned}$$
(4)

where i means the inclusion, is a topologically exact sequence of Fréchet spaces because \({\overline{\partial }}\) is surjective by assumption. Let us denote by \(J_{0}:\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\rightarrow \mathcal {EV}_{{\overline{\partial }}}(\varOmega )''\) and \(J_{1}:\mathcal {EV}(\varOmega )\rightarrow \mathcal {EV}(\varOmega )''\) the canonical embeddings which are topological isomorphisms since \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) and \(\mathcal {EV}(\varOmega )\) are reflexive. Then the exactness of (4) implies that

$$\begin{aligned} 0\rightarrow \mathcal {EV}_{{\overline{\partial }}}(\varOmega )''\overset{i_{0}}{\rightarrow } \mathcal {EV}(\varOmega )'' \overset{{\overline{\partial }}_{1}}{\rightarrow }\mathcal {EV}(\varOmega )''\rightarrow 0, \end{aligned}$$
(5)

where \(i_{0}:=J_{0}\circ i\circ J^{-1}_{0}\) and \({\overline{\partial }}_{1}:=J_{1}\circ {\overline{\partial }}\circ J^{-1}_{1}\), is an exact topological sequence. Topological as the (strong) bidual of a Fréchet space is again a Fréchet space by [54, Corollary 25.10, p. 298].

(a) Let \(E:=F_{b}'\) where F is a Fréchet space with (DN). Then \({\text {Ext}}^{1}(F,\mathcal {EV}_{{\overline{\partial }}}(\varOmega )'')=0\) by [65, 5.1 Theorem, p. 186] since \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) satisfies \((\varOmega )\) and therefore \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )''\) as well. Combined with the exactness of (5) this implies that the sequence

$$\begin{aligned} 0\rightarrow L(F,\mathcal {EV}_{{\overline{\partial }}}(\varOmega )'')\overset{i^{*}_{0}}{\rightarrow }L(F,\mathcal {EV}(\varOmega )'') \overset{{\overline{\partial }}^{*}_{1}}{\rightarrow }L(F,\mathcal {EV}(\varOmega )'')\rightarrow 0 \end{aligned}$$

is exact by [57, Proposition 2.1, p. 13-14] where \(i^{*}_{0}(B):=i_{0}\circ B\) and \({\overline{\partial }}^{*}_{1}(D):={\overline{\partial }}_{1}\circ D\) for \(B\in L(F,\mathcal {EV}_{{\overline{\partial }}}(\varOmega )'')\) and \(D\in L(F,\mathcal {EV}(\varOmega )'')\). In particular, we obtain that

$$\begin{aligned} {\overline{\partial }}^{*}_{1}:L(F,\mathcal {EV}(\varOmega )'')\rightarrow L(F,\mathcal {EV}(\varOmega )'') \end{aligned}$$
(6)

is surjective. Via \(E=F_{b}'\) and Proposition 4 (\(X=\mathcal {EV}(\varOmega )\) and \(Y=F\)) we have the topological isomorphism

$$\begin{aligned} \psi :=S\circ {^{t}(\cdot )}:L(F,\mathcal {EV}(\varOmega )'')\rightarrow \mathcal {EV}(\varOmega ,E),\; \psi (u)=\bigl (S\circ {^{t}(\cdot )}\bigr )(u)=\bigl [z\mapsto {^{t}u}(\delta _{z})\bigr ], \end{aligned}$$

and the inverse

$$\begin{aligned} \psi ^{-1}(f) =(S\circ {^{t}(\cdot )})^{-1}(f) =({^{t}(\cdot )}\circ S^{-1})(f) ={^{t}({\mathcal {J}}^{-1}\circ R_{f}^{t})},\quad f\in \mathcal {EV}(\varOmega ,E). \end{aligned}$$

Let \(g\in \mathcal {EV}(\varOmega ,E)\). Then \(\psi ^{-1}(g)\in L(F,\mathcal {EV}(\varOmega )'')\) and by the surjectivity of (6) there is \(u\in L(F,\mathcal {EV}(\varOmega )'')\) such that \({\overline{\partial }}^{*}_{1}u=\psi ^{-1}(g)\). So we get \(\psi (u)\in \mathcal {EV}(\varOmega ,E)\). Next, we show that \({\overline{\partial }}^{E}\psi (u)=g\) is valid. Let \(x\in F\), \(z\in \varOmega \) and \(h\in \mathbb {R}\), \(h\ne 0\), and \(e_{k}\) denote the kth unit vector in \(\mathbb {R}^{2}\). From

$$\begin{aligned} \bigl (\frac{\delta _{z+he_{k}}-\delta _{z}}{h}\bigr )(f)=\frac{f(z+he_{k})-f(z)}{h}\underset{h\rightarrow 0}{\rightarrow }\partial ^{e_{k}}f(z), \end{aligned}$$

for every \(f\in \mathcal {EV}(\varOmega )\) it follows that \(\frac{\delta _{z+he_{k}}-\delta _{z}}{h}\) converges to \(\delta _{z}\circ \partial ^{e_{k}}\) in \(\mathcal {EV}(\varOmega )_{\sigma }'\). Since the Fréchet–Schwartz space \(\mathcal {EV}(\varOmega )\) is in particular a Montel space, we deduce that \(\frac{\delta _{z+he_{k}}-\delta _{z}}{h}\) converges to \(\delta _{z}\circ \partial ^{e_{k}}\) in \(\mathcal {EV}(\varOmega )_{\gamma }'=\mathcal {EV}(\varOmega )_{b}'\) by the Banach–Steinhaus theorem. Let \(B\subset F\) be bounded. As \({^{t}u}\in L(\mathcal {EV}(\varOmega )_{b}',F_{b}')\), there are a bounded set \(B_{0}\subset \mathcal {EV}(\varOmega )\) and \(C>0\) such that

$$\begin{aligned}&\sup _{x\in B}\bigl |\bigl (\frac{{^{t}u}(\delta _{z+he_{k}})-{^{t}u}(\delta _{z})}{h}\bigr )(x) -{^{t}u}\bigl (\delta _{z}\circ \partial ^{e_{k}}\bigr )(x)\bigr |\\&=\sup _{x\in B}\bigl |{^{t}u}\bigl (\frac{\delta _{z+he_{k}}-\delta _{z}}{h}-\delta _{z}\circ \partial ^{e_{k}}\bigr )(x)\bigr | \le C\sup _{f\in B_{0}}\bigl |\bigl (\frac{\delta _{z+he_{k}}-\delta _{z}}{h}-\delta _{z}\circ \partial ^{e_{k}}\bigr )(f)\bigr | \underset{h\rightarrow 0}{\rightarrow }0, \end{aligned}$$

yielding to \((\partial ^{e_{k}})^{E}(\psi (u))(z)={^{t}u}(\delta _{z}\circ \partial ^{e_{k}})\). This implies \({\overline{\partial }}^{E}(\psi (u))(z)={^{t}u}(\delta _{z}\circ {\overline{\partial }})\). So for all \(x\in F\) and \(z\in \varOmega \) we have

$$\begin{aligned} {\overline{\partial }}^{E}(\psi (u))(z)(x)&={^{t}u}(\delta _{z}\circ {\overline{\partial }})(x) =u(x)(\delta _{z}\circ {\overline{\partial }}) =\langle \delta _{z}\circ {\overline{\partial }},J^{-1}_{1}(u(x))\rangle \\&=\langle \delta _{z},{\overline{\partial }}J^{-1}_{1}(u(x))\rangle =\langle [J_{1}\circ {\overline{\partial }}\circ J^{-1}_{1}](u(x)),\delta _{z}\rangle =\langle ({\overline{\partial }}_{1}\circ u)(x),\delta _{z}\rangle \\&=\langle ({\overline{\partial }}^{*}_{1}u)(x),\delta _{z}\rangle =\psi ^{-1}(g)(x)(\delta _{z}) ={^{t}({\mathcal {J}}^{-1}\circ R_{g}^{t})}(x)(\delta _{z})\\&=({\mathcal {J}}^{-1}\circ R_{g}^{t})(\delta _{z})(x) ={\mathcal {J}}^{-1}({\mathcal {J}}(g(z))(x) =g(z)(x). \end{aligned}$$

Thus \({\overline{\partial }}^{E}(\psi (u))(z)=g(z)\) for every \(z\in \varOmega \), which proves the surjectivity.

(b) Let E be an ultrabornological PLS-space satisfying (PA). Since the nuclear Fréchet space \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) is also a Schwartz space, its strong dual \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )_{b}'\) is a DFS-space. By [6, Theorem 4.1, p. 577] we obtain \({\text {Ext}}^{1}_{PLS}(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )_{b}',E)=0\) as the bidual \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )''\) satisfies \((\varOmega )\), E is a PLS-space satisfying (PA) and condition (c) in the theorem is fulfilled because \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )_{b}'\) is the strong dual of a nuclear Fréchet space. Moreover, we have \({\text {Proj}}^{1} E=0\) due to [66, Corollary 3.3.10, p. 46] because E is an ultrabornological PLS-space. Then the exactness of the sequence (5), [6, Theorem 3.4, p. 567] and [6, Lemma 3.3, p. 567] (in the lemma the same condition (c) as in [6, Theorem 4.1, p. 577] is fulfilled and we choose \(H=\mathcal {EV}_{{\overline{\partial }}}(\varOmega )''\) and \(F=G=\mathcal {EV}(\varOmega )''\)), imply that the sequence

$$\begin{aligned} 0\rightarrow L(E_{b}',\mathcal {EV}_{{\overline{\partial }}}(\varOmega )'') \overset{i^{*}_{0}}{\rightarrow }L(E_{b}',\mathcal {EV}(\varOmega )'') \overset{{\overline{\partial }}^{*}_{1}}{\rightarrow }L(E_{b}',\mathcal {EV}(\varOmega )'')\rightarrow 0 \end{aligned}$$

is exact. The maps \(i^{*}_{0}\) and \({\overline{\partial }}^{*}_{1}\) are defined like in part (a). Especially, we get that

$$\begin{aligned} {\overline{\partial }}^{*}_{1}:L(E_{b}',\mathcal {EV}(\varOmega )'')\rightarrow L(E_{b}',\mathcal {EV}(\varOmega )'') \end{aligned}$$
(7)

is surjective.

By [19, Remark 4.4, p. 1114] we have \(L_{b}(\mathcal {EV}(\varOmega )_{b}',E'') \cong L_{b}(E_{b}',\mathcal {EV}(\varOmega )'')\) via taking adjoints since \(\mathcal {EV}(\varOmega )\), being a Fréchet–Schwartz space, is a PLS-space and hence its strong dual an LFS-space, which is regular by [66, Corollary 6.7, \(10.\Leftrightarrow 11.\), p. 114], and E is an ultrabornological PLS-space, in particular, reflexive by [17, Theorem 3.2, p. 58]. In addition, the map

$$\begin{aligned} T:L_{b}(\mathcal {EV}(\varOmega )_{b}',E'')\rightarrow L_{b}(\mathcal {EV}(\varOmega )_{b}',E), \end{aligned}$$

defined by \(T(u)(y):={\mathcal {J}}^{-1}(u(y))\) for \(u\in L(\mathcal {EV}(\varOmega )_{b}',E'')\) and \(y\in \mathcal {EV}(\varOmega )'\), is a topological isomorphism because E is reflexive. Due to Proposition 4 (b) we obtain the topological isomorphism

$$\begin{aligned} \psi :=S\circ {\mathcal {J}}^{-1}\circ {^{t}(\cdot )}:L_{b}(E_{b}',\mathcal {EV}(\varOmega )'') \rightarrow \mathcal {EV}(\varOmega ,E),\\ \psi (u)=[S\circ {\mathcal {J}}^{-1}\circ {^{t}(\cdot )}](u)=\bigl [z\mapsto {\mathcal {J}}^{-1}({^{t}u}(\delta _{z}))\bigr ], \end{aligned}$$

with the inverse given by

$$\begin{aligned} \psi ^{-1}(f) =(S\circ {\mathcal {J}}^{-1}\circ {^{t}(\cdot )})^{-1}(f) =[{^{t}(\cdot )}\circ {\mathcal {J}}\circ S^{-1}](f) ={^{t}({\mathcal {J}}\circ {\mathcal {J}}^{-1}\circ R_{f}^{t})}={^{t}(R_{f}^{t})} \end{aligned}$$

for \(f\in \mathcal {EV}(\varOmega ,E)\).

Let \(g\in \mathcal {EV}(\varOmega ,E)\). Then \(\psi ^{-1}(g)\in L(E_{b}',\mathcal {EV}(\varOmega )'')\) and by the surjectivity of (7) there exists \(u\in L(E_{b}',\mathcal {EV}(\varOmega )'')\) such that \({\overline{\partial }}^{*}_{1}u=\psi ^{-1}(g)\). So we have \(\psi (u)\in \mathcal {EV}(\varOmega ,E)\). The last step is to show that \({\overline{\partial }}^{E}\psi (u)=g\). Like in part (a) we gain for every \(z\in \varOmega \)

$$\begin{aligned} {\overline{\partial }}^{E}(\psi (u))(z)={\mathcal {J}}^{-1}({^{t}u} (\delta _{z}\circ {\overline{\partial }})) \end{aligned}$$

and for every \(x\in E'\)

$$\begin{aligned} {^{t}u}(\delta _{z}\circ {\overline{\partial }})(x)&=u(x)(\delta _{z}\circ {\overline{\partial }}) =({\overline{\partial }}^{*}_{1}u)(x)(\delta _{z}) =\psi ^{-1}(g)(x)(\delta _{z}) ={^{t}(R_{g}^{t})}(x)(\delta _{z})\\&=\delta _{z}(x\circ g) =x(g(z)) ={\mathcal {J}}(g(z))(x). \end{aligned}$$

Thus we have \({^{t}u}(\delta _{z}\circ {\overline{\partial }})={\mathcal {J}}(g(z))\) and therefore \({\overline{\partial }}^{E}(\psi (u))(z)=g(z)\) for all \(z\in \varOmega \). \(\square \)

By Remark 2 case (a) is included in case (b) if F is a Fréchet–Schwartz space. Therefore (a) is only interesting for Fréchet spaces F which are not Schwartz spaces. In the next more technical section we will present sufficient conditions for \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) to have \((\varOmega )\) as well as concrete examples of such spaces.

4 \((\varOmega )\) for \(\mathcal {OV}\)-spaces on strips and applications of the main result

In this section we give some sufficient conditions such that the assumptions of our main result Theorem 5 are fulfilled. The outline is as follows. First, we show that \(\mathcal {OV}(\varOmega )\) and \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) coincide topologically under mild assumptions on the weights \({\mathcal {V}}\) and the sequence of sets \((\varOmega _{n})\). These mild conditions also imply that \(\mathcal {EV}(\varOmega )\) is nuclear, in particular Schwartz, and thus its subspace \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )=\mathcal {OV}(\varOmega )\) too. Second, we reduce the problem whether the projective limit \(\mathcal {OV}(\varOmega )\) has \((\varOmega )\) to the problem whether it is weakly reduced in the case that the \(\varOmega _{n}\) are strips along the real axis and the weights have a certain structure. Third, we use a similar result for \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) which was obtained in [33] to prove the weak reducibility of \(\mathcal {OV}(\varOmega )\). For corresponding results in the case that \(\varOmega _{n}=\varOmega \) for all \(n\in \mathbb {N}\) see [20, Theorem 3, p. 56], [39, 1.3 Lemma, p. 418] and [58, Theorem 1, p. 145]. We close this section with some examples of our main result. Let us start with the sufficient conditions, guaranteeing that the projective limit \(\mathcal {EV}(\varOmega )\) is nuclear (if \(q=1\)). They also allow to switch from \(\sup \)- to weighted \(L^{q}\)-seminorms which is important for the proof of surjectivity of the scalar-valued \({\overline{\partial }}\)-operator given in [33], using Hörmander’s \(L^{2}\)-machinery (if \(q=2\)).

Condition (PN)

([33, 3.3 Condition, p. 7]) Let \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) be a directed family of continuous weights on an open set \(\varOmega \subset \mathbb {R}^{2}\) and \((\varOmega _{n})_{n\in \mathbb {N}}\) a family of non-empty open sets such that \(\varOmega _{n}\subset \varOmega _{n+1}\) and \(\varOmega =\bigcup _{n\in \mathbb {N}} \varOmega _{n}\). For every \(k\in \mathbb {N}\) let there be \(\rho _{k}\in \mathbb {R}\) such that \(0<\rho _{k}<\mathrm {d}^{\Vert \cdot \Vert _{\infty }}(\{x\},\partial \varOmega _{k+1})\) for all \(x\in \varOmega _{k}\) and let there be \(q\in \mathbb {N}\) such that for any \(n\in \mathbb {N}\) there is \(\psi _{n}\in L^{q}(\varOmega _{k})\), \(\psi _{n}>0\), and \(\mathbb {N}\ni J_{i}(n)\ge n\) and \(C_{i}(n)>0\) such that for any \(x\in \varOmega _{k}\):

(PN.1):

\(\sup _{\zeta \in \mathbb {R}^{2},\,\Vert \zeta \Vert _{\infty }\le \rho _{k}}\nu _{n}(x+\zeta ) \le C_{1}(n)\inf _{\zeta \in \mathbb {R}^{2},\,\Vert \zeta \Vert _{\infty }\le \rho _{k}}\nu _{J_{1}(n)}(x+\zeta )\)

\((PN.2)^{q}\):

\(\nu _{n}(x)\le C_{2}(n)\psi _{n}(x)\nu _{J_{2}(n)}(x)\)

Example 6

Let \(\varOmega :=\mathbb {R}^{2}\) and \( \varOmega _{n}:=\{x=(x_{i})\in \mathbb {R}^{2}\;||x_{2}|<n\}. \) Let \(0<\gamma \le 1\) and \((a_{n})_{n\in \mathbb {N}}\) be strictly increasing such that \(a_{n}\ge 0\) for all \(n\in \mathbb {N}\) or \(a_{n}\le 0\) for all \(n\in \mathbb {N}\). The family \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) of positive continuous functions on \(\varOmega \) given by

$$\begin{aligned} \nu _{n}:\varOmega \rightarrow (0,\infty ),\;\nu _{n}(x):=e^{a_{n}|x_{1}|^{\gamma }}, \end{aligned}$$

fulfils \(\nu _{n}\le \nu _{n+1}\) all \(n\in \mathbb {N}\) and (PN) for every \(q\in \mathbb {N}\) with \(\psi _{n}(x):= (1+|x|^{2})^{-2}\), \(x\in \mathbb {R}^{2}\), for every \(n\in \mathbb {N}\).

The space \(\mathcal {OV}({\mathbb {C}})\) with this kind of weights consists of functions which are entire and exponentially growing \((a_{n}<0)\) resp. decreasing (\(a_{n}>0\)) with order \(\gamma \) on strips along the real axis. This example of weights and many more are included in [33, 3.7 Example, p. 9]. We restrict to this particular weights because we use it in an example for our main result.

Proposition 7

Let \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) be a directed family of continuous weights on an open set \(\varOmega \subset \mathbb {R}^{2}\) and \((\varOmega _{n})_{n\in \mathbb {N}}\) a family of non-empty open sets such that \(\varOmega _{n}\subset \varOmega _{n+1}\) and \(\varOmega =\bigcup _{n\in \mathbb {N}} \varOmega _{n}\). If (PN.1) is fulfilled, then

  1. (a)

    for every \(n\in \mathbb {N}\) and \(m\in \mathbb {N}_{0}\) there is \(C>0\) such that

    $$\begin{aligned} |f|_{n,m}\le C|f|_{2J_{1}(n)},\quad f\in {\mathcal {O}}\nu _{2J_{1}(n)}(\varOmega _{2J_{1}(n)}). \end{aligned}$$
  2. (b)

    \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )=\mathcal {OV}(\varOmega )\) as Fréchet spaces.

Proof

(a) Let \(n\in \mathbb {N}\) and \(m\in \mathbb {N}_{0}\). We note that \(\varOmega _{n+1}\subset \varOmega _{2J_{1}(n)}\) and \(\partial ^{\beta }f(x)=i^{\beta _{2}}f^{(|\beta |)}(x)\), \(x\in \varOmega _{2J_{1}(n)}\), holds for all \(\beta =(\beta _{1},\beta _{2})\in \mathbb {N}_{0}^{2}\) and \(f\in {\mathcal {O}}\nu _{2J_{1}(n)}(\varOmega _{2J_{1}(n)})\) where \(f^{(|\beta |)}\) is the \(|\beta |\)th complex derivative of f. Then we obtain via (PN.1) and Cauchy’s inequality

(b) The space \(\mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) is a Fréchet space since it is a closed subspace of the Fréchet space \(\mathcal {EV}(\varOmega )\) by [34, Proposition 3.7, p. 240]. From part (a) and \(|f|_{n}=|f|_{n,0}\) for all \(n\in \mathbb {N}\) and \(f\in \mathcal {EV}_{{\overline{\partial }}}(\varOmega )\) follows the statement. \(\square \)

Let us come to the second part. Using special weight functions, strips along the real axis as \(\varOmega _{n}\) and a decomposition theorem of Langenbruch, we will see that answering the question whether \(\mathcal {OV}(\varOmega )\) satisfies the property \((\varOmega )\) of Vogt boils down to answering whether the projective limit \(\mathcal {OV}(\varOmega )\) is weakly reduced. The special weights we want to consider are generated by a function \(\mu \) with the following properties.

Definition 8

(strong weight generator) A continuous function \(\mu :{\mathbb {C}}\rightarrow [0,\infty )\) is called a weight generator if \(\mu (z)=\mu (|{{\,\mathrm{Re}\,}}(z)|)\) for all \(z\in {\mathbb {C}}\), the restriction \(\mu _{\mid [0,\infty )}\) is strictly increasing,

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow \infty \\ x\in \mathbb {R} \end{array}}\frac{\ln (1+|x|)}{\mu (x)}=0 \end{aligned}$$

and

$$\begin{aligned} \exists \;\varGamma>1,\,C>0\;\forall \;x\in [0,\infty ):\; \mu (x+1)\le \varGamma \mu (x)+C. \end{aligned}$$

If \(\mu \) is a weight generator which fulfils the stronger condition

$$\begin{aligned} \exists \;\varGamma>1\;\forall \;n\in \mathbb {N}\;\exists \;C>0\;\forall \;x\in [0,\infty ):\; \mu (x+n)\le \varGamma \mu (x)+C, \end{aligned}$$

then \(\mu \) is called a strong weight generator.

Weight generators are introduced in [46, Definition 2.1, p. 225] and strong weight generators in [60, Definition 2.2.2, p. 43] where they are simply called weight functions resp. strong weight functions. For a weight generator \(\mu \) we define the space

$$\begin{aligned} H_{\tau }(S_{t}):=\{f\in {\mathcal {O}}(S_{t})\;|\;\Vert f\Vert _{\tau ,t}:=\sup _{z\in S_{t}}|f(z)|e^{\tau \mu (z)}<\infty \} \end{aligned}$$

for \(t>0\) and \(\tau \in \mathbb {R}\) with the strip \(S_{t}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<t\}\) .

Theorem 9

[46, Theorem 2.2, p. 225]Footnote 1 Let \(\mu \) be a weight generator. There are \({\widetilde{t}}\), \(K_{1}\), \(K_{2}>0\) such that for any \(\tau _{0}<\tau <\tau _{2}\) there is \(C_{0}=C_{0}({\text {sign}}(\tau ))\) such that for any \(0<2t_{0}<t<t_{2}<{\widetilde{t}}\) with

$$\begin{aligned} t_{0}\le \min \Bigl [K_{1},K_{2}\sqrt{\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}}}\Bigr ] \end{aligned}$$

there is \(C_{1}\ge 1\) such that for any \(r\ge 0\) and any \(f\in H_{\tau }(S_{t})\) with \(\Vert f\Vert _{\tau ,t}\le 1\) the following holds: there are \(f_{2}\in {\mathcal {O}}(S_{t_{2}})\) and \(f_{0}\in {\mathcal {O}}(S_{t_{0}})\) such that \(f=f_{0}+f_{2}\) on \(S_{t_{0}}\) and

$$\begin{aligned} \Vert f_{0}\Vert _{C_{0}\tau _{0},t_{0}}\le C_{1}e^{-Gr}\quad \text {and}\quad \Vert f_{2}\Vert _{\tau _{2},t_{2}}\le e^{r} \end{aligned}$$

where

$$\begin{aligned} G:= K_{1} \min \Bigl [1,\frac{t-t_{0}}{2{\widetilde{t}}},\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}}\Bigr ]. \end{aligned}$$

To apply this theorem, we have to know the constants involved. In the following the notation of [46] is used and it is referred to the corresponding positions resp. conditions for these constants. We have

$$\begin{aligned} {\widetilde{t}}:=\frac{1}{4 \ln (\varGamma )} \end{aligned}$$

by [46, Lemma 2.4, (2.15), p. 228] with \(\varGamma \) from Definition 8 such that \(\varGamma \ge e^{1/4}\). The choice \(\varGamma \ge e^{1/4}\) comes from wanting \({\widetilde{t}}\le 1\) in [46, Lemma 2.4, p. 228]. By [46, Corollary 2.6, p. 230-231] we have

$$\begin{aligned} C_{0}:= {\left\{ \begin{array}{ll}4 \varGamma B_{3}=\frac{64\cosh (1)}{\cos (1/2)}\varGamma ^{2}>1 &{},\tau<0,\\ \frac{1}{4 \varGamma B_{3}}=\frac{\cos (1/2)}{64\cosh (1)\varGamma ^{2}}<1 &{},\tau \ge 0, \end{array}\right. } \end{aligned}$$

where \(B_{3}:=\frac{16\cosh (1)}{\cos (1/2)}\varGamma \) by [46, Lemma 2.4, p. 228–229].Footnote 2 To get the constants \(K_{1}\) and \(K_{2}\), we have to analyze the conditions for \(t_{0}\) in the proof of [46, Theorem 2.2, p. 225]. By the assumptions on \(\tau _{0}\), \(\tau \) and \(\tau _{2}\) and the choice of \(C_{0}\) we obtain

$$\begin{aligned} \tau _{2}-C_{0}\tau _{0}>\tau _{2}-C_{0}\tau \ge \tau _{2}-\tau >0 \end{aligned}$$
(8)

and

$$\begin{aligned} \tau -C_{0}\tau _{0}>\tau -C_{0}\tau =\tau (1-C_{0})>0. \end{aligned}$$
(9)

By choosing \(D>0\) in the proof of [46, Theorem 2.2, (2.22), p. 232–233] as \(D:=\frac{\tau -C_{0}\tau _{0}}{(\tau _{2}-C_{0}\tau _{0})2\varGamma _{0}}\), the estimate

$$\begin{aligned} D=\frac{\tau -C_{0}\tau _{0}}{(\tau _{2}-C_{0}\tau _{0})2\varGamma _{0}} =\min \Bigl (\frac{1}{2{\widetilde{\varGamma }}},\frac{1}{2{\widehat{\varGamma }}}\Bigr ) \frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}} \underset{(8),\,(9)}{\le }\min \Bigl (\frac{1}{2{\widetilde{\varGamma }}}, \frac{1}{2{\widehat{\varGamma }}}\Bigr ) \frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau } \end{aligned}$$

holds where \(\varGamma _{0}:=\max ({\widetilde{\varGamma }},{\widehat{\varGamma }})\) with \({\widetilde{\varGamma }}\), \({\widehat{\varGamma }}>1\) from the proof. With \(\theta \ge \frac{t-t_{0}}{2{\widetilde{t}}}\) (p. 232) we get on p. 233, below (2.24), due to the condition \(t_{0}\le T_{0}:=\min (\frac{t}{2},\frac{1}{4a^{2}B_{1}{\widetilde{t}}})\),

$$\begin{aligned} \min \Bigl (\frac{\theta }{2}, D, 1\Bigr )&\ge \min \Bigl (\frac{1}{2},\frac{1}{2\varGamma _{0}}\Bigr ) \min \Bigl (\theta ,\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}},1\Bigr ) \ge \frac{1}{2\varGamma _{0}}\min \Bigl (\frac{t-t_{0}}{2{\widetilde{t}}},\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}},1\Bigr )\\&\ge \min \Bigl (\frac{1}{2\varGamma _{0}},\frac{1}{4a^{2}B_{1}{\widetilde{t}}}\Bigr )\min \Bigl (\frac{t-t_{0}}{2{\widetilde{t}}}, \frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}},1\Bigr )\\&=\underbrace{\min \Bigl (\frac{1}{2\varGamma _{0}},\frac{1}{2\cosh (1)\ln (\varGamma )}\Bigr )}_{=:K_{1}} \min \Bigl (\frac{t-t_{0}}{2{\widetilde{t}}},\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}},1\Bigr )=:G \end{aligned}$$

where \(a:=\ln (\varGamma )\) (in the middle of p. 231) and \(B_{1}:=2\cosh (1)\) by the proof of [46, Lemma 2.3, p. 226–227]. The assumptions \(2t_{0}<t\) and \(t_{0}\le K_{1}\) in Theorem 9 guarantee that the condition \(t_{0}\le T_{0}\) is satisfied. Looking at the condition \(t_{0}\le T_{1}:=\sqrt{\frac{D}{a^{2}B_{1}}}\) (p. 232), we derive

$$\begin{aligned} T_{1}=\frac{1}{\sqrt{2\varGamma _{0}a^{2}B_{1}}} \sqrt{\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}}} =\underbrace{\frac{1}{2\sqrt{\cosh (1)\varGamma _{0}} \ln (\varGamma )}}_{=:K_{2}}\sqrt{\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}}}. \end{aligned}$$

For the subsequent theorem we merge and modify the proofs of [60, Satz 2.2.3, p. 44]Footnote 3 (\(a_{n}=n\), \(n\in \mathbb {N}\), and \(\mu \) a strong weight generator) and [32, 5.20 Theorem, p. 84] (\(a_{n}=-1/n\), \(n\in \mathbb {N}\), and \(\mu =|{{\,\mathrm{Re}\,}}(\cdot )|\)).

Theorem 10

Let \(\mu \) be a strong weight generator, \(a_{n}\nearrow 0\) or \(a_{n}\nearrow \infty \), \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) and \(\varOmega _{n}:=S_{n}\) for all \(n\in \mathbb {N}\). If \(\mathcal {OV}({\mathbb {C}})\) is weakly reduced, then \(\mathcal {OV}({\mathbb {C}})\) satisfies \((\varOmega )\).

Proof

Since \(\mathcal {OV}({\mathbb {C}})\) is weakly reduced, for every \(n\in \mathbb {N}\) there exists \(m_{n}\in \mathbb {N}\) such that \(\pi _{n}(\mathcal {OV}({\mathbb {C}}))\) is dense in \({\mathcal {O}}\nu _{m_{n}}(\varOmega _{m_{n}})\) w.r.t. the topology of \({\mathcal {O}}\nu _{n}(\varOmega _{n})\) where

$$\begin{aligned} \pi _{n}:\mathcal {OV}({\mathbb {C}})\rightarrow {\mathcal {O}}\nu _{n} (\varOmega _{n}),\;\pi _{n}(f):=f_{\mid \varOmega _{n}}, \end{aligned}$$

is the canonical projection. Let \(p,k\in \mathbb {N}\). As \((a_{n})_{n\in \mathbb {N}}\) is strictly increasing and \(\lim _{n\rightarrow \infty }a_{n}=0\) or \(\lim _{n\rightarrow \infty }a_{n}=\infty \), we may choose \(q\in \mathbb {N}\) such that \(a_{m_{p}}/C_{0}<a_{q}\) and \(2m_{p}<q\). To use the decomposition from Theorem 9, we need a linear transformation between strips to get the decomposition on the desired strip \(S_{m_{p}}\). We choose \(\varGamma \ge e^{1/4}\) and \(T\in \mathbb {R}\) such that

$$\begin{aligned} 0<T<\frac{1}{4\max (q+1,m_{k})\ln (\varGamma )} \end{aligned}$$
(10)

which also fulfils

$$\begin{aligned} T\le \frac{1}{m_{p}}\min \Biggl (\frac{1}{2\varGamma _{0}},\frac{1}{2\cosh (1)\ln (\varGamma )}, \frac{1}{2\sqrt{\cosh (1)\varGamma _{0}}\ln (\varGamma )} \sqrt{\frac{a_{q}-a_{m_{p}}}{\max (a_{q+1},a_{m_{k}})-a_{m_{p}}}}\Biggr ).\nonumber \\ \end{aligned}$$
(11)

Let

$$\begin{aligned} \tau _{0}&:=\frac{a_{m_{p}}}{C_{0}},&\tau :=a_{q},&\tau _{2}:=\max (a_{q+1},a_{m_{k}}),\\ t_{0}&:= m_{p}T,&t:=qT,&t_{2}:=\max (q+1,m_{k})T. \end{aligned}$$

By the choice of q we have

$$\begin{aligned} \tau _{0}=\frac{a_{m_{p}}}{C_{0}}<a_{q}=\tau <\max (a_{q+1},a_{m_{k}})=\tau _{2}. \end{aligned}$$

By the choice of q and (10) we get

$$\begin{aligned} 0<2t_{0}=2m_{p}T<qT=t<\max (q+1,m_{k})T=t_{2}<\frac{1}{4\ln (\varGamma )}={\widetilde{t}}. \end{aligned}$$

Further, we deduce from (11) that

$$\begin{aligned} t_{0}=m_{p}T\le \min \Bigl [K_{1},K_{2} \sqrt{\frac{\tau -C_{0}\tau _{0}}{\tau _{2}-C_{0}\tau _{0}}}\Bigr ]. \end{aligned}$$

Let \(r\ge 0\) and \(f\in \mathcal {OV}({\mathbb {C}})\) such that \(|f|_{q}=\Vert f\Vert _{a_{q},q}\le 1\). We set \({\widetilde{f}}:S_{qT}\rightarrow {\mathbb {C}}\), \({\widetilde{f}}(z):=f(z/T)\), and define

$$\begin{aligned} H_{\tau }^{\sim }(S_{t}):=\{g\in {\mathcal {O}}(S_{t})\;|\;\Vert g\Vert _{\tau ,t}^{\sim }:=\sup _{z\in S_{t}}|g(z)|e^{\tau {\widetilde{\mu }}(z)}<\infty \} \end{aligned}$$

where \({\widetilde{\mu }}:=\mu (\cdot /T)\). We note that for \({\widetilde{n}}:=\lceil 1/T\rceil \), where \(\lceil \cdot \rceil \) is the ceiling function, there is \(C>0\) such that for all \(x\ge 0\)

$$\begin{aligned} {\widetilde{\mu }}(x+1)=\mu \Bigl (\frac{x+1}{T}\Bigr ) \le \mu \Bigl (\frac{x}{T}+\Bigl \lceil \frac{1}{T}\Bigr \rceil \Bigr ) =\mu \Bigl (\frac{x}{T}+{\widetilde{n}}\Bigr ) \le \varGamma \mu \Bigl (\frac{x}{T}\Bigr )+C =\varGamma {\widetilde{\mu }}(x)+C \end{aligned}$$

because \(\mu \) is a strong weight generator. We conclude that \({\widetilde{\mu }}\) is also a weight generator with the same \(\varGamma \) as \(\mu \) which is independent of T. Moreover, from

$$\begin{aligned} \Vert {\widetilde{f}}\Vert _{\tau ,t}^{\sim }=\sup _{z\in S_{qT}}|{\widetilde{f}}(z)|e^{a_{q}{\widetilde{\mu }}(z)} =\sup _{z\in S_{q}}|f(z)|e^{a_{q}\mu (z)}=|f|_{q}\le 1 \end{aligned}$$

it follows by Theorem 9 that there are \({\widetilde{f}}_{j}\in {\mathcal {O}}(S_{t_{j}})\), \(j\in \{0,2\}\), such that

$$\begin{aligned} {\widetilde{f}}(z)={\widetilde{f}}_{0}(z)+{\widetilde{f}}_{2}(z),\quad z\in S_{t_{0}}, \end{aligned}$$
(12)

and

$$\begin{aligned} C_{1}e^{-Gr} \ge \Vert {\widetilde{f}}_{0}\Vert _{C_{0}\tau _{0},t_{0}}^{\sim } =\sup _{z\in S_{t_{0}/T}}|\underbrace{{\widetilde{f}}_{0}(Tz)}_{=:f_{0}(z)} |e^{C_{0}\tau _{0}{\widetilde{\mu }}(Tz)} =\sup _{z\in S_{m_{p}}}|f_{0}(z)|e^{a_{m_{p}}\mu (z)}=|f_{0}|_{m_{p}},\nonumber \\ \end{aligned}$$
(13)

where \(f_{0}\in {\mathcal {O}}(S_{m_{p}})\), as well as

$$\begin{aligned} e^{r} \ge \Vert {\widetilde{f}}_{2}\Vert _{\tau _{2},t_{2}}^{\sim } =\sup _{z\in S_{t_{2}/T}}|\underbrace{{\widetilde{f}}_{2}(Tz)}_{=:f_{2}(z)}|e^{\tau _{2}{\widetilde{\mu }}(Tz)} \ge \sup _{z\in S_{m_{k}}}|f_{2}(z)|e^{a_{m_{k}}\mu (z)}=|f_{2}|_{m_{k}} \end{aligned}$$
(14)

where \(f_{2}\in {\mathcal {O}}(S_{t_{2}/T})\subset {\mathcal {O}}(S_{m_{k}})\) and the inclusion is justified by the identity theorem. Furthermore, for \(z\in S_{t_{0}/T}=S_{m_{p}}\) the equation

$$\begin{aligned} f(z)={\widetilde{f}}(Tz)\underset{(12)}{=}{\widetilde{f}}_{0} (Tz)+{\widetilde{f}}_{2}(Tz)=f_{0}(z)+f_{2}(z) \end{aligned}$$

holds, thus \(f=f_{0}+f_{2}\) on \(S_{m_{p}}\). By virtue of the weak reducibility of \(\mathcal {OV}({\mathbb {C}})\) and the choice of \(m_{p},m_{k}\) the following is valid:

$$\begin{aligned} \forall \;\varepsilon >0\;\exists \;{\widehat{f}}_{0},\, {\widehat{f}}_{2}\in \mathcal {OV}({\mathbb {C}}):\; (i)\;\;|{\widehat{f}}_{0}-f_{0}|_{p}<\varepsilon \quad \text {and}\quad (ii)\;\;|{\widehat{f}}_{2}-f_{2}|_{k}<\varepsilon . \end{aligned}$$
(15)

Now, we have to consider two cases. Let \(\varepsilon :=C_{1}e^{-Gr}\). For \(k\le p\) we get via (15) (i) that \( f={\widehat{f}}_{0}+(f_{2}+f_{0}-{\widehat{f}}_{0}) \) on \(S_{m_{p}}\) so

$$\begin{aligned} f_{2}+f_{0}-{\widehat{f}}_{0}=f-{\widehat{f}}_{0}=:{\overline{f}}_{2} \quad \text {on}\;S_{m_{p}} \end{aligned}$$
(16)

where the function \({\overline{f}}_{2}\in \mathcal {OV}({\mathbb {C}})\) and thus is a holomorphic extension of the left-hand side on \({\mathbb {C}}\). Hence we clearly have \(f={\widehat{f}}_{0}+{\overline{f}}_{2}\) and

$$\begin{aligned} |{\widehat{f}}_{0}|_{p} \le |{\widehat{f}}_{0}-f_{0}|_{p}+|f_{0}|_{p} \underset{(15) (i)}{\le }\varepsilon +|f_{0}|_{p} \le \varepsilon +|f_{0}|_{m_{p}} \underset{(13)}{\le }2C_{1}e^{-Gr}=:C_{2}e^{-Gr} \end{aligned}$$
(17)

as well as

(18)

Analogously, for \(k>p\) we obtain via (15) (ii) that \( f={\widehat{f}}_{2}+(f_{0}+f_{2}-{\widehat{f}}_{2}) \) on \(S_{m_{p}}\) so

$$\begin{aligned} f_{0}+f_{2}-{\widehat{f}}_{2}=f-{\widehat{f}}_{2} =:{\overline{f}}_{0}\quad \text {on}\;S_{m_{p}} \end{aligned}$$
(19)

where the function \({\overline{f}}_{0}\in \mathcal {OV}({\mathbb {C}})\) and thus is a holomorphic extension of the left-hand side on \({\mathbb {C}}\). Hence we clearly have \(f={\overline{f}}_{0}+{\widehat{f}}_{2}\) and

(20)

as well as

$$\begin{aligned} |{\widehat{f}}_{2}|_{k} \le |{\widehat{f}}_{2}-f_{2}|_{k}+|f_{2}|_{k} \underset{(15) (ii)}{\le }\varepsilon +|f_{2}|_{m_{k}} \underset{(14)}{\le } C_{1}e^{-Gr}+e^{r}\le C_{3}e^{r}. \end{aligned}$$
(21)

Next, we set \(n:=\lceil 1/G \rceil \) and \(C:=C_{3}e^{\ln (C_{2})/G}\). Let \({\widetilde{r}}>0\). For \({\widetilde{r}}\ge 1\) there is \(r\ge 0\) such that

$$\begin{aligned} {\widetilde{r}}=e^{Gr-\ln (C_{2})}=\frac{e^{Gr}}{C_{2}} \end{aligned}$$

and we have by (17) and (18) for \(k\le p\)

$$\begin{aligned} |{\widehat{f}}_{0}|_{p}\le C_{2}e^{-Gr}=\frac{1}{{\widetilde{r}}}, \quad |{\overline{f}}_{2}|_{k} \le C_{3}e^{r}=C_{3}e^{\frac{1}{G}\ln (C_{2})}e^{\frac{1}{G}(Gr-\ln (C_{2}))} =C\,{\widetilde{r}}^{\,\frac{1}{G}}\underset{{\widetilde{r}}\ge 1}{\le }C\,{\widetilde{r}}^{\;n}, \end{aligned}$$

as well as by (20) and (21) for \(k>p\)

$$\begin{aligned} |{\overline{f}}_{0}|_{p}\le \frac{1}{{\widetilde{r}}},\quad |{\widehat{f}}_{2}|_{k}\le C\,{\widetilde{r}}^{\;n}. \end{aligned}$$

For \(0<{\widetilde{r}}<1\) we have, since \(q\ge p\),

$$\begin{aligned} |f|_{p}\le |f|_{q}\le 1<\frac{1}{{\widetilde{r}}}. \end{aligned}$$

Thus our statement is proved. \(\square \)

Let us remark that the choice of the sequence \((a_{n})_{n\in \mathbb {N}}\) in the preceding theorem does not really matter.

Remark 11

Let \(\mu :{\mathbb {C}}\rightarrow [0,\infty )\) be continuous, \(a_{n}\nearrow 0\) or \(a_{n}\nearrow \infty \), \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) and \(\varOmega _{n}:=S_{n}\) for all \(n\in \mathbb {N}\). Set \({\mathcal {V}}_{-}:=(\exp ((-1/n)\mu ))_{n\in \mathbb {N}}\) and \({\mathcal {V}}_{+}:=(\exp (n\mu ))_{n\in \mathbb {N}}\). Then

$$\begin{aligned} \mathcal {OV}({\mathbb {C}})\cong \mathcal {OV}_{-}({\mathbb {C}}),\quad \text {if}\;a_{n}\nearrow 0,\quad \text {and}\quad \mathcal {OV}({\mathbb {C}})\cong \mathcal {OV}_{+}({\mathbb {C}}),\quad \text {if}\;a_{n}\nearrow \infty , \end{aligned}$$

which is easily seen. Thus one may choose the most suitable sequence \((a_{n})_{n\in \mathbb {N}}\) for one’s purpose without changing the space.

Let us turn to the third part. The following quite technical conditions guarantee a kind of weak reducibility of the projective limit \(\mathcal {EV}(\varOmega )\) and in combination with (PN.1) the weak reducibility of \(\mathcal {OV}(\varOmega )\) too.

Condition (WR)

Let \({\mathcal {V}}:=(\nu _{n})_{n\in \mathbb {N}}\) be a directed family of continuous weights on an open set \(\varOmega \subset \mathbb {R}^{2}\) and \((\varOmega _{n})_{n\in \mathbb {N}}\) a family of non-empty open sets such that \(\varOmega _{n}\ne \mathbb {R}^{2}\), \(\varOmega _{n}\subset \varOmega _{n+1}\) for all \(n\in \mathbb {N}\), \(\mathrm {d}_{n,k}:=\mathrm {d}^{|\cdot |}(\varOmega _{n},\partial \varOmega _{k})>0\) for all \(n,k\in \mathbb {N}\), \(k>n\), and \(\varOmega =\bigcup _{n\in \mathbb {N}} \varOmega _{n}\).

(WR.1) For every \(n\in \mathbb {N}\) let there be \(g_{n}\in {\mathcal {O}}({\mathbb {C}})\) with \(g_{n}(0)=1\) and \(\mathbb {N}\ni I_{j}(n)> n\) such that

  1. (a)

    for every \(\varepsilon >0\) there is a compact set \(K\subset {\overline{\varOmega }}_{n}\) with \(\nu _{n}(x)\le \varepsilon \nu _{I_{1}(n)}(x)\) for all \(x\in \varOmega _{n}{\setminus } K\).

  2. (b)

    there is an open set \(X_{I_{2}(n)}\subset \mathbb {R}^{2}{\setminus } {\overline{\varOmega }}_{I_{2}(n)}\) such that there are \(R_{n},r_{n}\in \mathbb {R}\) with \(0<2R_{n}<\mathrm {d}^{|\cdot |}(X_{I_{2}(n)},\varOmega _{I_{2}(n)}):=\mathrm {d}_{X,I_{2}(n)}\) and \(R_{n}<r_{n}<\mathrm {d}_{X,I_{2}(n)}-R_{n}\) as well as \(A_{2}(\cdot ,n):X_{I_{2}(n)}+\mathbb {B}_{R_{n}}(0)\rightarrow (0,\infty )\), \(A_{2}(\cdot ,n)_{\mid X_{I_{2}(n)}}\) locally bounded, satisfying

    $$\begin{aligned} \max \{|g_{n}(\zeta )|\nu _{I_{2}(n)}(z)\;|\;\zeta \in \mathbb {R}^{2},\,|\zeta -(z-x)|=r_{n}\}\le A_{2}(x,n) \end{aligned}$$

    for all \(z\in \varOmega _{I_{2}(n)}\) and \(x\in X_{I_{2}(n)}+\mathbb {B}_{R_{n}}(0)\).

  3. (c)

    for every compact set \(K\subset \mathbb {R}^{2}\) there is \(A_{3}(n,K)>0\) with

    $$\begin{aligned} \int _{K}{\frac{|g_{n}(x-y)|\nu _{n}(x)}{|x-y|}\mathrm {d}y}\le A_{3}(n,K),\quad x\in \varOmega _{n}. \end{aligned}$$

(WR.2) Let (WR.1a) be fulfilled. For every \(n\in \mathbb {N}\) let there be \(\mathbb {N}\ni I_{4}(n)>n\) and \(A_{4}(n)>0\) such that

$$\begin{aligned} \int _{\varOmega _{I_{4}(n)}}{\frac{|g_{I_{14}(n)}(x-y)|\nu _{p}(x)}{|x-y|\nu _{k}(y)}\mathrm {d}y}\le A_{4}(n), \quad x\in \varOmega _{p}, \end{aligned}$$

for \((k,p)=(I_{4}(n),n)\) and \((k,p)=(I_{14}(n),I_{14}(n))\) where \(I_{14}(n):=I_{1}(I_{4}(n))\).

(WR.3) Let (WR.1a), (WR.1b) and (WR.2) be fulfilled. For every \(n\in \mathbb {N}\), every closed subset \(M\subset {\overline{\varOmega }}_{n}\) and every component N of \(M^{C}\) we have

$$\begin{aligned} N\cap {\overline{\varOmega }}_{n}^{C}\ne \varnothing \;\Rightarrow \; N\cap X_{I_{214}(n)}\ne \varnothing \end{aligned}$$

where \(I_{214}(n):=I_{2}(I_{14}(n))\), fulfilling \(I_{214}(n)\ge I_{14}(n+1)\).

(WR) is [33, 4.2 Condition, p. 10] combined with the assumption \(I_{214}(n)\ge I_{14}(n+1)\), \(n\in \mathbb {N}\). We will see that \(\varOmega _{n}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<n\}\) and \(\nu _{n}(z):=\exp (a_{n}|{{\,\mathrm{Re}\,}}(z)|^{\gamma })\) for some \(0<\gamma \le 1\) and \(a_{n}\nearrow 0\) or \(a_{n}\nearrow \infty \) fulfil the conditions above with \(g_{n}(z):=\exp (-z^2)\).

Theorem 12

[33, 4.3 Theorem, p. 10] Let \(n\in \mathbb {N}\). Then \(\pi _{I_{214}(n),n}({\mathcal {E}}\nu _{I_{214}(n),{\overline{\partial }}} (\varOmega _{I_{214}(n)}))\) is dense in \(\pi _{I_{14}(n),n}({\mathcal {E}}\nu _{I_{14}(n), {\overline{\partial }}}(\varOmega _{I_{14}(n)}))\) w.r.t. \((|\cdot |_{n,m})_{m\in \mathbb {N}_{0}}\) if (WR) is fulfilled.

As a consequence of this theorem, whose proof does not need the assumption \(I_{214}(n)\ge I_{14}(n+1)\), we obtain that the projective limit \(\mathcal {OV}(\varOmega )\) is weakly reduced, which is a generalisation of [32, 5.6 Corollary, p. 69] and [32, 5.11 Corollary, p. 75].

Corollary 13

\(\mathcal {OV}(\varOmega )\) is weakly reduced if (WR) and (PN.1) are satisfied.

Proof

Let \(n\in \mathbb {N}\). We show that \(\pi _{n}(\mathcal {OV}(\varOmega ))\) is dense in \(\pi _{2J_{1}I_{14}(n),n}({\mathcal {O}} \nu _{2J_{1}I_{14}(n)}(\varOmega _{2J_{1}I_{14}(n)}))\) w.r.t. \(|\cdot |_{n}\) where \(J_{1}I_{14}(n):=J_{1}(I_{14}(n))\) and

$$\begin{aligned} \pi _{n}:\mathcal {OV}(\varOmega )\rightarrow {\mathcal {O}}\nu _{n} (\varOmega _{n}),\;\pi _{n}(f):=f_{\mid \varOmega _{n}}. \end{aligned}$$

We omit the restriction maps in our proof. Due to Proposition 7 (a) the restrictions to \(\varOmega _{I_{14}(n)}\) of functions from \({\mathcal {O}}\nu _{2J_{1}I_{14}(n)}(\varOmega _{2J_{1}I_{14}(n)})\) are elements of \({\mathcal {E}}\nu _{I_{14}(n),{\overline{\partial }}}(\varOmega _{I_{14}(n)})\). Let \(\varepsilon >0\) and \(f_{0}\in {\mathcal {O}}\nu _{2J_{1}I_{14}(n)}(\varOmega _{2J_{1}I_{14}(n)})\). For every \(j\in \mathbb {N}\) there exists

  1. (i)

    \(f_{j}\in {\mathcal {E}}\nu _{I_{214}(n+j-1),{\overline{\partial }}}(\varOmega _{I_{214}(n+j-1)})\) with

  2. (ii)

    \({f_{j}}_{\mid \varOmega _{I_{14}(n+j)}}\in {\mathcal {E}}\nu _{I_{14}(n+j),{\overline{\partial }}}(\varOmega _{I_{14}(n+j)}) \subset {\mathcal {O}}\nu _{I_{14}(n+j)}(\varOmega _{I_{14}(n+j)})\)

such that

$$\begin{aligned} |f_{j}-f_{j-1}|_{n+j-1}=|f_{j}-f_{j-1}|_{n+j-1,0}<\frac{\varepsilon }{2^{j+1}} \end{aligned}$$
(22)

by Theorem 12 and the condition \(I_{214}(k)\ge I_{14}(k+1)\) for all \(k\in \mathbb {N}\) from (WR). Therefore we obtain for every \(k\in \mathbb {N}\)

(23)

Now, let \(\varepsilon _{0}>0\) and \(l\in \mathbb {N}\). We choose \(l_{0}\in \mathbb {N}\), \(l_{0}\ge l\), such that \(\frac{\varepsilon }{2^{l_{0}+1}}<\varepsilon _{0}\). Similarly, we get for all \(p\ge k\ge l_{0}\)

Hence \((f_{k})_{k\ge n_{0}}\) is a Cauchy sequence in the Banach space \({\mathcal {O}}\nu _{I_{14}(n+n_{0})}(\varOmega _{I_{14}(n+n_{0})})\) for every \(n_{0}\in \mathbb {N}_{0}\) and thus has a limit \(F_{n_{0}}\in {\mathcal {O}}\nu _{I_{14}(n+n_{0})}(\varOmega _{I_{14}(n+n_{0})})\). These limits coincide on their common domain because for every \(n_{1},n_{2}\in \mathbb {N}_{0}\) with \(I_{14}(n+n_{1})<I_{14}(n+n_{2})\) and \(\varepsilon _{1}>0\) there exists \(N\in \mathbb {N}\) such that for all \(k\ge N\)

$$\begin{aligned} |F_{n_{1}}-F_{n_{2}}|_{I_{14}(n+n_{1})}&\le |F_{n_{1}}-f_{k}|_{I_{14}(n+n_{1})}+|f_{k}-F_{n_{2}}|_{I_{14}(n+n_{1})}\\&\le |F_{n_{1}}-f_{k}|_{I_{14}(n+n_{1})}+|f_{k}-F_{n_{2}}|_{I_{14}(n+n_{2})} <\frac{\varepsilon _{1}}{2}+\frac{\varepsilon _{1}}{2}=\varepsilon _{1}. \end{aligned}$$

We deduce that the glued limit function f given by \(f:=F_{n_{0}}\) on \(\varOmega _{I_{14}(n+n_{0})}\) for all \(n_{0}\in \mathbb {N}_{0}\) is well-defined and we have \(f\in \bigcap _{n_{0}\in \mathbb {N}_{0}}{\mathcal {O}}\nu _{I_{14}(n+n_{0})}(\varOmega _{I_{14}(n+n_{0})}) =\mathcal {OV}(\varOmega )\) since \(I_{14}(n+n_{0})\ge n+n_{0}\). By the definition of f there exists \(N\in \mathbb {N}\) such that for every \(k\ge N\)

$$\begin{aligned} |f-f_{0}|_{n}\le |f-f_{k}|_{n}+|f_{k}-f_{0}|_{n} \underset{n\le I_{14}(n+0)}{<}\frac{\varepsilon }{2}+|f_{k}-f_{0}|_{n} \underset{(23)}{\le }\frac{\varepsilon }{2}+\frac{\varepsilon }{2}=\varepsilon , \end{aligned}$$

which proves our statement. \(\square \)

Combining Theorem 10 and Corollary 13, we obtain the following corollary.

Corollary 14

Let \(a_{n}\nearrow 0\) or \(a_{n}\nearrow \infty \), \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) and \(\varOmega _{n}:=S_{n}\) for all \(n\in \mathbb {N}\) where

$$\begin{aligned} \mu :{\mathbb {C}}\rightarrow [0,\infty ),\;\mu (z):=|{{\,\mathrm{Re}\,}}(z)|^{\gamma }, \end{aligned}$$

for some \(0<\gamma \le 1\). Then \(\mathcal {OV}({\mathbb {C}})\) satisfies \((\varOmega )\).

Proof

We only need to check that the conditions of Theorem 10 are fulfilled. Obviously, \(\mu (z)=\mu (|{{\,\mathrm{Re}\,}}(z)|)\) for all \(z\in {\mathbb {C}}\), \(\mu \) is strictly increasing on \([0,\infty )\) and \(\lim _{x\rightarrow \infty ,\, x\in \mathbb {R}}\frac{\ln (1+|x|)}{\mu (x)}=0\). The observation

$$\begin{aligned} \mu (x+n)-\mu (x)=|x+n|^{\gamma }-|x|^{\gamma }\le |x+n-n|^{\gamma }=n^{\gamma },\quad n\in \mathbb {N},\;x\in [0,\infty ), \end{aligned}$$

implies that \(\mu \) is a strong weight generator with any \(\varGamma >1\) and \(C:=n^{\gamma }\) by Definition 8. Let us turn to the conditions (WR) and (PN.1) which we need for the weak reducibility of \(\mathcal {OV}({\mathbb {C}})\) by Corollary 13. Condition (PN.1) is fulfilled by Example 6. If \(a_{n}<0\) for all \(n\in \mathbb {N}\), then (WR) is fulfilled by [33, 4.10 Example a), p. 22] where we used \({\widetilde{\mu }}(z):=|z|^{\gamma }\) instead of \(\mu \), which does not make a difference since

$$\begin{aligned} |{{\,\mathrm{Re}\,}}(z)|^{\gamma }\le |z|^{\gamma }\le |{{\,\mathrm{Re}\,}}(z)|^{\gamma }+n^{\gamma },\quad z\in \varOmega _{n}=S_{n}. \end{aligned}$$

If \(a_{n}\ge 0\) for all \(n\in \mathbb {N}\), we only have to modify [33, 4.10 Example a), p. 22] a bit. We choose \(I_{j}(n):=2n\) for \(j\in \{1,2,4\}\) and define the open set \(X_{I_{2}(n)}:={\overline{S}}_{4n}^{C}\). Then we have

$$\begin{aligned} I_{214}(n)=8n\ge 4n+4=I_{14}(n+1),\quad n\in \mathbb {N}. \end{aligned}$$

Furthermore, we have \(\mathrm {d}_{n,k}=|n-k|\) for all \(n,k\in \mathbb {N}\).

(WR.1a) and (WR.3): Verbatim as in [33, 4.10 Example a), p. 22].

(WR.1b): We have \(\mathrm {d}_{X,I_{2}}=2n\). We choose \(g_{n}:{\mathbb {C}}\rightarrow {\mathbb {C}}\), \(g_{n}(z):=\exp (-z^2)\), as well as \(r_{n}:=1/(4n)\) and \(R_{n}:=1/(6n)\) for \(n\in \mathbb {N}\). Let \(z=z_{1}+iz_{2}\in \varOmega _{I_{2}(n)}=S_{2n}\) and \(x\in X_{I_{2}(n)}+\mathbb {B}_{R_{n}}(0)\). For \(\zeta =\zeta _{1}+i\zeta _{2}\in {\mathbb {C}}\) with \(|\zeta -(z-x)|=r_{n}\) we have

$$\begin{aligned} |g_{n}(\zeta )|e^{a_{2n}\mu (z)}&=e^{-{{\,\mathrm{Re}\,}}(\zeta ^{2})}e^{a_{2n}|{{\,\mathrm{Re}\,}}(z)|^{\gamma }} \le e^{-\zeta _{1}^{2}+\zeta _{2}^{2}}e^{a_{2n}(1+|z_{1}|)}\\&\le e^{(r_{n}+|z_{2}|+|x_{2}|)^{2}+a_{2n}(1+r_{n}+|x_{1}|)}e^{-|\zeta _{1}|^{2}+a_{2n}|\zeta _{1}|}\\&\le e^{(r_{n}+2n+|x_{2}|)^{2}+a_{2n}(1+r_{n}+|x_{1}|)}\sup _{t\in \mathbb {R}}e^{-t^{2}+a_{2n}t}\\&= e^{(r_{n}+2n+|x_{2}|)^{2}+a_{2n}(1+r_{n}+|x_{1}|)+a_{2n}^{2}/4}=:A_{2}(x,n) \end{aligned}$$

and observe that \(A_{2}(\cdot ,n)\) is continuous and thus locally bounded on \(X_{I_{2}(n)}\).

(WR.1c): Let \(K\subset {\mathbb {C}}\) be compact and \(x=x_{1}+ix_{2}\in \varOmega _{n}\). Then there is \(b>0\) such that \(|y|\le b\) for all \(y=y_{1}+iy_{2}\in K\) and from polar coordinates and Fubini’s theorem it follows that

$$\begin{aligned}&\int _{K}\frac{|g_{n}(x-y)|}{|x-y|}\mathrm {d}y\\&\quad \le \underbrace{\sup _{w\in K}e^{a_{2n}|{{\,\mathrm{Re}\,}}(w)|}}_{=:C_{1}} \int _{K}\frac{e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}}{|x-y|}e^{-a_{2n}|y_{1}|}\mathrm {d}y\\&\quad \le C_{1}\bigl ( \int _{\mathbb {B}_{1}(x)}\frac{e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}}{|x-y|}e^{-a_{2n}|{{\,\mathrm{Re}\,}}(y)|}\mathrm {d}y +\int _{K{\setminus } \mathbb {B}_{1}(x)}\frac{e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}}{|x-y|}e^{-a_{2n}|{{\,\mathrm{Re}\,}}(y)|}\mathrm {d}y\bigr )\\&\quad \le C_{1}\bigl (\int _{0}^{2\pi }\int _{0}^{1}\frac{e^{-r^{2}\cos (2\varphi )}}{r}e^{-a_{2n}|x_{1}+r\cos (\varphi )|}r\mathrm {d}r\mathrm {d}\varphi + \int _{K{\setminus } \mathbb {B}_{1}(x)}e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}e^{-a_{2n}|{{\,\mathrm{Re}\,}}(y)|}\mathrm {d}y\bigr )\\&\quad \le C_{1}\bigl (2\pi e^{1+a_{2n}}e^{-a_{2n}|x_{1}|} +\int _{-b}^{b}e^{(x_{2}-y_{2})^{2}}\mathrm {d}y_{2} \int _{\mathbb {R}}e^{-(x_{1}-y_{1})^{2}+a_{2n}|x_{1}-y_{1}|}\mathrm {d}y_{1}e^{-a_{2n}|x_{1}|}\bigr )\\&\quad \le C_{1}\bigl (2\pi e^{1+a_{2n}}+2be^{(|x_{2}|+b)^{2}}\int _{\mathbb {R}}e^{-y_{1}^{2}+a_{2n}|y_{1}|}\mathrm {d}y_{1}\bigr )e^{-a_{2n}|x_{1}|}\\&\quad = C_{1}\bigl (2\pi e^{1+a_{2n}}+2be^{(|x_{2}|+b)^{2}}e^{a_{2n}^{2}/4} \int _{\mathbb {R}}e^{-(|y_{1}|-a_{2n}/2)^{2}}\mathrm {d}y_{1}\bigr )e^{-a_{2n}|x_{1}|}\\&\quad = C_{1}\bigl (2\pi e^{1+a_{2n}}+4be^{(|x_{2}|+b)^{2}}e^{a_{2n}^{2}/4} \int _{-a_{2n}/2}^{\infty }e^{-y_{1}^{2}}\mathrm {d}y_{1}\bigr )e^{-a_{2n}|x_{1}|}\\&\quad \le C_{1}\bigl (2\pi e^{1+a_{2n}}+4\sqrt{\pi }be^{(n+b)^{2}+a_{2n}^{2}/4}\bigr )e^{-a_{2n}|x_{1}|}. \end{aligned}$$

We conclude that (WR.1c) holds since

$$\begin{aligned} e^{-a_{2n}|x_{1}|}e^{a_{n}|{{\,\mathrm{Re}\,}}(x)|^{\gamma }}\le e^{(a_{n}-a_{2n})|x_{1}|+a_{n}}\le e^{a_{n}}. \end{aligned}$$

(WR.2): Let \(p,k\in \mathbb {N}\) with \(p\le k\). For all \(x=x_{1}+ix_{2}\in \varOmega _{p}\) and \(y=y_{1}+iy_{2}\in \varOmega _{I_{4}(n)}\) we note that

$$\begin{aligned} a_{p}|{{\,\mathrm{Re}\,}}(x)|^{\gamma }-a_{k}|{{\,\mathrm{Re}\,}}(y)|^{\gamma } \le a_{k}|x_{1}-y_{1}|^{\gamma }\le a_{k}(1+|x_{1}-y_{1}|) \end{aligned}$$

because \((a_{n})_{n\in \mathbb {N}}\) is non-negative and increasing and \(0<\gamma \le 1\). Like before we deduce that

$$\begin{aligned}&\int _{\varOmega _{I_{4}(n)}}\frac{|g_{n}(x-y)|\nu _{p}(x)}{|x-y|\nu _{k}(y)}\mathrm {d}y\\&=\int _{\varOmega _{2n}}\frac{e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}}{|x-y|}e^{a_{p}|{{\,\mathrm{Re}\,}}(x)|^{\gamma }-a_{k}|{{\,\mathrm{Re}\,}}(y)|^{\gamma }}\mathrm {d}y \le \int _{\varOmega _{2n}}\frac{e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}}{|x-y|}e^{a_{k}|{{\,\mathrm{Re}\,}}(x)-{{\,\mathrm{Re}\,}}(y)|^{\gamma }}\mathrm {d}y\\&\le \int _{0}^{2\pi }\int _{0}^{1}\frac{e^{-r^{2}\cos (2\varphi )}}{r}e^{a_{k}r^{\gamma }}r\mathrm {d}r\mathrm {d}\varphi + \int _{\varOmega _{2n}{\setminus } \mathbb {B}_{1}(x)}e^{-{{\,\mathrm{Re}\,}}((x-y)^{2})}e^{a_{k}|{{\,\mathrm{Re}\,}}(x)-{{\,\mathrm{Re}\,}}(y)|^{\gamma }}\mathrm {d}y\\&\le 2\pi e^{1+a_{k}}+e^{a_{k}}\int _{-2n}^{2n}e^{(x_{2}-y_{2})^{2}}\mathrm {d}y_{2} \int _{\mathbb {R}}e^{-(x_{1}-y_{1})^{2}+a_{k}|x_{1}-y_{1}|}\mathrm {d}y_{1}\\&\le 2\pi e^{1+a_{k}}+8\sqrt{\pi }ne^{a_{k}+(|x_{2}|+2n)^{2}+a_{k}^{2}/4}\\&\le 2\pi e^{1+a_{I_{14}(n)}}+8\sqrt{\pi }ne^{a_{I_{14}(n)}+(I_{14}(n)+2n)^{2}+a_{I_{14}(n)}^{2}/4} \end{aligned}$$

for \((k,p)=(I_{4}(n),n)\) and \((k,p)=(I_{14}(n),I_{14}(n))\) as \((a_{n})_{n\in \mathbb {N}}\) is non-negative and increasing. \(\square \)

We close this section with a special case of our main result on the surjectivity of the Cauchy–Riemann operator on \(\mathcal {EV}(\varOmega ,E)\). We recall the corresponding result for \(E={\mathbb {C}}\) which we will need for the application of our main result. It is a consequence of the approximation Theorem 12 in combination with Hörmander’s solution of the \({\overline{\partial }}\)-problem in weighted \(L^{2}\)-spaces [27, Theorem 4.4.2, p. 94] and the Mittag–Leffler procedure.

Theorem 15

[33, 4.8 Theorem, p. 20] Let (PN) with \(\psi _{n}(z):=(1+|z|^{2})^{-2}\), \(z\in \varOmega \), and (WR) be fulfilled and \(-\ln \nu _{n}\) be subharmonic on \(\varOmega \) for every \(n\in \mathbb {N}\). Then

$$\begin{aligned} {\overline{\partial }}:\mathcal {EV}(\varOmega )\rightarrow \mathcal {EV}(\varOmega ) \end{aligned}$$

is surjective.

A consequence of this theorem is the following corollary.

Corollary 16

[33, 4.10 Example a), p. 22] Let \((a_{n})_{n\in \mathbb {N}}\) be strictly increasing, \(a_{n}<0\) for all \(n\in \mathbb {N}\), \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) and \(\varOmega _{n}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<n\}\) for all \(n\in \mathbb {N}\) where

$$\begin{aligned} \mu :{\mathbb {C}}\rightarrow [0,\infty ),\;\mu (z):=|{{\,\mathrm{Re}\,}}(z)|^{\gamma }, \end{aligned}$$

for some \(0<\gamma \le 1\). Then

$$\begin{aligned} {\overline{\partial }}:\mathcal {EV}({\mathbb {C}})\rightarrow \mathcal {EV}({\mathbb {C}}) \end{aligned}$$

is surjective.

The restriction to negative \(a_{n}\) comes from the condition that \(-\ln \nu _{n}\) should be subharmonic. We note that the E-valued versions of Theorem 15 and Corollary 16 where E is a Fréchet space over \({\mathbb {C}}\) hold as well by the classical theory of tensor products for nuclear Fréchet spaces (see [33, 4.9 Corollary, p. 21]). Now, we use the results obtained so far to obtain a special case of our main result.

Corollary 17

Let \(\mu \) be a subharmonic strong weight generator and \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) with \(a_{n}\nearrow 0\). Let (PN) with \(\psi _{n}(z):=(1+|z|^{2})^{-2}\), \(z\in {\mathbb {C}}\), and (WR) with \(\varOmega _{n}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<n\}\) for all \(n\in \mathbb {N}\) be fulfilled. If

  1. (a)

    \(E:=F_{b}'\) where F is a Fréchet space over \({\mathbb {C}}\) satisfying (DN), or

  2. (b)

    E is an ultrabornological PLS-space over \({\mathbb {C}}\) satisfying (PA),

then

$$\begin{aligned} {\overline{\partial }}^{E}:\mathcal {EV}({\mathbb {C}},E)\rightarrow \mathcal {EV}({\mathbb {C}},E) \end{aligned}$$

is surjective.

Proof

The space \(\mathcal {EV}({\mathbb {C}})\) is nuclear, in particular Schwartz, by [37, Theorem 3.1, p. 188], [37, Remark 2.7, p. 178-179] and [37, Remark 2.3 (b), p. 177] because (PN.1) and \((PN.2)^{1}\) from (PN) are fulfilled. Hence the subspace \(\mathcal {EV}_{{\overline{\partial }}}({\mathbb {C}})=\mathcal {OV}({\mathbb {C}})\) is nuclear by Proposition 7 (b) as well. Further, \(\mathcal {OV}({\mathbb {C}})\) is weakly reduced by Corollary 13 due to (WR) and thus satisfies \((\varOmega )\) by Theorem 10. Therefore, the assertion is a consequence of the surjectivity of \({\overline{\partial }}\) in the \({\mathbb {C}}\)-valued case by Theorem 15 and our main result Theorem 5. \(\square \)

Corollary 17 generalises a part of [32, 5.24 Theorem, p. 95] (\(K=\varnothing \)) which is the case \(\gamma =1\) of the next corollary.

Corollary 18

Let \(a_{n}\nearrow 0\), \({\mathcal {V}}:=(\exp (a_{n}\mu ))_{n\in \mathbb {N}}\) and \(\varOmega _{n}:=\{z\in {\mathbb {C}}\;|\;|{{\,\mathrm{Im}\,}}(z)|<n\}\) for all \(n\in \mathbb {N}\) where

$$\begin{aligned} \mu :{\mathbb {C}}\rightarrow [0,\infty ),\;\mu (z):=|{{\,\mathrm{Re}\,}}(z)|^{\gamma }, \end{aligned}$$

for some \(0<\gamma \le 1\). If

  1. (a)

    \(E:=F_{b}'\) where F is a Fréchet space over \({\mathbb {C}}\) satisfying (DN), or

  2. (b)

    E is an ultrabornological PLS-space over \({\mathbb {C}}\) satisfying (PA),

then

$$\begin{aligned} {\overline{\partial }}^{E}:\mathcal {EV}({\mathbb {C}},E)\rightarrow \mathcal {EV}({\mathbb {C}},E) \end{aligned}$$

is surjective.

Proof

Follows from Corollary 17, (the proof of) Corollary 14 and Example 6. \(\square \)