This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A note on the Lumer–Phillips theorem for bi-continuous semigroups

Karsten Kruse 0000-0003-1864-4915 Hamburg University of Technology
Institute of Mathematics
Am Schwarzenberg-Campus 3
21073 Hamburg
Germany
karsten.kruse@tuhh.de
 and  Christian Seifert 0000-0001-9182-8687 christian.seifert@tuhh.de
(Date: March 16, 2025)
Abstract.

Given a Banach space XX and an additional coarser Hausdorff locally convex topology τ\tau on XX we characterise the generators of τ\tau-bi-continuous semigroups in the spirit of the Lumer–Phillips theorem, i.e. by means of dissipativity w.r.t. a directed system of seminorms and a range condition.

Key words and phrases:
dissipativity, bi-continuous semigroup, Lumer–Phillips theorem, Saks space, mixed topology
2020 Mathematics Subject Classification:
Primary 47B44 Secondary 47D06, 46A70
K. Kruse acknowledges the support by the Deutsche Forschungsgemeinschaft (DFG) within the Research Training Group GRK 2583 “Modeling, Simulation and Optimization of Fluid Dynamic Applications”.

1. Introduction

Characterising generators of strongly continuous semigroups on Banach spaces is a classical topic due to its relation to well-posedness of the corresponding abstract Cauchy problem [8, Chap. II, 6.7 Theorem, p. 150]. The two main generation theorems go back to Hille–Yosida [16, 34] and Feller–Miyadera–Phillips [14, 28, 29] for general semigroups and Lumer–Phillips [25] for contraction semigroups.

However, there are important examples of semigroups which are not strongly continuous for the Banach space norm, e.g. the Gauß–Weierstraß semigroup on Cb(d)\mathrm{C}_{\operatorname{b}}(\mathbb{R}^{d}). To circumvent this issue the concept of bi-continuous semigroups which are strongly continuous only w.r.t. to a weaker Hausdorff locally convex topology τ\tau has been introduced by Kühnemund [22]. Thus a natural question is to characterise the generators of bi-continuous semigroups in the spirit of the Hille–Yosida theorem and the Lumer–Phillips theorem. While a version of the Hille–Yosida theorem for bi-continuous semigroups was established directly at the beginning of the theory [23], a corresponding version of the Lumer–Phillips theorem for bi-continuous contraction semigroups was missing. Recently, in [5], Budde and Wegner introduced the notion of bi-dissipativity to characterise the generators of bi-continuous contraction semigroups. However, the result in [5] does not cover the key example of the Gauß–Weierstraß semigroup on Cb(d)\mathrm{C}_{\operatorname{b}}(\mathbb{R}^{d}) (see [5, Example 3.9, p. 8]), see also Remark 3.25 below.

In this paper we make use of the notion of Γ\Gamma-dissipativity, where Γ\Gamma is a directed set of seminorms generating a topology related to the mixed topology γγ(,τ)\gamma\coloneqq\gamma(\|\cdot\|,\tau) (see [33] and Definition 2.1 for the mixed topology), introduced in [1] to prove versions of the Lumer–Phillips theorem for bi-continuous (contraction) semigroups in Theorem 3.9 and Theorem 3.10. Note that also [5] used Γ\Gamma-dissipativity to define bi-dissipativity; however, there Γ\Gamma is a fundamental system of seminorms generating the topology τ\tau. Working instead with the mixed topology yields a more natural concept which can also be applied to the Gauß–Weierstraß semigroup, see Remark 3.25.

Let us mention that strongly continuous semigroups have also been considered in locally convex spaces, in particular including a Lumer–Phillips-type generation theorem, see [1].

Note that with these results we also answer a question on characterising generators of transition semigroups raised by Markus Kunze, see Problem 3 in [15, p. 4].

In Section 2 we review the notion of bi-continuous semigroups and their generators as well as the (sub-)mixed topology. We also recall further properties of locally convex spaces as well as operators which we make use of later on. In Section 3 we recall the notion of Γ\Gamma-dissipativity and then prove the version of the Lumer–Phillips theorem, where we also comment on its relation to [5]. Further, we also provide some examples.

2. Notions and preliminaries

In this short section we recall some basic notions and results in the context of bi-continuous semigroups. For a vector space XX over the field \mathbb{R} or \mathbb{C} with a Hausdorff locally convex topology τ\tau we denote by (X,τ)(X,\tau)^{\prime} the topological linear dual space and just write X(X,τ)X^{\prime}\coloneqq(X,\tau)^{\prime} if (X,τ)(X,\tau) is a Banach space. By Γτ\Gamma_{\tau} we always denote a directed system of continuous seminorms that generates the Hausdorff locally convex topology τ\tau on XX. Further, for two Hausdorff locally convex spaces (X,τ)(X,\tau) and (Y,σ)(Y,\sigma) we use the symbol ((X,τ);(Y,σ))\mathcal{L}((X,\tau);(Y,\sigma)) for the space of continuous linear operators from (X,τ)(X,\tau) to (Y,σ)(Y,\sigma), and abbreviate (X,τ)((X,τ);(X,τ))\mathcal{L}(X,\tau)\coloneqq\mathcal{L}((X,\tau);(X,\tau)). If (X,X)(X,\|\cdot\|_{X}) and (Y,Y)(Y,\|\cdot\|_{Y}) are Banach spaces, we denote by τX\tau_{\|\cdot\|_{X}} and τY\tau_{\|\cdot\|_{Y}} the corresponding topologies induced by the norms and just write (X;Y)((X,τX);(Y,τY))\mathcal{L}(X;Y)\coloneqq\mathcal{L}((X,\tau_{\|\cdot\|_{X}});(Y,\tau_{\|\cdot\|_{Y}})) with operator norm (X;Y)\|\cdot\|_{\mathcal{L}(X;Y)}, and (X)(X;X)\mathcal{L}(X)\coloneqq\mathcal{L}(X;X).

Let us recall the definition of the mixed topology, [33, Section 2.1], and the notion of a Saks space, [7, I.3.2 Definition, p. 27–28], which will be important for the rest of the paper.

2.1 Definition ([18, Definition 2.2, p. 3], [5, Proposition 3.11 (a), p. 9]).

Let (X,)(X,\|\cdot\|) be a Banach space and τ\tau a Hausdorff locally convex topology on XX that is coarser than the \|\cdot\|-topology τ\tau_{\|\cdot\|}. Then

  1. (a)

    the mixed topology γγ(,τ)\gamma\coloneqq\gamma(\|\cdot\|,\tau) is the finest linear topology on XX that coincides with τ\tau on \|\cdot\|-bounded sets and such that τγτ\tau\subseteq\gamma\subseteq\tau_{\|\cdot\|};

  2. (b)

    a directed system of continuous seminorms Γτ\Gamma_{\tau} that generates the topology τ\tau is called norming if

    x=suppΓτp(x),xX;\|x\|=\sup_{p\in\Gamma_{\tau}}p(x),\quad x\in X; (1)
  3. (c)

    the triple (X,,τ)(X,\|\cdot\|,\tau) is called a Saks space if there exists a norming directed system of continuous seminorms Γτ\Gamma_{\tau} that generates the topology τ\tau.

The mixed topology is actually Hausdorff locally convex and the definition given above is equivalent to the one introduced by Wiweger [33, Section 2.1] due to [33, Lemmas 2.2.1, 2.2.2, p. 51].

2.2 Definition ([20, Definitions 2.2, 5.4, p. 2, 8]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space.

  1. (a)

    We call (X,,τ)(X,\|\cdot\|,\tau) (sequentially) complete if (X,γ)(X,\gamma) is (sequentially) complete.

  2. (b)

    We call (X,,τ)(X,\|\cdot\|,\tau) semi-reflexive if (X,γ)(X,\gamma) is semi-reflexive.

  3. (c)

    We call (X,,τ)(X,\|\cdot\|,\tau) C-sequential if (X,γ)(X,\gamma) is C-sequential, i.e. every convex sequentially open subset of (X,γ)(X,\gamma) is already open (see [31, p. 273]).

2.3 Remark.

If (X,,τ)(X,\|\cdot\|,\tau) is a sequentially complete Saks space, then (X,,γ)(X,\|\cdot\|,\gamma) is also a sequentially complete Saks space by [21, Lemma 5.5, p. 2680–2681] and [18, Remark 2.3 (c), p. 3]. In particular, there exists a norming directed system of continuous seminorms Γγ\Gamma_{\gamma} that generates γ\gamma.

There is another kind of mixed topology (see [7, p. 41]) which becomes quite handy if one has to deal with the mixed topology because it is generated by a quite simple directed system of continuous seminorms and often coincides with the mixed topology.

2.4 Definition ([18, Definition 3.9, p. 9]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space and Γτ\Gamma_{\tau} a norming directed system of continuous seminorms that generates the topology τ\tau. We set

𝒩{(pn,an)n|(pn)nΓτ,(an)nc0,an0for all n}\mathcal{N}\coloneqq\{(p_{n},a_{n})_{n\in\mathbb{N}}\;|\;(p_{n})_{n\in\mathbb{N}}\subseteq\Gamma_{\tau},\,(a_{n})_{n\in\mathbb{N}}\in c_{0},\,a_{n}\geq 0\;\text{for all }n\in\mathbb{N}\}

where c0c_{0} is the space of real null-sequences. For (pn,an)n𝒩(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N} we define the seminorm

|x|(pn,an)nsupnpn(x)an,xX.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}\coloneqq\sup_{n\in\mathbb{N}}p_{n}(x)a_{n},\quad x\in X.

We denote by γsγs(,τ)\gamma_{s}\coloneqq\gamma_{s}(\|\cdot\|,\tau) the Hausdorff locally convex topology that is generated by the system of seminorms (||||||(pn,an)n)(pn,an)n𝒩({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}})_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}} and call it the submixed topology.

Due to [7, I.1.10 Proposition, p. 9], [7, I.4.5 Proposition, p. 41–42] and [11, Lemma A.1.2, p. 72] we have the following observation.

2.5 Remark ([18, Remark 3.10, p. 9]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space, Γτ\Gamma_{\tau} a norming directed system of continuous seminorms that generates the topology τ\tau, γγ(,τ)\gamma\coloneqq\gamma(\|\cdot\|,\tau) the mixed and γsγs(,τ)\gamma_{s}\coloneqq\gamma_{s}(\|\cdot\|,\tau) the submixed topology.

  1. (a)

    We have τγsγ\tau\subseteq\gamma_{s}\subseteq\gamma and γs\gamma_{s} has the same convergent sequences as γ\gamma.

  2. (b)

    If

    1. (i)

      for every xXx\in X, ε>0\varepsilon>0 and pΓτp\in\Gamma_{\tau} there are y,zXy,z\in X such that x=y+zx=y+z, p(z)=0p(z)=0 and yp(x)+ε\|y\|\leq p(x)+\varepsilon, or

    2. (ii)

      the \|\cdot\|-unit ball B={xX|x1}B_{\|\cdot\|}=\{x\in X\;|\;\|x\|\leq 1\} is τ\tau-compact,

    then γ=γs\gamma=\gamma_{s} holds.

The submixed topology γs\gamma_{s} was originally introduced in [33, Theorem 3.1.1, p. 62] where a proof of Remark 2.5 (b) can be found as well.

2.6 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space and Γτ\Gamma_{\tau} a norming directed system of continuous seminorms that generates the topology τ\tau. Then there is a norming directed system of continuous seminorms Γγs\Gamma_{\gamma_{s}} that generates the submixed topology γs=γs(,τ)\gamma_{s}=\gamma_{s}(\|\cdot\|,\tau). Indeed, we set

𝒩1{(pn,an)n|(pn)nΓτ,(an)nc0, 0an1for all n}\mathcal{N}_{1}\coloneqq\{(p_{n},a_{n})_{n\in\mathbb{N}}\;|\;(p_{n})_{n\in\mathbb{N}}\subseteq\Gamma_{\tau},\,(a_{n})_{n\in\mathbb{N}}\in c_{0},\,0\leq a_{n}\leq 1\;\text{for all }n\in\mathbb{N}\}

and Γγs{||||||(pn,an)n|(pn,an)n𝒩1}\Gamma_{\gamma_{s}}\coloneqq\{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}\;|\;(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}_{1}\}. Let (pn)nΓτ(p_{n})_{n\in\mathbb{N}}\subseteq\Gamma_{\tau} and (an)nc0(a_{n})_{n\in\mathbb{N}}\in c_{0} with an0a_{n}\geq 0 for all nn\in\mathbb{N}. Then Csupnan<C\coloneqq\sup_{n\in\mathbb{N}}a_{n}<\infty and w.l.o.g. C>0C>0. We have

|x|(pn,an)n=supnpn(x)anCsupnpn(x)anC=C|x|(pn,anC)n{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}=\sup_{n\in\mathbb{N}}p_{n}(x)a_{n}\leq C\sup_{n\in\mathbb{N}}p_{n}(x)\frac{a_{n}}{C}=C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\left(p_{n},\tfrac{a_{n}}{C}\right)_{n\in\mathbb{N}}}

for all xXx\in X. In combination with 𝒩1𝒩\mathcal{N}_{1}\subseteq\mathcal{N} this shows that Γγs\Gamma_{\gamma_{s}} generates γs\gamma_{s}. Furthermore, for every pΓτp\in\Gamma_{\tau} we have with pnpp_{n}\coloneqq p for all nn\in\mathbb{N} that p(x)|x|(pn,1/n)np(x)\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},1/n)_{n\in\mathbb{N}}} for all xXx\in X. Together with the norming property of Γτ\Gamma_{\tau} this implies

x=suppΓτp(x)sup(pn,an)n𝒩1|x|(pn,an)nsup(an)nc0, 0an1xanx\|x\|=\sup_{p\in\Gamma_{\tau}}p(x)\leq\sup_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}_{1}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}\leq\sup_{(a_{n})_{n\in\mathbb{N}}\in c_{0},\,0\leq a_{n}\leq 1}\|x\|a_{n}\leq\|x\|

for all xXx\in X. Hence Γγs\Gamma_{\gamma_{s}} is norming.

Recall that for a Banach space (X,)(X,\|\cdot\|) and a Hausdorff locally convex topology τ\tau on XX the triple (X,,τ)(X,\|\cdot\|,\tau) is called bi-admissible space if τ\tau is coarser than τ\tau_{\|\cdot\|}, τ\tau is sequentially complete on the \|\cdot\|-closed unit ball (or equivalently on \|\cdot\|-bounded sets), and (X,τ)(X,\tau)^{\prime} is norming for XX; cf. [5, Assumption 2.1, p. 3].

2.7 Lemma.

Let (X,)(X,\|\cdot\|) be a Banach space and τ\tau a Hausdorff locally convex topology on XX that is coarser than the \|\cdot\|-topology τ\tau_{\|\cdot\|}. Then the following are equivalent:

  1. (i)

    (X,,τ)(X,\|\cdot\|,\tau) is a sequentially complete Saks space.

  2. (ii)

    (X,,τ)(X,\|\cdot\|,\tau) is a bi-admissible space.

Proof.

By [33, Corollary 2.3.2, p. 55] a Saks space (X,,τ)(X,\|\cdot\|,\tau) is sequentially complete if and only if (X,τ)(X,\tau) is sequentially complete on \|\cdot\|-bounded sets, meaning that every \|\cdot\|-bounded τ\tau-Cauchy sequence converges in XX. Combined with [18, Remark 2.3 (c), p. 3] it follows that a triple (X,,τ)(X,\|\cdot\|,\tau) fulfils [23, Assumptions 1, p. 206], which provides a bi-admissible space, if and only if it is a sequentially complete Saks space. ∎

Let us recall the notion of a bi-continuous semigroup.

2.8 Definition ([23, Definition 3, p. 207]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space. A family (T(t))t0(T(t))_{t\geq 0} in (X)\mathcal{L}(X) is called τ\tau-bi-continuous semigroup if

  1. (i)

    (T(t))t0(T(t))_{t\geq 0} is a semigroup, i.e. T(t+s)=T(t)T(s)T(t+s)=T(t)T(s) and T(0)=idT(0)=\operatorname{id} for all t,s0t,s\geq 0,

  2. (ii)

    (T(t))t0(T(t))_{t\geq 0} is τ\tau-strongly continuous, i.e. the map Tx:[0,)(X,τ)T_{x}\colon[0,\infty)\to(X,\tau), Tx(t)T(t)xT_{x}(t)\coloneqq T(t)x, is continuous for all xXx\in X,

  3. (iii)

    (T(t))t0(T(t))_{t\geq 0} is exponentially bounded (of type ω\omega), i.e. there exists M1M\geq 1, ω\omega\in\mathbb{R} such that T(t)(X)Meωt\|T(t)\|_{\mathcal{L}(X)}\leq M\mathrm{e}^{\omega t} for all t0t\geq 0,

  4. (iv)

    (T(t))t0(T(t))_{t\geq 0} is locally bi-equicontinuous, i.e. for every sequence (xn)n(x_{n})_{n\in\mathbb{N}} in XX, xXx\in X with supnxn<\sup\limits_{n\in\mathbb{N}}\|x_{n}\|<\infty and τ-limnxn=x\tau\text{-}\lim\limits_{n\to\infty}x_{n}=x it holds that

    τ-limnT(t)(xnx)=0\tau\text{-}\lim_{n\to\infty}T(t)(x_{n}-x)=0

    locally uniformly for all t[0,)t\in[0,\infty).

2.9 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space.

  1. (a)

    A sequence in XX is γ\gamma-convergent if and only if it is \|\cdot\|-bounded and τ\tau-convergent by [7, I.1.10 Proposition, p. 9].

  2. (b)

    Let (X,,τ)(X,\|\cdot\|,\tau) be sequentially complete. A semigroup of linear operators (T(t))t0(T(t))_{t\geq 0} from XX to XX is γ\gamma-strongly continuous and locally sequentially γ\gamma-equicontinuous (i.e. for all γ\gamma-null sequences (xn)n(x_{n})_{n\in\mathbb{N}} in XX, t0>0t_{0}>0 and pΓγp\in\Gamma_{\gamma} we have limnsupt[0,t0]p(T(t)xn)=0\lim_{n\to\infty}\sup_{t\in[0,t_{0}]}p(T(t)x_{n})=0) if and only if it is a τ\tau-bi-continuous semigroup on XX. This follows directly fom part (a) and [18, Remark 2.6 (b), p. 5], and remains true if γ\gamma is replaced by any other Hausdorff locally convex topology on XX that has the same convergent sequences as γ\gamma (cf. [11, Proposition A.1.3, p. 73] for γ\gamma replaced by γs\gamma_{s}).

We already observed in Remark 2.9 (b) that τ\tau-bi-continuous semigroups are locally sequentially γ\gamma-equicontinuous. Under some mild conditions on the Saks space (X,,τ)(X,\|\cdot\|,\tau) they are even quasi-γ\gamma-equicontinuous. Let us recall what that means.

2.10 Definition.

Let (X,υ)(X,\upsilon) be a Hausdorff locally convex space and Γυ\Gamma_{\upsilon} a directed system of continuous seminorms that generates υ\upsilon. A family (T(t))tI(T(t))_{t\in I} of linear maps from XX to XX is called υ\upsilon-equicontinuous if

pΓυp~Γυ,C0tI,xX:p(T(t)x)Cp~(x).\forall\;p\in\Gamma_{\upsilon}\;\exists\;\widetilde{p}\in\Gamma_{\upsilon},\;C\geq 0\;\forall\;t\in I,\,x\in X:\;p(T(t)x)\leq C\widetilde{p}(x).

The family (T(t))t0(T(t))_{t\geq 0} is called locally υ\upsilon-equicontinuous if (T(t))t[0,t0](T(t))_{t\in[0,t_{0}]} is υ\upsilon-equicontinuous for all t00t_{0}\geq 0. The family (T(t))t0(T(t))_{t\geq 0} is called quasi-υ\upsilon-equicontinuous if there is α\alpha\in\mathbb{R} such that (eαtT(t))t0(\mathrm{e}^{-\alpha t}T(t))_{t\geq 0} is υ\upsilon-equicontinuous. Note that one often drops the υ\upsilon if the topology is clear.

2.11 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space. Due to Remark 2.9 (b) a γ\gamma-strongly continuous, locally γ\gamma-equicontinuous semigroup of linear operators (T(t))t0(T(t))_{t\geq 0} from XX to XX is a τ\tau-bi-continuous semigroup on XX. The converse is not true in general by [13, Example 4.1, p. 320]. However, if (X,,τ)(X,\|\cdot\|,\tau) is C-sequential, then the converse also holds by [18, Theorem 3.17 (a), p. 13]. Even more is true, namely, that every τ\tau-bi-continuous semigroup on XX is quasi-γ\gamma-equicontinuous if (X,,τ)(X,\|\cdot\|,\tau) is C-sequential.

There is another related notion to equicontinuity on Saks spaces.

2.12 Definition ([18, Definitions 3.4, 3.5, p. 6, 7]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space and Γτ\Gamma_{\tau} a directed system of continuous seminorms generating the topology τ\tau. A family of linear maps (T(t))tI(T(t))_{t\in I} from XX to XX is called (,τ)(\|\cdot\|,\tau)-equitight if

ε>0,pΓτp~Γτ,C0tI,xX:p(T(t)x)Cp~(x)+εx.\forall\;\varepsilon>0,\,p\in\Gamma_{\tau}\;\exists\;\widetilde{p}\in\Gamma_{\tau},\,C\geq 0\;\forall\;t\in I,\,x\in X:\;p(T(t)x)\leq C\widetilde{p}(x)+\varepsilon\|x\|.

The family (T(t))t0(T(t))_{t\geq 0} is called locally (,τ)(\|\cdot\|,\tau)-equitight if (T(t))t[0,t0](T(t))_{t\in[0,t_{0}]} is (,τ)(\|\cdot\|,\tau)-equitight for all t00t_{0}\geq 0. The family (T(t))t0(T(t))_{t\geq 0} is called quasi-(,τ)(\|\cdot\|,\tau)-equitight if there is α\alpha\in\mathbb{R} such that (eαtT(t))t0(\mathrm{e}^{-\alpha t}T(t))_{t\geq 0} is (,τ)(\|\cdot\|,\tau)-equitight.

At first, tight operators T(X)T\in\mathcal{L}(X) as well as families of equitight operators (T(t))t[0,t0](T(t))_{t\in[0,t_{0}]} in (X)\mathcal{L}(X) for t00t_{0}\geq 0 appeared in [11, Definitions 1.2.20, 1.2.21, p. 12] under the name local. In the setting of τ\tau-bi-continuous semigroups (T(t))t0(T(t))_{t\geq 0} the notion of tightness is used in [9, Definition 1.1, p. 668], meaning that (T(t))t[0,t0](T(t))_{t\in[0,t_{0}]} is equitight (or local) for all t00t_{0}\geq 0. Local equitightness plays an important role in perturbation results for bi-continuous semigroups, see e.g. [9, Theorem 1.2, p. 669], [12, Theorems 2.4, 3.2, p. 92, 94–95], [12, Remark 4.1, p. 101], [2, Theorem 5, p. 8], [3, Theorem 3.3, p. 582], and the corrections regarding [2] in [18, Remark 3.8, p. 8–9].

Due to [18, Proposition 3.16, p. 12–13] (,τ)(\|\cdot\|,\tau)-equitightness of a family of linear maps (T(t))tI(T(t))_{t\in I} from XX to XX implies γ\gamma-equicontinuity. If (X,,τ)(X,\|\cdot\|,\tau) is a sequentially complete C-sequential Saks space and γ=γs\gamma=\gamma_{s}, then any τ\tau-bi-continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX is quasi-γ\gamma-equicontinuous and quasi-(,τ)(\|\cdot\|,\tau)-equitight by [18, Theorem 3.17, p. 13], and both properties are equivalent by [18, Proposition 3.16, p. 12–13].

We close this section by recalling the definition of the generator of a τ\tau-bi-continuous semigroup and two of its properties which we will need.

2.13 Definition ([11, Definition 1.2.6, p. 7]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space and (T(t))t0(T(t))_{t\geq 0} a τ\tau-bi-continuous semigroup on XX. The generator (A,D(A))(A,D(A)) is defined by

D(A)\displaystyle D(A)\coloneqq {xX|τ-limt0+T(t)xxtexists in Xand supt(0,1]T(t)xxt<},\displaystyle\Bigl{\{}x\in X\;|\;\tau\text{-}\lim_{t\to 0\mathchoice{\vbox{\hbox{${\scriptstyle{+}}$}}}{\vbox{\hbox{$\scriptstyle{+}$}}}{\vbox{\hbox{$\scriptscriptstyle{+}$}}}{\vbox{\hbox{$\scriptscriptstyle{+}$}}}}\frac{T(t)x-x}{t}\;\text{exists in }X\;\text{and }\sup_{t\in(0,1]}\frac{\|T(t)x-x\|}{t}<\infty\Bigr{\}},
Ax\displaystyle Ax\coloneqq τ-limt0+T(t)xxt,xD(A).\displaystyle\tau\text{-}\lim_{t\to 0\mathchoice{\vbox{\hbox{${\scriptstyle{+}}$}}}{\vbox{\hbox{$\scriptstyle{+}$}}}{\vbox{\hbox{$\scriptscriptstyle{+}$}}}{\vbox{\hbox{$\scriptscriptstyle{+}$}}}}\frac{T(t)x-x}{t},\quad x\in D(A).
2.14 Proposition ([23, Corollary 13, p. 215]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space and (T(t))t0(T(t))_{t\geq 0} a τ\tau-bi-continuous semigroup on XX with generator (A,D(A))(A,D(A)). Then the following assertions hold:

  1. (a)

    The generator (A,D(A))(A,D(A)) is bi-closed, i.e. whenever (xn)n(x_{n})_{n\in\mathbb{N}} is a sequence in D(A)D(A) such that τ-limnxn=x\tau\text{-}\lim_{n\to\infty}x_{n}=x and τ-limnAxn=y\tau\text{-}\lim_{n\to\infty}Ax_{n}=y for some x,yXx,y\in X and both sequences are \|\cdot\|-bounded, then xD(A)x\in D(A) and Ax=yAx=y.

  2. (b)

    The generator (A,D(A))(A,D(A)) is bi-densely defined, i.e. for each xXx\in X there exists a \|\cdot\|-bounded sequence (xn)n(x_{n})_{n\in\mathbb{N}} in D(A)D(A) such that τ-limnxn=x\tau\text{-}\lim_{n\to\infty}x_{n}=x.

3. Lumer–Phillips for bi-continuous semigroups

First, we recall the relevant notions from [1] concerning dissipative linear operators on Hausdorff locally convex spaces. We write in short that (A,D(A))(A,D(A)) is a linear operator on a Hausdorff locally convex space XX if A:D(A)XXA\colon D(A)\subseteq X\to X is a linear operator.

3.1 Definition ([1, Definitions 3.1, 3.5, p. 923]).

Let (X,υ)(X,\upsilon) be a Hausdorff locally convex space and (A,D(A))(A,D(A)) a linear operator on XX.

  1. (a)

    (A,D(A))(A,D(A)) is called υ\upsilon-closed if for each net (xi)iID(A)(x_{i})_{i\in I}\subseteq D(A) satisfying xixx_{i}\to x and AxiyAx_{i}\to y w.r.t. υ\upsilon for some x,yXx,y\in X, we have xD(A)x\in D(A) and Ax=yAx=y.

  2. (b)

    A linear operator (B,D(B))(B,D(B)) on XX is called an extension of (A,D(A))(A,D(A)) if D(A)D(B)D(A)\subseteq D(B) and BD(A)=AB_{\mid D(A)}=A. The operator (A,D(A))(A,D(A)) is called υ\upsilon-closable if it admits a υ\upsilon-closed extension. The smallest υ\upsilon-closed extension of an υ\upsilon-closable operator (A,D(A))(A,D(A)) is called the υ\upsilon-closure of (A,D(A))(A,D(A)) and denoted by (A¯,D(A¯))(\overline{A},D(\overline{A})).

  3. (c)

    (A,D(A))(A,D(A)) is called (sequentially) υ\upsilon-densely defined if D(A)D(A) is (sequentially) υ\upsilon-dense in XX.

  4. (d)

    Let (A,D(A))(A,D(A)) be υ\upsilon-densely defined. The υ\upsilon-dual operator (A,D(A))(A^{\prime},D(A^{\prime})) of (A,D(A))(A,D(A)) on (X,υ)(X,\upsilon)^{\prime} is defined by setting

    D(A){x(X,υ)|y(X,υ)xD(A):Ax,x=x,y}D(A^{\prime})\coloneqq\{x^{\prime}\in(X,\upsilon)^{\prime}\;|\;\exists\;y^{\prime}\in(X,\upsilon)^{\prime}\;\forall\;x\in D(A):\;\langle Ax,x^{\prime}\rangle=\langle x,y^{\prime}\rangle\}

    and AxyA^{\prime}x^{\prime}\coloneqq y^{\prime} for xD(A)x^{\prime}\in D(A^{\prime}).

We have the following relation to the notions of a bi-closed resp. bi-densely defined operator.

3.2 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space and (A,D(A))(A,D(A)) a linear operator. Due to Remark 2.9 (a) (A,D(A))(A,D(A)) is bi-closed (see Proposition 2.14 (a)) if and only if it is sequentially γ\gamma-closed (which is defined analogously to γ\gamma-closedness but with nets replaced by sequences). Again by Remark 2.9 (a) (A,D(A))(A,D(A)) is bi-densely defined (see Proposition 2.14 (b)) if and only if it is sequentially γ\gamma-densely defined. Moreover, if (A,D(A))(A,D(A)) is sequentially γ\gamma-densely defined, then it is obviously γ\gamma-densely defined, too (as every sequence is a net).

3.3 Definition ([1, p. 922]).

Let (X,υ)(X,\upsilon) be a Hausdorff locally convex space and (A,D(A))(A,D(A)) a linear operator on XX. If λ\lambda\in\mathbb{C} is such that λAλidA:D(A)X\lambda-A\coloneqq\lambda\operatorname{id}-A\colon D(A)\to X is injective, then the linear operator (λA)1(\lambda-A)^{-1} exists and is defined on the domain Ran(λA){(λA)x|xD(A)}\operatorname{Ran}(\lambda-A)\coloneqq\{(\lambda-A)x\;|\;x\in D(A)\}, i.e. the range of λA\lambda-A. The resolvent set of AA is defined by

ρυ(A){λ|λA is bijective and (λA)1(X,υ)}.\rho_{\upsilon}(A)\coloneqq\{\lambda\in\mathbb{C}\;|\;\lambda-A\text{ is bijective and }(\lambda-A)^{-1}\in\mathcal{L}(X,\upsilon)\}.

If υ=τ\upsilon=\tau_{\|\cdot\|} for a Banach space (X,)(X,\|\cdot\|), we just write R(λ,A)(λA)1R(\lambda,A)\coloneqq(\lambda-A)^{-1} and ρ(A)ρυ(A)\rho(A)\coloneqq\rho_{\upsilon}(A).

3.4 Definition ([1, Definition 3.9, p. 925]).

Let (X,υ)(X,\upsilon) be a Hausdorff locally convex space and Γυ\Gamma_{\upsilon} a directed system of continuous seminorms that generates υ\upsilon. A linear operator (A,D(A))(A,D(A)) on XX is called Γυ\Gamma_{\upsilon}-dissipative if

λ>0,xD(A),pΓυ:p((λA)x)λp(x).\forall\;\lambda>0,\,x\in D(A),\,p\in\Gamma_{\upsilon}:\;p((\lambda-A)x)\geq\lambda p(x).

It is important to note that in contrast to equicontinuity or equitightness the notion of dissipativity depends on the selection of the directed system of continuous seminorms that generates the topology υ\upsilon by [1, Remark 3.10, p. 925–926].

3.5 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space, υ\upsilon a Hausdorff locally convex topology on XX and (A,D(A))(A,D(A)) a Γυ\Gamma_{\upsilon}-dissipative operator on XX. If Γυ\Gamma_{\upsilon} is norming, then it follows from the Γυ\Gamma_{\upsilon}-dissipativity and (1) that

λ>0,xD(A):(λA)xλx.\forall\;\lambda>0,\,x\in D(A):\;\|(\lambda-A)x\|\geq\lambda\|x\|.

Thus (A,D(A))(A,D(A)) is also a dissipative operator on the Banach space (X,)(X,\|\cdot\|) in the sense of [8, Chap. II, 3.13 Definition, p. 82] (cf. [5, Remark 3.3 (i), p. 5] for υ=τ\upsilon=\tau). We also denote such kind of dissipativity on a Banach space (X,)(X,\|\cdot\|) by \|\cdot\|-dissipativity.

In [5] another notion of dissipativity on Saks spaces was introduced.

3.6 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space. In [5, Definition 3.2, p. 5] a linear operator (A,D(A))(A,D(A)) on XX is called bi-dissipative if there exists a norming directed system of continuous seminorms Γτ\Gamma_{\tau} that generates τ\tau such that (A,D(A))(A,D(A)) is Γτ\Gamma_{\tau}-dissipative. It is then shown in the proof of [5, Theorem 3.15, p. 11] that a bi-dissipative operator (A,D(A))(A,D(A)) is also (||||||(pn,an)n)(pn,an)n𝒩({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}})_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}}-dissipative since

|(λA)x|(pn,an)n=supnpn((λA)x)ansupnλpn(x)an=λ|x|(pn,an)n{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|(\lambda-A)x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}=\sup_{n\in\mathbb{N}}p_{n}((\lambda-A)x)a_{n}\geq\sup_{n\in\mathbb{N}}\lambda p_{n}(x)a_{n}=\lambda{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}}

for all λ>0\lambda>0, xD(A)x\in D(A) and (pn,an)n𝒩(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}, where (||||||(pn,an)n)(pn,an)n𝒩({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}})_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}} is the system of seminorms that generates the submixed topology γs\gamma_{s} from Definition 2.4. We observe that this also implies that (A,D(A))(A,D(A)) is Γγs\Gamma_{\gamma_{s}}-dissipative w.r.t the norming directed system of continuous seminorms Γγs\Gamma_{\gamma_{s}} that generates the submixed topology γs\gamma_{s} from Remark 2.6.

3.7 Proposition.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space, υ\upsilon a Hausdorff locally convex topology on XX, (A,D(A))(A,D(A)) a Γυ\Gamma_{\upsilon}-dissipative operator on XX. Then the following assertions hold:

  1. (a)

    λA\lambda-A is injective for all λ>0\lambda>0. Moreover, we have

    λ>0,xRan(λA),pΓυ:p((λA)1x)1λp(x).\forall\;\lambda>0,\,x\in\operatorname{Ran}(\lambda-A),\,p\in\Gamma_{\upsilon}:\;p((\lambda-A)^{-1}x)\leq\frac{1}{\lambda}p(x). (2)
  2. (b)

    If Ran(λA)\operatorname{Ran}(\lambda-A) is (sequentially) υ\upsilon-closed for some λ>0\lambda>0, then (A,D(A))(A,D(A)) is (sequentially) υ\upsilon-closed. If υ=γ\upsilon=\gamma, then the converse even holds for all λ>0\lambda>0.

  3. (c)

    Let Γυ\Gamma_{\upsilon} be norming. Then λA\lambda-A is surjective for some λ>0\lambda>0 if and only if it is surjective for all λ>0\lambda>0. In such a case, (0,)ρ(A)(0,\infty)\subseteq\rho(A).

  4. (d)

    Let υ=γ\upsilon=\gamma. Then λA\lambda-A is surjective for some λ>0\lambda>0 if and only if it is surjective for all λ>0\lambda>0. In such a case, (0,)ργ(A)(0,\infty)\subseteq\rho_{\gamma}(A).

Proof.

Parts (a), (b) and (d) are just [1, Proposition 3.11, p. 927] in combination with the sequential completeness of (X,γ)(X,\gamma). Part (c) is a consequence of [8, Chap. II, 3.14 Proposition (ii), p. 82] and Remark 3.5. ∎

In the case υ=τ\upsilon=\tau parts (a) and (c) of Proposition 3.7 are [5, Proposition 3.4, p. 6].

3.8 Definition.

Let (X,)(X,\|\cdot\|) be a Banach space. We call a semigroup of linear operators (T(t))t0(T(t))_{t\geq 0} from XX to XX a contraction semigroup if T(t)(X)1\|T(t)\|_{\mathcal{L}(X)}\leq 1 for all t0t\geq 0.

3.9 Theorem.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space, υ\upsilon a Hausdorff locally convex topology on XX with τυτ\tau\subseteq\upsilon\subseteq\tau_{\|\cdot\|} such that γ\gamma-convergent sequences are υ\upsilon-convergent, (A,D(A))(A,D(A)) a bi-densely defined, Γυ\Gamma_{\upsilon}-dissipative operator on XX and Γυ\Gamma_{\upsilon} norming. Then the following assertions are equivalent:

  1. (a)

    (A,D(A))(A,D(A)) generates a τ\tau-bi-continuous contraction semigroup on XX.

  2. (b)

    λA\lambda-A is surjective for some λ>0\lambda>0.

Proof.

We use the Hille–Yosida theorem for bi-continuous semigroups to prove both implications (see [23, Theorem 16, p. 217] and [4, Theorem 5.6, p. 340]).

(a)\Rightarrow(b): Let (A,D(A))(A,D(A)) generate a τ\tau-bi-continuous contraction semigroup on XX. Due to [23, Theorem 16, p. 217] with ω=0\omega=0 we obtain that (0,)ρ(A)(0,\infty)\subseteq\rho(A), in particular, that λA\lambda-A is surjective for all λ>0\lambda>0.

(b)\Rightarrow(a): Let λA\lambda-A be surjective for some λ>0\lambda>0. Due to [4, Theorem 5.6, p. 340] we only need to prove that

  1. (i)

    (0,)ρ(A)(0,\infty)\subseteq\rho(A),

  2. (ii)

    R(λ,A)n(X)1λn\|R(\lambda,A)^{n}\|_{\mathcal{L}(X)}\leq\frac{1}{\lambda^{n}} for all nn\in\mathbb{N} and all λ>0\lambda>0, and

  3. (iii)

    {(λα)nR(λ,A)n|n,λα}\{(\lambda-\alpha)^{n}R(\lambda,A)^{n}\;|\;n\in\mathbb{N},\,\lambda\geq\alpha\} is bi-equicontinuous for each α>0\alpha>0, i.e. for each α>0\alpha>0 and each \|\cdot\|-bounded τ\tau-null sequence (xm)m(x_{m})_{m\in\mathbb{N}} in XX one has that τ\tau-limm(λα)nR(λ,A)nxm=0\lim_{m\to\infty}(\lambda-\alpha)^{n}R(\lambda,A)^{n}x_{m}=0 uniformly for all nn\in\mathbb{N} and all λα\lambda\geq\alpha.

Since (A,D(A))(A,D(A)) is Γυ\Gamma_{\upsilon}-dissipative, we get that λA\lambda-A is bijective for all λ>0\lambda>0 and ρ(A)(0,)\rho(A)\subseteq(0,\infty) by Proposition 3.7 (a) and (c). From Remark 3.5 and Γυ\Gamma_{\upsilon} being norming we deduce that R(λ,A)x1λx\|R(\lambda,A)x\|\leq\tfrac{1}{\lambda}\|x\| for all λ>0\lambda>0 and xRan(λA)=Xx\in\operatorname{Ran}(\lambda-A)=X, yielding R(λ,A)n(X)1λn\|R(\lambda,A)^{n}\|_{\mathcal{L}(X)}\leq\tfrac{1}{\lambda^{n}} for all nn\in\mathbb{N} and λ>0\lambda>0. Let Γτ\Gamma_{\tau} be a directed system of continuous seminorms that generates the topology τ\tau and qΓτq\in\Gamma_{\tau}. Thanks to (2) we know that p((λA)1x)1λp(x)p((\lambda-A)^{-1}x)\leq\frac{1}{\lambda}p(x) for all pΓυp\in\Gamma_{\upsilon}, λ>0\lambda>0 and xRan(λA)=Xx\in\operatorname{Ran}(\lambda-A)=X. As τυ\tau\subseteq\upsilon, there are pΓυp\in\Gamma_{\upsilon} and C0C\geq 0 such that for each α>0\alpha>0 we have

q((λα)nR(λ,A)nx)\displaystyle q((\lambda-\alpha)^{n}R(\lambda,A)^{n}x) C(λα)np((λA)nx)C(λα)nλnp(x)\displaystyle\leq C(\lambda-\alpha)^{n}p((\lambda-A)^{-n}x)\leq C\frac{(\lambda-\alpha)^{n}}{\lambda^{n}}p(x)
C(1αλ)np(x)Cp(x)\displaystyle\leq C\left(1-\frac{\alpha}{\lambda}\right)^{n}p(x)\leq Cp(x)

for all xXx\in X, nn\in\mathbb{N} and λα\lambda\geq\alpha. Since \|\cdot\|-bounded τ\tau-null sequences are exactly the γ\gamma-null-sequences by Remark 2.9 (a) and γ\gamma-convergent sequences are assumed to be υ\upsilon-convergent, this inequality implies that {(λα)nR(λ,A)n|n,λα}\{(\lambda-\alpha)^{n}R(\lambda,A)^{n}\;|\;n\in\mathbb{N},\,\lambda\geq\alpha\} is bi-equicontinuous for all α>0\alpha>0. This finishes the proof. ∎

In the case υ=τ\upsilon=\tau we know that γ\gamma-convergent sequences are τ\tau-convergent and thus Theorem 3.9 is [5, Theorem 3.6, p. 6] (without the superfluous assumption that AA should be norm-closed) in this case. Another possible choice is υ=γs\upsilon=\gamma_{s} since the submixed topology γs\gamma_{s} has the same convergent sequences as γ\gamma by Remark 2.5 (a). However, we are mostly interested in the choice υ=γ\upsilon=\gamma. Our second generation result involves complete Saks spaces.

3.10 Theorem.

Let (X,,τ)(X,\|\cdot\|,\tau) be a complete Saks space and (A,D(A))(A,D(A)) a γ\gamma-densely defined, Γγ\Gamma_{\gamma}-dissipative operator. Assume that Ran(λA)\operatorname{Ran}(\lambda-A) is γ\gamma-dense in XX for some λ>0\lambda>0. Then the following assertions hold:

  1. (a)

    The γ\gamma-closure (A¯,D(A¯))(\overline{A},D(\overline{A})) generates a γ\gamma-strongly continuous, γ\gamma-equicontinuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX.

  2. (b)

    If Γγ\Gamma_{\gamma} is norming, then (T(t))t0(T(t))_{t\geq 0} is a contraction semigroup.

  3. (c)

    If Γγ\Gamma_{\gamma} is norming and γ=γs\gamma=\gamma_{s}, then (T(t))t0(T(t))_{t\geq 0} is (,τ)(\|\cdot\|,\tau)-equitight.

Proof.

(a) Due to [1, Theorem 3.14, p. 929] (A¯,D(A¯))(\overline{A},D(\overline{A})) generates a γ\gamma-equicontinuous, γ\gamma-strongly continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX.

(b) By [1, Proposition 3.13, p. 929] the operator (A¯,D(A¯))(\overline{A},D(\overline{A})) is also Γγ\Gamma_{\gamma}-dissipative and λA¯\lambda-\overline{A} is surjective for all λ>0\lambda>0. As a consequence of part (a) and Remark 2.9 (T(t))t0(T(t))_{t\geq 0} is also a τ\tau-bi-continuous semigroup on XX. By [19, p. 5] the generator (A¯,D(A¯))(\overline{A},D(\overline{A})) of (T(t))t0(T(t))_{t\geq 0} as a γ\gamma-strongly continuous, γ\gamma-equicontinuous semigroup (see [1, p. 922]) and the generator of (T(t))t0(T(t))_{t\geq 0} as a τ\tau-bi-continuous semigroup (see Definition 2.13) coincide. Thus (A¯,D(A¯))(\overline{A},D(\overline{A})) is bi-densely defined by Proposition 2.14 (b). Hence we get that (T(t))t0(T(t))_{t\geq 0} is a contraction semigroup by Theorem 3.9 with υ=γ\upsilon=\gamma and the norming property of Γγ\Gamma_{\gamma}.

(c) It follows from part (b) that T(t)(X)1\|T(t)\|_{\mathcal{L}(X)}\leq 1 for all t0t\geq 0. Further, (T(t))t0(T(t))_{t\geq 0} is γ\gamma-equicontinuous by part (a). In combination with γ=γs\gamma=\gamma_{s} we derive that (T(t))t0(T(t))_{t\geq 0} is (,τ)(\|\cdot\|,\tau)-equitight by [18, Proposition 3.16, p. 12–13]. ∎

Let us compare Theorem 3.10 with one of the main theorems of [5], namely, [5, Theorem 3.15, p. 11]. We note that the topology that is called mixed topology (and denoted by γ\gamma there) in [5, p. 10] is actually the submixed topology γs\gamma_{s}. With this observation at hand let us phrase [5, Theorem 3.15, p. 11] in our terminology.

3.11 Theorem ([5, Theorem 3.15, p. 11]).

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space such that (X,γs)(X,\gamma_{s}) is complete, and (A,D(A))(A,D(A)) a bi-densely defined, bi-dissipative operator. Assume that Ran(λA)\operatorname{Ran}(\lambda-A) is bi-dense, i.e. sequentially γ\gamma-dense, in XX for some λ>0\lambda>0. Then the γs\gamma_{s}-closure (A¯γs,D(A¯γs))(\overline{A}^{\gamma_{s}},D(\overline{A}^{\gamma_{s}})) generates a τ\tau-bi-continuous contraction semigroup on XX.

3.12 Remark.
  1. (a)

    First, it is actually shown in the proof of Theorem 3.11 that (A¯γs,D(A¯γs))(\overline{A}^{\gamma_{s}},D(\overline{A}^{\gamma_{s}})) generates a γs\gamma_{s}-strongly continuous, γs\gamma_{s}-equicontinuous semigroup on XX, in particular, a τ\tau-bi-continuous semigroup by Remark 2.9 and Remark 2.5 (a). However, the proof that the generated semigroup is contractive is missing. In this proof it is used that a bi-dissipative operator is (||||||(pn,an)n)(pn,an)n𝒩({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}})_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}}-dissipative as well (see Remark 3.6). In order to prove that the generated semigroup is also contractive, the only available tool in [5] is [5, Theorem 3.6, p. 6]. However, to apply the latter theorem one has to show that (A¯γs,D(A¯γs))(\overline{A}^{\gamma_{s}},D(\overline{A}^{\gamma_{s}})) is also bi-dissipative. Due to [1, Proposition 3.13, p. 929] we only know that (A¯γs,D(A¯γs))(\overline{A}^{\gamma_{s}},D(\overline{A}^{\gamma_{s}})) is (||||||(pn,an)n)(pn,an)n𝒩({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{(p_{n},a_{n})_{n\in\mathbb{N}}})_{(p_{n},a_{n})_{n\in\mathbb{N}}\in\mathcal{N}}-dissipative (and λA¯γs\lambda-\overline{A}^{\gamma_{s}} is surjective for all λ>0\lambda>0). To circumvent this obstacle, we relaxed [5, Theorem 3.6, p. 6] to Theorem 3.9 where one has several possible choices for the topology υ\upsilon, not only υ=τ\upsilon=\tau as in [5, Theorem 3.6, p. 6]. Using Remark 3.6 and [1, Proposition 3.13, p. 929], we see that (A¯γs,D(A¯γs))(\overline{A}^{\gamma_{s}},D(\overline{A}^{\gamma_{s}})) is Γγs\Gamma_{\gamma_{s}}-dissipative w.r.t the norming directed system of continuous seminorms Γγs\Gamma_{\gamma_{s}} from Remark 2.6. Now, it is possible to apply Theorem 3.9 with υ=γs\upsilon=\gamma_{s} to conclude that the generated semigroup is contractive. This closes the gap in the proof of Theorem 3.11.

  2. (b)

    There is no nice characterisation (known to us) of the completeness of (X,γs)(X,\gamma_{s}), that is assumed in Theorem 3.11. However, there is a nice characterisation of the completeness of the Saks space (X,,τ)(X,\|\cdot\|,\tau). By definition the Saks space is complete if and only if (X,γ)(X,\gamma) is complete. The space (X,γ)(X,\gamma) is complete if and only if B={xX|x1}B_{\|\cdot\|}=\{x\in X\;|\;\|x\|\leq 1\} is τ\tau-complete by [7, I.1.14 Proposition, p. 11]. But, since γs\gamma_{s} is in general a weaker topology than γ\gamma by Remark 2.5 (a), the completeness of (X,γ)(X,\gamma) does in general not imply the completeness of (X,γs)(X,\gamma_{s}).

  3. (c)

    Let us suppose that γ=γs\gamma=\gamma_{s}. Then Theorem 3.11 is covered by Theorem 3.10 (a) and (b). Further, we point out that in comparison to Theorem 3.11 we weakened the assumptions from (A,D(A))(A,D(A)) being a bi-densely defined, bi-dissipative operator and Ran(λA)\operatorname{Ran}(\lambda-A) being bi-dense for some λ>0\lambda>0 to (A,D(A))(A,D(A)) being a γ\gamma-densely defined, Γγ\Gamma_{\gamma}-dissipative operator and Ran(λA)\operatorname{Ran}(\lambda-A) being γ\gamma-dense for some λ>0\lambda>0 (see Remark 3.2) in Theorem 3.10.

Let us take a closer look at the completeness assumption on the Saks space (X,,τ)(X,\|\cdot\|,\tau) in Theorem 3.10, which is actually fulfilled for many important examples, and its characterisation in Remark 3.12 (b). Especially, (X,γ)(X,\gamma) is complete, thus (X,,τ)(X,\|\cdot\|,\tau) as well, if BB_{\|\cdot\|} is τ\tau-compact, which is condition (ii) of Remark 2.5 (b) and also a sufficient condition for γ=γs\gamma=\gamma_{s}. We recall the following observations from [18, Examples 2.4, 3.11, p. 4–5, 10], [18, Remark 3.20 (a), p. 15], [18, Example 4.12, p. 24–25] and [18, Corollary 3.23, p. 17], and add a proof of the completeness of the Saks spaces considered in Remark 3.13 (c), (d) and (f) below.

3.13 Remark.
  1. (a)

    Let Ω\Omega be a Hausdorff kk_{\mathbb{R}}-space and recall that a completely regular space Ω\Omega is called kk_{\mathbb{R}}-space if any map f:Ωf\colon\Omega\to\mathbb{R} whose restriction to each compact KΩK\subset\Omega is continuous, is already continuous on Ω\Omega (see [27, p. 487]). Further, let Cb(Ω)\mathrm{C}_{\operatorname{b}}(\Omega) be the space of bounded continuous functions on Ω\Omega, and \|\cdot\|_{\infty} the sup-norm as well as τco\tau_{\operatorname{co}} the compact-open topology, i.e. the topology of uniform convergence on compact subsets of Ω\Omega. Then (Cb(Ω),,τco)(\mathrm{C}_{\operatorname{b}}(\Omega),\|\cdot\|_{\infty},\tau_{\operatorname{co}}) is a complete Saks space and γ(,τco)=γs(,τco)\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}})=\gamma_{s}(\|\cdot\|_{\infty},\tau_{\operatorname{co}}).

    Let 𝒱\mathcal{V} denote the set of all non-negative bounded functions ν\nu on Ω\Omega that vanish at infinity, i.e. for every ε>0\varepsilon>0 the set {xΩ|ν(x)ε}\{x\in\Omega\;|\;\nu(x)\geq\varepsilon\} is compact. Let β0\beta_{0} be the Hausdorff locally convex topology on Cb(Ω)\mathrm{C}_{\operatorname{b}}(\Omega) that is induced by the seminorms

    |f|νsupxΩ|f(x)|ν(x),fCb(Ω),|f|_{\nu}\coloneqq\sup_{x\in\Omega}|f(x)|\nu(x),\quad f\in\mathrm{C}_{\operatorname{b}}(\Omega),

    for ν𝒱\nu\in\mathcal{V}. Then we have γ(,τco)=β0\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}})=\beta_{0}. If Ω\Omega is locally compact, then 𝒱\mathcal{V} may be replaced by the functions in C0(Ω)\mathrm{C}_{0}(\Omega) that are non-negative where C0(Ω)\mathrm{C}_{0}(\Omega) is the space of real-valued continuous functions on Ω\Omega that vanish at infinity.

    If Ω\Omega is a hemicompact Hausdorff kk_{\mathbb{R}}-space or a Polish space, then we even have

    γ(,τco)=β0=μ(Cb(Ω),Mt(Ω))\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}})=\beta_{0}=\mu(\mathrm{C}_{\operatorname{b}}(\Omega),\mathrm{M}_{\operatorname{t}}(\Omega))

    where Mt(Ω)=(Cb(Ω),β0)\mathrm{M}_{\operatorname{t}}(\Omega)=(\mathrm{C}_{\operatorname{b}}(\Omega),\beta_{0})^{\prime} is the space of bounded Radon measures and μ(Cb(Ω),Mt(Ω))\mu(\mathrm{C}_{\operatorname{b}}(\Omega),\mathrm{M}_{\operatorname{t}}(\Omega)) the Mackey-topology of the dual pair (Cb(Ω),Mt(Ω))(\mathrm{C}_{\operatorname{b}}(\Omega),\mathrm{M}_{\operatorname{t}}(\Omega)).

  2. (b)

    Let (X,)(X,\|\cdot\|) be a Banach space and σσ(X,X)\sigma^{\ast}\coloneqq\sigma(X^{\prime},X) the weak-topology. Then condition (ii) of Remark 2.5 (b) is fulfilled, (X,X,σ)(X^{\prime},\|\cdot\|_{X^{\prime}},\sigma^{\ast}) is a complete Saks space and γ(X,σ)=γs(,σ)=τc(X,X)\gamma(\|\cdot\|_{X^{\prime}},\sigma^{\ast})=\gamma_{s}(\|\cdot\|_{\infty},\sigma^{\ast})=\tau_{\operatorname{c}}(X^{\prime},X) where τc(X,X)\tau_{\operatorname{c}}(X^{\prime},X) is the topology of uniform convergence on compact subsets of XX.

  3. (c)

    Let (X,)(X,\|\cdot\|) be a Banach space and μμ(X,X)\mu^{\ast}\coloneqq\mu(X^{\prime},X) the dual Mackey-topology. Then (X,X,μ)(X^{\prime},\|\cdot\|_{X^{\prime}},\mu^{\ast}) is a complete Saks space, where the completeness follows from [17, p. 74], and γ(X,μ)=μ\gamma(\|\cdot\|_{X^{\prime}},\mu^{\ast})=\mu^{\ast}. If XX is a Schur space, i.e. every σ(X,X)\sigma(X,X^{\prime})-convergent sequence is \|\cdot\|-convergent (see [10, p. 253]), then condition (ii) of Remark 2.5 (b) is fulfilled and γ(X,μ)=γs(X,μ)\gamma(\|\cdot\|_{X^{\prime}},\mu^{\ast})=\gamma_{s}(\|\cdot\|_{X^{\prime}},\mu^{\ast}).

  4. (d)

    Let (X,X)(X,\|\cdot\|_{X}) and (Y,Y)(Y,\|\cdot\|_{Y}) be Banach spaces, and τsot\tau_{\operatorname{sot}} the strong operator topology on (X;Y)\mathcal{L}(X;Y). Then ((X;Y),(X;Y),τsot)(\mathcal{L}(X;Y),\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}}) is a Saks space. Let (Ti)iI(T_{i})_{i\in I} be a τsot\tau_{\operatorname{sot}}-Cauchy net in B(X;Y)={T(X;Y)|T(X;Y)1}B_{\|\cdot\|_{\mathcal{L}(X;Y)}}=\{T\in\mathcal{L}(X;Y)\;|\;\|T\|_{\mathcal{L}(X;Y)}\leq 1\}. Then for each xXx\in X the net (Tix)iI(T_{i}x)_{i\in I} is Y\|\cdot\|_{Y}-convergent to some TxYTx\in Y with TxYxX\|Tx\|_{Y}\leq\|x\|_{X} in the Banach space (Y,Y)(Y,\|\cdot\|_{Y}). Thus the map T:xTxT\colon x\mapsto Tx belongs to (X;Y)\mathcal{L}(X;Y) with T(X;Y)1\|T\|_{\mathcal{L}(X;Y)}\leq 1 and (Ti)iI(T_{i})_{i\in I} is τsot\tau_{\operatorname{sot}}-convergent to TT. Hence B(X;Y)B_{\|\cdot\|_{\mathcal{L}(X;Y)}} is τsot\tau_{\operatorname{sot}}-complete and so ((X;Y),(X;Y),τsot)(\mathcal{L}(X;Y),\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}}) is complete. If YY is in addition finite-dimensional, then condition (ii) of Remark 2.5 (b) is fulfilled and γ((X;Y),τsot)=γs((X;Y),τsot)\gamma(\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}})=\gamma_{s}(\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}}).

  5. (e)

    Let HH be a separable Hilbert space and 𝒩(H)\mathcal{N}(H) the space of trace class operators in (H)\mathcal{L}(H) and note that (H)=𝒩(H)\mathcal{L}(H)=\mathcal{N}(H)^{\prime}. Let τsot\tau_{\operatorname{sot}^{\ast}} be the symmetric strong operator topology, i.e. the Hausdorff locally convex topology on (H)\mathcal{L}(H) generated by the directed system of seminorms

    pN(R)max(supxNRxH,supxNRxH),R(H),p_{N}(R)\coloneqq\max\bigl{(}\sup_{x\in N}\|Rx\|_{H},\sup_{x\in N}\|R^{\ast}x\|_{H}\bigr{)},\quad R\in\mathcal{L}(H),

    for finite NHN\subset H where RR^{\ast} is the adjoint of RR. We denote by βsot\beta_{\operatorname{sot}^{\ast}} the mixed topology γ((H),τsot)\gamma(\|\cdot\|_{\mathcal{L}(H)},\tau_{\operatorname{sot}^{\ast}}). Then the triple ((H),(H),τsot)(\mathcal{L}(H),\|\cdot\|_{\mathcal{L}(H)},\tau_{\operatorname{sot}^{\ast}}) is a complete Saks space and βsot=μ((H),𝒩(H))\beta_{\operatorname{sot}^{\ast}}=\mu(\mathcal{L}(H),\mathcal{N}(H)).

  6. (f)

    Let Ω\Omega be a completely regular Hausdorff space, Mt(Ω)\mathrm{M}_{\operatorname{t}}(\Omega) the space of bounded Radon measures on Ω\Omega, and Mt(Ω)\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)} the total variation norm on Mt(Ω)\mathrm{M}_{\operatorname{t}}(\Omega). Then (Mt(Ω),Mt(Ω),σ(Mt(Ω),Cb(Ω)))(\mathrm{M}_{\operatorname{t}}(\Omega),\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)},\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega))) is a complete Saks space where the completeness follows from BMt(Ω)B_{\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)}} being σ(Mt(Ω),Cb(Ω))\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega))-compact by [18, Corollary 3.23 (a), p. 17], i.e. condition (ii) of Remark 2.5 (b) is fulfilled. Furthermore, we have

    β0\displaystyle\beta_{0}^{\prime}\coloneqq γ(Mt(Ω),σ(Mt(Ω),Cb(Ω)))=γs(Mt(Ω),σ(Mt(Ω),Cb(Ω)))\displaystyle\gamma(\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)},\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega)))=\gamma_{s}(\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)},\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega)))
    =\displaystyle= τc(Mt(Ω),(Cb(Ω),)).\displaystyle\tau_{\operatorname{c}}(\mathrm{M}_{\operatorname{t}}(\Omega),(\mathrm{C}_{\operatorname{b}}(\Omega),\|\cdot\|_{\infty})).

Let us consider a toy example for an application of Theorem 3.10, namely, the multiplication operator on Cb(Ω)\mathrm{C}_{\operatorname{b}}(\Omega), which we will revisit for other generation results.

3.14 Example.

Let Ω\Omega be a Hausdorff kk_{\mathbb{R}}-space and q:Ωq\colon\Omega\to\mathbb{C} be continuous with CsupxΩReq(x)<C\coloneqq\sup_{x\in\Omega}\operatorname{Re}q(x)<\infty. We define the multiplication operator (Mq,D(Mq))(M_{q},D(M_{q})) by setting

D(Mq){fCb(Ω)|qfCb(Ω)}D(M_{q})\coloneqq\{f\in\mathrm{C}_{\operatorname{b}}(\Omega)\;|\;qf\in\mathrm{C}_{\operatorname{b}}(\Omega)\}

and MqqfM_{q}\coloneqq qf for fD(Mq)f\in D(M_{q}). By solving the equation (λq)f=g(\lambda-q)f=g we can compute the resolvent R(λ,Mq)R(\lambda,M_{q}) of MqM_{q} explicitly by

R(λ,Mq)f=1λqf,fCb(Ω),R(\lambda,M_{q})f=\frac{1}{\lambda-q}f,\quad f\in\mathrm{C}_{\operatorname{b}}(\Omega),

for all λ(q(Ω)¯)=ρ(Mq)\lambda\in(\mathbb{C}\setminus\overline{q(\Omega)})=\rho(M_{q}), which shows that λMq\lambda-M_{q} is surjective, i.e. Ran(λMq)=Cb(Ω)\operatorname{Ran}(\lambda-M_{q})=\mathrm{C}_{\operatorname{b}}(\Omega), for all λq(Ω)¯\lambda\in\mathbb{C}\setminus\overline{q(\Omega)}. Suppose that C0C\leq 0. Then (0,)ρ(Mq)(0,\infty)\in\rho(M_{q}) and Ran(λMq)=Cb(Ω)\operatorname{Ran}(\lambda-M_{q})=\mathrm{C}_{\operatorname{b}}(\Omega) for all λ>0\lambda>0. Furthermore, we have for all λ>0\lambda>0, fCb(Ω)f\in\mathrm{C}_{\operatorname{b}}(\Omega) and ν𝒱\nu\in\mathcal{V} from Remark 3.13 (a) that

|R(λ,Mq)f|ν\displaystyle|R(\lambda,M_{q})f|_{\nu} =supxΩ1|λq(x)||f(x)|ν(x)C0supxΩ1λReq(x)|f(x)|ν(x)\displaystyle=\sup_{x\in\Omega}\frac{1}{|\lambda-q(x)|}|f(x)|\nu(x)\underset{C\leq 0}{\leq}\sup_{x\in\Omega}\frac{1}{\lambda-\operatorname{Re}q(x)}|f(x)|\nu(x)
1λsupxΩ|f(x)|ν(x)=1λ|f|ν.\displaystyle\leq\frac{1}{\lambda}\sup_{x\in\Omega}|f(x)|\nu(x)=\frac{1}{\lambda}|f|_{\nu}.

Therefore (Mq,D(Mq))(M_{q},D(M_{q})) is Γβ0\Gamma_{\beta_{0}}-dissipative for the directed system of seminorms Γβ0(||ν)ν𝒱\Gamma_{\beta_{0}}\coloneqq(|\cdot|_{\nu})_{\nu\in\mathcal{V}} that generates the mixed topology β0=γ(,τco)\beta_{0}=\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}}). Moreover, due to Proposition 3.7 (b) and Ran(λMq)=Cb(Ω)\operatorname{Ran}(\lambda-M_{q})=\mathrm{C}_{\operatorname{b}}(\Omega) for all λ>0\lambda>0 the operator (Mq,D(Mq))(M_{q},D(M_{q})) is β0\beta_{0}-closed and thus generates a β0\beta_{0}-strongly continuous, β0\beta_{0}-equicontinuous semigroup (T(t))t0(T(t))_{t\geq 0} on Cb(Ω)\mathrm{C}_{\operatorname{b}}(\Omega) by Theorem 3.10 (a) and Remark 3.13 (a). Choosing 𝒱1{ν𝒱|xΩ:ν(x)1}\mathcal{V}_{1}\coloneqq\{\nu\in\mathcal{V}\;|\;\forall\;x\in\Omega:\;\nu(x)\leq 1\} instead of 𝒱\mathcal{V}, we get a norming directed system of continuous seminorms that generates β0\beta_{0} for which (Mq,D(Mq))(M_{q},D(M_{q})) is dissipative, too. Hence (T(t))t0(T(t))_{t\geq 0} is also a (,τco)(\|\cdot\|_{\infty},\tau_{\operatorname{co}})-equitight contraction semigroup by Theorem 3.10 (b) and (c) since β0=γ(,τco)=γs(,τco)\beta_{0}=\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}})=\gamma_{s}(\|\cdot\|_{\infty},\tau_{\operatorname{co}}) by Remark 3.13 (a).

Our next generation result involves the γ\gamma-dual operator. Let us recall some observations from [19, Remark 4.5, p. 9]. Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space. Then Xγ(X,γ)X_{\gamma}^{\prime}\coloneqq(X,\gamma)^{\prime} is a closed linear subspace of XX^{\prime}, in particular a Banach space, by [7, I.1.18 Proposition, p. 15], and we denote by Xγ\|\cdot\|_{X_{\gamma}^{\prime}} the restriction of X\|\cdot\|_{X^{\prime}} to XγX_{\gamma}^{\prime}. We note that (X,γ)(X,\gamma) is a Mazur space, i.e. XγX_{\gamma}^{\prime} coincides with the space of linear γ\gamma-sequentially continuous functionals on XX (see [32, p. 50]), if and only if

Xγ={xX|xτ-sequentially continuous on -bounded sets}XX_{\gamma}^{\prime}=\{x^{\prime}\in X^{\prime}\;|\;x^{\prime}\;\tau\text{-sequentially continuous on }\|\cdot\|\text{-bounded sets}\}\eqqcolon X^{\circ}

by [7, I.1.10 Proposition, p. 9]. The space XX^{\circ} was introduced in [13, Proposition 2.1, p. 314] in the context of dual semigroups of bi-continuous semigroups.

3.15 Corollary.

Let (X,,τ)(X,\|\cdot\|,\tau) be a complete Saks space. Let both (A,D(A))(A,D(A)) and its γ\gamma-dual operator (A,D(A))(A^{\prime},D(A^{\prime})) be Γγ\Gamma_{\gamma}-dissipative and Xγ\|\cdot\|_{X_{\gamma}^{\prime}}-dissipative operators, respectively. Then the following assertions hold:

  1. (a)

    The γ\gamma-closure (A¯,D(A¯))(\overline{A},D(\overline{A})) generates a γ\gamma-equicontinuous, γ\gamma-strongly continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX.

  2. (b)

    If Γγ\Gamma_{\gamma} is norming, then (T(t))t0(T(t))_{t\geq 0} is a contraction semigroup.

  3. (c)

    If Γγ\Gamma_{\gamma} is norming and γ=γs\gamma=\gamma_{s}, then (T(t))t0(T(t))_{t\geq 0} is (,τ)(\|\cdot\|,\tau)-equitight.

Proof.

By [7, I.1.18 Proposition (i), p. 15] we have (Xγ,τb)=(Xγ,Xγ)(X_{\gamma}^{\prime},\tau_{b})=(X_{\gamma}^{\prime},\|\cdot\|_{X_{\gamma}^{\prime}}) where τb\tau_{b} denotes the topology of uniform convergence on γ\gamma-bounded sets. Due to [1, Corollary 3.17, p. 931] we get that (A¯,D(A¯))(\overline{A},D(\overline{A})) generates a γ\gamma-strongly continuous, γ\gamma-equicontinuous semigroup on XX. Parts (b) and (c) follow as in Theorem 3.10. ∎

3.16 Example.

Let Ω\Omega\coloneqq\mathbb{N} be equipped with the metric induced by the absolute value. Then Ω\Omega is a Polish space, in particular, a Hausdorff kk_{\mathbb{R}}-space. Moreover, Cb()=\mathrm{C}_{\operatorname{b}}(\mathbb{N})=\ell^{\infty} and Mt()=1\mathrm{M}_{\operatorname{t}}(\mathbb{N})=\ell^{1} (see e.g. [6, p. 477]). It follows from Remark 3.13 (a) that β0=μ(,1)\beta_{0}=\mu(\ell^{\infty},\ell^{1}) and so

(,β0)=(,μ(,1))=1.(\ell^{\infty},\beta_{0})^{\prime}=(\ell^{\infty},\mu(\ell^{\infty},\ell^{1}))^{\prime}=\ell^{1}.

Let q:q\colon\mathbb{N}\to\mathbb{C} be a function with CsupnReq(n)0C\coloneqq\sup_{n\in\mathbb{N}}\operatorname{Re}q(n)\leq 0. Again, we consider the multiplication operator MqM_{q} from Example 3.14, i.e.

D(Mq)={f|qf}D(M_{q})=\{f\in\ell^{\infty}\;|\;qf\in\ell^{\infty}\}

and Mq=qfM_{q}=qf for fD(Mq)f\in D(M_{q}). We already know that (Mq,D(Mq))(M_{q},D(M_{q})) is Γβ0\Gamma_{\beta_{0}}-dissipative with Γβ0\Gamma_{\beta_{0}} from Example 3.14. Furthermore, we have for the β0\beta_{0}-dual operator (Mq,D(Mq))(M_{q}^{\prime},D(M_{q}^{\prime})) that

D(Mq)={f1|qf1}D(M_{q}^{\prime})=\{f\in\ell^{1}\;|\;qf\in\ell^{1}\}

and Mq=qfM_{q}^{\prime}=qf for fD(Mq)f\in D(M_{q}^{\prime}). For all λ>0\lambda>0 and fD(Mq)f\in D(M_{q}^{\prime}) we get

(λMq)f1=n=1|(λq(n))fn|C0n=1(λReq(n))|fn|λn=1|fn|=λf1,\|(\lambda-M_{q}^{\prime})f\|_{\ell^{1}}=\sum_{n=1}^{\infty}|(\lambda-q(n))f_{n}|\underset{C\leq 0}{\geq}\sum_{n=1}^{\infty}(\lambda-\operatorname{Re}q(n))|f_{n}|\geq\lambda\sum_{n=1}^{\infty}|f_{n}|=\lambda\|f\|_{\ell^{1}},

meaning that (Mq,D(Mq))(M_{q}^{\prime},D(M_{q}^{\prime})) is 1\|\cdot\|_{\ell^{1}}-dissipative. Thus we may apply Corollary 3.15 (a) to deduce that (Mq,D(Mq))(M_{q},D(M_{q})) generates a μ(,1)\mu(\ell^{\infty},\ell^{1})-strongly continuous, μ(,1)\mu(\ell^{\infty},\ell^{1})-equicontinuous semigroup on \ell^{\infty}.

Instead of Remark 3.13 (a) we may also use Remark 3.13 (c) in Example 3.16 since 1\ell^{1} is a Schur space by [10, Theorem 5.36, p. 252].

Next, we would like to transfer [1, Theorem 3.18, p. 931] to the setting of Saks spaces (X,,τ)(X,\|\cdot\|,\tau). However, looking at the assumptions of [1, Theorem 3.18, p. 931], we see that this requires (X,γ)(X,\gamma) to be reflexive. Since reflexive spaces are barrelled, this requirement implies that τ=γ=τ\tau=\gamma=\tau_{\|\cdot\|} by [7, I.1.15 Proposition, p. 12] and so we are in an uninteresting situation from the perspective of τ\tau-bi-continuous semigroups. But if we could relax the assumption to (X,γ)(X,\gamma) being semi-reflexive, then there are non-trivial (i.e. not Banach) Saks spaces. This is actually possible by the following observation.

3.17 Remark.

[1, Theorem 3.18, p. 931] is stated for reflexive Hausdorff locally convex spaces (X,υ)(X,\upsilon). However, a closer look at its proof reveals that it is actually valid for semi-reflexive (X,υ)(X,\upsilon) because the only part where reflexivity comes into play is that it implies that a υ\upsilon-bounded set BXB\subset X is relatively σ(X,(X,υ))\sigma(X,(X,\upsilon)^{\prime})-compact; see [1, p. 931, l. 9–10 from below]. But the latter assertion is equivalent to semi-reflexivity by [26, Proposition 23.18, p. 270].

3.18 Theorem.

Let (X,,τ)(X,\|\cdot\|,\tau) be a complete, semi-reflexive Saks space, (A,D(A))(A,D(A)) a γ\gamma-densely defined, Γγ\Gamma_{\gamma}-dissipative operator and Ran(λA)=X\operatorname{Ran}(\lambda-A)=X for some λ>0\lambda>0. Then the following assertions hold:

  1. (a)

    (A,D(A))(A,D(A)) generates a γ\gamma-equicontinuous, γ\gamma-strongly continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX.

  2. (b)

    If Γγ\Gamma_{\gamma} is norming, then (T(t))t0(T(t))_{t\geq 0} is a contraction semigroup.

  3. (c)

    If Γγ\Gamma_{\gamma} is norming and γ=γs\gamma=\gamma_{s}, then (T(t))t0(T(t))_{t\geq 0} is (,τ)(\|\cdot\|,\tau)-equitight.

Proof.

Part (a) follows from [1, Theorem 3.18, p. 931] (noting that 1A1-A is surjective by Proposition 3.7 (d)) and Remark 3.17. Parts (b) and (c) follow as in Theorem 3.10. ∎

Let (X,,τ)(X,\|\cdot\|,\tau) be a Saks space. By definition the Saks space is semi-reflexive if and only if (X,γ)(X,\gamma) is semi-reflexive. The space (X,γ)(X,\gamma) is semi-reflexive if and only if B={xX|x1}B_{\|\cdot\|}=\{x\in X\;|\;\|x\|\leq 1\} is σ(X,(X,τ))\sigma(X,(X,\tau)^{\prime})-compact by [7, I.1.21 Corollary, p. 16]. Due to [7, I.1.20 Proposition, p. 16], BB_{\|\cdot\|} is σ(X,(X,τ))\sigma(X,(X,\tau)^{\prime})-compact if and only if it is σ(X,(X,γ))\sigma(X,(X,\gamma)^{\prime})-compact. Further, (X,γ)(X,\gamma) is a semi-Montel space, thus semi-reflexive, if and only if BB_{\|\cdot\|} is τ\tau-compact by [7, I.1.13 Proposition, p. 11] which is condition (ii) in Remark 2.5 (b) again and also a sufficient condition for γ=γs\gamma=\gamma_{s}. Therefore we have by Remark 3.13 the following observations where we only have to add an additional argument in parts (a), (c) and (e) of Remark 3.19 below.

3.19 Remark.
  1. (a)

    Let Ω\Omega be a discrete space. Then (Cb(Ω),,τco)(\mathrm{C}_{\operatorname{b}}(\Omega),\|\cdot\|_{\infty},\tau_{\operatorname{co}}) is a complete, semi-reflexive Saks space by [7, II.1.24 Remark 4), p. 88–89].

  2. (b)

    Let (X,)(X,\|\cdot\|) be a Banach space. Then (X,X,σ)(X^{\prime},\|\cdot\|_{X^{\prime}},\sigma^{\ast}) is a complete, semi-reflexive Saks space.

  3. (c)

    Let (X,)(X,\|\cdot\|) be a Banach space. Then (X,X,μ)(X^{\prime},\|\cdot\|_{X^{\prime}},\mu^{\ast}) is a complete, semi-reflexive Saks space where the semi-reflexivity follows from (X,μ)′′=X(X^{\prime},\mu^{\ast})^{\prime\prime}=X^{\prime} by the Mackey–Arens theorem.

  4. (d)

    Let (X,X)(X,\|\cdot\|_{X}) and (Y,Y)(Y,\|\cdot\|_{Y}) be Banach spaces and YY finite-dimensional. Then ((X;Y),(X;Y),τsot)(\mathcal{L}(X;Y),\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}}) is a complete, semi-reflexive Saks space.

  5. (e)

    Let HH be a separable Hilbert space. Then ((H),(H),τsot)(\mathcal{L}(H),\|\cdot\|_{\mathcal{L}(H)},\tau_{\operatorname{sot}^{\ast}}) is a complete, semi-reflexive Saks space where the semi-reflexivity follows from ((H),βsot)′′=𝒩(H)=(H)(\mathcal{L}(H),\beta_{\operatorname{sot}^{\ast}})^{\prime\prime}=\mathcal{N}(H)^{\prime}=\mathcal{L}(H).

  6. (f)

    Let Ω\Omega be a completely regular Hausdorff space. Then we have that the triple (Mt(Ω),Mt(Ω),σ(Mt(Ω),Cb(Ω)))(\mathrm{M}_{\operatorname{t}}(\Omega),\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)},\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega))) is a complete, semi-reflexive Saks space.

3.20 Example.

Due to Example 3.16 and Remark 3.19 (a) (,,τco)(\ell^{\infty},\|\cdot\|_{\infty},\tau_{\operatorname{co}}) is a complete, semi-reflexive Saks space. Therefore we may also apply Theorem 3.18 (a) to prove that the multiplication operator (Mq,D(Mq))(M_{q},D(M_{q})) with supnReq(n)0\sup_{n\in\mathbb{N}}\operatorname{Re}q(n)\leq 0 generates a μ(,1)\mu(\ell^{\infty},\ell^{1})-strongly continuous, μ(,1)\mu(\ell^{\infty},\ell^{1})-equicontinuous semigroup on \ell^{\infty} (we already checked in Example 3.14 that the other assumptions of Theorem 3.18 are satisfied).

We close this section with a characterisation of the bi-continuous semigroups with dissipative generators. First, we start with a refinement of Theorem 3.9. In the case υ=τ\upsilon=\tau this was already done in [5, Proposition 3.11, p. 9] whose prove needs some adaptations in the case of more general Hausdorff locally convex topologies υ\upsilon with τυτ\tau\subseteq\upsilon\subseteq\tau_{\|\cdot\|} for sequentially complete Saks spaces (X,,τ)(X,\|\cdot\|,\tau).

3.21 Proposition.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space, υ\upsilon a Hausdorff locally convex topology on XX with τυτ\tau\subseteq\upsilon\subseteq\tau_{\|\cdot\|} such that γ\gamma-convergent sequences are υ\upsilon-convergent, and (A,D(A))(A,D(A)) bi-densely defined. Then the following assertions are equivalent:

  1. (a)

    (A,D(A))(A,D(A)) generates a τ\tau-bi-continuous contraction semigroup (T(t))t0(T(t))_{t\geq 0} on XX and there exists a norming directed system of continuous seminorms Γυ\Gamma_{\upsilon} that generates υ\upsilon such that p(T(t)x)p(x)p(T(t)x)\leq p(x) for all t0t\geq 0, pΓυp\in\Gamma_{\upsilon} and xXx\in X.

  2. (b)

    λA\lambda-A is surjective for some λ>0\lambda>0 and (A,D(A))(A,D(A)) is a Γυ\Gamma_{\upsilon}-dissipative operator on XX for some norming directed system of continuous seminorms Γυ\Gamma_{\upsilon} that generates υ\upsilon.

Proof.

(a)\Rightarrow(b): First, we show that (A,D(A))(A,D(A)) is Γυ\Gamma_{\upsilon}-dissipative. We note that (0,)ρ(A)(0,\infty)\subseteq\rho(A) and

R(λ,A)x=0eλtT(t)xdtR(\lambda,A)x=\int_{0}^{\infty}\mathrm{e}^{-\lambda t}T(t)x\mathrm{d}t

for all λ>0\lambda>0 and xXx\in X by [23, Theorem 12, p. 215] and [23, Definition 9, p. 213] where the integral is an improper τ\tau-Riemann integral. The sequence of Riemann sums that approximate the integral on the right-hand side w.r.t. τ\tau are \|\cdot\|-bounded for each λ>0\lambda>0 and xXx\in X. Due to Remark 2.9 (a) this means that this sequence of Riemann sums is actually γ\gamma-convergent and thus υ\upsilon-convergent by assumption. Therefore we have for all λ>0\lambda>0, pΓυp\in\Gamma_{\upsilon} and xXx\in X that

p(R(λ,A)x)0eλtp(T(t)x)dt0eλtp(x)dt=1λp(x)p\left(R(\lambda,A)x\right)\leq\int_{0}^{\infty}\mathrm{e}^{-\lambda t}p(T(t)x)\mathrm{d}t\leq\int_{0}^{\infty}\mathrm{e}^{-\lambda t}p(x)\mathrm{d}t=\frac{1}{\lambda}p(x)

where we used that pp is υ\upsilon-continuous for the first inequality. Hence (A,D(A))(A,D(A)) is Γυ\Gamma_{\upsilon}-dissipative. In combination with Theorem 3.9 this yields that λA\lambda-A is surjective for some λ>0\lambda>0.

(b)\Rightarrow(a): Due to Theorem 3.9, (A,D(A))(A,D(A)) generates a τ\tau-bi-continuous contraction semigroup on XX, and (0,)ρ(A)(0,\infty)\subseteq\rho(A) by Proposition 3.7 (c). Furthermore, we have by the Post–Widder inversion formula [22, Corollary 2.10, p. 47] that

T(t)x=τ-limn(ntR(nt,A))nxT(t)x=\tau\text{-}\lim_{n\to\infty}\left(\frac{n}{t}R\left(\frac{n}{t},A\right)\right)^{n}x

for all t>0t>0 and xXx\in X. As a consequence of Remark 3.5 the τ\tau-convergent sequence ((ntR(nt,A))nx)n((\tfrac{n}{t}R(\tfrac{n}{t},A))^{n}x)_{n\in\mathbb{N}} is \|\cdot\|-bounded for each t>0t>0 and xXx\in X, thus γ\gamma-convergent by Remark 2.9 (a) and so υ\upsilon-convergent by assumption. We deduce from (2) that for all t>0t>0, pΓυp\in\Gamma_{\upsilon} and xXx\in X it holds that

p(T(t)x)=limn(nt)np(R(nt,A)nx)(2)p(x)p(T(t)x)=\lim_{n\to\infty}\left(\frac{n}{t}\right)^{n}p\left(R\left(\frac{n}{t},A\right)^{n}x\right)\underset{\eqref{eq:res_cont}}{\leq}p(x)

where we used that pp is υ\upsilon-continuous for the first equality. Further, for t=0t=0 we have p(T(t)x)=p(x)p(T(t)x)=p(x). We conclude that statement (a) holds. ∎

The assumptions of Proposition 3.21 (a) are up to rescaling fulfilled for any τ\tau-bi-continuous semigroup on XX if it is υ\upsilon-equicontinuous for υ=τ\upsilon=\tau or γs\gamma_{s} or γ\gamma (the assumption on υ\upsilon-equicontinuity may fail if υ=τ\upsilon=\tau by Remark 3.25).

3.22 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a sequentially complete Saks space and υ=τ\upsilon=\tau, γs\gamma_{s} or γ\gamma. Then there exists a norming directed system of continuous seminorms that generates the topology υ\upsilon by Definition 2.1 (c) for υ=τ\upsilon=\tau, by Remark 2.6 for υ=γs\upsilon=\gamma_{s} and by Remark 2.3 for υ=γ\upsilon=\gamma. Further, let (T(t))t0(T(t))_{t\geq 0} be a τ\tau-bi-continuous, υ\upsilon-equicontinuous semigroup on XX and ω\omega\in\mathbb{R} be its type (see Definition 2.8 (iii)). By modifying the proof of [5, Remark 3.12, p. 9–10] one can show that for the τ\tau-bi-continuous, υ\upsilon-equicontinuous contraction semigroup (eωtT(t))t0(e^{-\omega t}T(t))_{t\geq 0} on XX there exists a norming directed system of continuous seminorms Γυ\Gamma_{\upsilon} that generates υ\upsilon such that p(eωtT(t)x)p(x)p(e^{-\omega t}T(t)x)\leq p(x) for all t0t\geq 0, pΓυp\in\Gamma_{\upsilon} and xXx\in X.

In addition, we have the following characterisation in the case of complete, C-sequential Saks spaces (X,,τ)(X,\|\cdot\|,\tau) and υ=γ\upsilon=\gamma.

3.23 Proposition.

Let (X,,τ)(X,\|\cdot\|,\tau) be a complete, C-sequential Saks space and (A,D(A))(A,D(A)) the generator of a τ\tau-bi-continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX. Then the following assertions are eqivalent:

  1. (a)

    (T(t))t0(T(t))_{t\geq 0} is γ\gamma-equicontinuous.

  2. (b)

    There is a directed system of continuous seminorms Γγ\Gamma_{\gamma} that generates the mixed topology γ\gamma such that (A,D(A))(A,D(A)) is Γγ\Gamma_{\gamma}-dissipative.

Proof.

By Remark 2.11 we know that (T(t))t0(T(t))_{t\geq 0} is quasi-γ\gamma-equicontinuous if (X,,τ)(X,\|\cdot\|,\tau) is a sequentially complete C-sequential Saks space. Hence the equivalence of the assertions (a) and (b) follows from [1, Propositions 4.2, 4.4, p. 933, 935]. ∎

The condition that (X,,τ)(X,\|\cdot\|,\tau) is a complete, C-sequential Saks space is quite often fulfilled, e.g. for the examples from Remark 3.13 under some minor constraints by [18, Remark 3.19, p. 14], [18, Remark 3.20 (c), p. 15], [18, Example 4.12, p. 24–25] and [18, Corollary 3.23 (b), p. 17].

3.24 Remark.
  1. (a)

    Let Ω\Omega be a hemicompact Hausdorff kk_{\mathbb{R}}-space or a Polish space. Then (Cb(Ω),,τco)(\mathrm{C}_{\operatorname{b}}(\Omega),\|\cdot\|_{\infty},\tau_{\operatorname{co}}) is a complete, C-sequential Saks space.

  2. (b)

    Let (X,)(X,\|\cdot\|) be a separable Banach space. Then (X,X,σ)(X^{\prime},\|\cdot\|_{X^{\prime}},\sigma^{\ast}) is a complete, C-sequential Saks space.

  3. (c)

    Let (X,)(X,\|\cdot\|) be an SWCG space (see [30, p. 387]), or a sequentially σ(X,X)\sigma(X,X^{\prime})-complete space with an almost shrinking basis (see [17, p. 75]). Then the triple (X,X,μ)(X^{\prime},\|\cdot\|_{X^{\prime}},\mu^{\ast}) is a complete, C-sequential Saks space.

  4. (d)

    Let (X,X)(X,\|\cdot\|_{X}) be a separable Banach space and (Y,Y)(Y,\|\cdot\|_{Y}) a Banach space. Then ((X;Y),(X;Y),τsot)(\mathcal{L}(X;Y),\|\cdot\|_{\mathcal{L}(X;Y)},\tau_{\operatorname{sot}}) is a complete, C-sequential Saks space.

  5. (e)

    Let HH be a separable Hilbert space. Then ((H),(H),τsot)(\mathcal{L}(H),\|\cdot\|_{\mathcal{L}(H)},\tau_{\operatorname{sot}^{\ast}}) is a complete, C-sequential Saks space.

  6. (f)

    Let Ω\Omega be a Polish space. Then (Mt(Ω),Mt(Ω),σ(Mt(Ω),Cb(Ω)))(\mathrm{M}_{\operatorname{t}}(\Omega),\|\cdot\|_{\mathrm{M}_{\operatorname{t}}(\Omega)},\sigma(\mathrm{M}_{\operatorname{t}}(\Omega),\mathrm{C}_{\operatorname{b}}(\Omega))) is a complete, C-sequential Saks space.

3.25 Remark.

Let (X,,τ)(X,\|\cdot\|,\tau) be a complete, C-sequential Saks space. Due to Remark 2.11 assertion (a) of Proposition 3.23 always holds up to rescaling, and thus for any τ\tau-bi-continuous semigroup (T(t))t0(T(t))_{t\geq 0} on XX there is a rescaling such that the generator of the rescaled semigroup is Γγ\Gamma_{\gamma}-dissipative for some system of continuous seminorms Γγ\Gamma_{\gamma} that generates the mixed topology γ\gamma.

On the other hand, there are important examples of τ\tau-bi-continuous semigroups that are not quasi-τ\tau-equicontinuous. For instance, the Gauß–Weierstraß semigroup on the complete, C-sequential Saks space (Cb(d),,τco)(\mathrm{C}_{\operatorname{b}}(\mathbb{R}^{d}),\|\cdot\|_{\infty},\tau_{\operatorname{co}}) is τco\tau_{\operatorname{co}}-bi-continuous but not locally τco\tau_{\operatorname{co}}-equicontinuous by [23, Examples 6 (a), p. 209–210]. Since there is some λ>0\lambda>0 such that λA\lambda-A is surjective for the generator (A,D(A))(A,D(A)) of the Gauß–Weierstraß semigroup by [23, Lemma 7, Proposition 8, Theorem 12, p. 211–212, 215], it follows from Proposition 3.21 with υ=τco\upsilon=\tau_{\operatorname{co}} that its generator (A,D(A))(A,D(A)) is not bi-dissipative (even after rescaling, cf. [5, Example 3.9, p. 8]). Another example of a τco\tau_{\operatorname{co}}-bi-continuous semigroup which has no bi-dissipative generator (even after rescaling) is the left translation semigroup on the complete, C-sequential Saks space (Cb(),,τco)(\mathrm{C}_{\operatorname{b}}(\mathbb{R}),\|\cdot\|_{\infty},\tau_{\operatorname{co}}) which is τco\tau_{\operatorname{co}}-bi-continuous, even locally τco\tau_{\operatorname{co}}-equicontinuous but not quasi-τco\tau_{\operatorname{co}}-equicontinuous by [23, Examples 6 (b), p. 209–210] and [24, Example 3.2, p. 549].

Nevertheless, the Gauß–Weierstraß semigroup and the left translation semigroup are quasi-β0\beta_{0}-equicontinuous by Remark 2.11 and Remark 3.24 (a), and thus both have (after rescaling) a Γβ0\Gamma_{\beta_{0}}-dissipative generator for some system of continuous seminorms Γβ0\Gamma_{\beta_{0}} that generates the mixed topology β0=γ(,τco)\beta_{0}=\gamma(\|\cdot\|_{\infty},\tau_{\operatorname{co}}). This underlines that in the framework of τ\tau-bi-continuous semigroups the concept of a bi-dissipative operator resp. generator is not the correct choice whereas the concept of a Γγ\Gamma_{\gamma}-dissipative operator resp. generator is the more reasonable one.

Acknowledgement

We would like to thank David Jornet for his kind explanations regarding the proof of [1, Proposition 3.11 (iii), p. 927].

References

  • Albanese and Jornet [2016] A.A. Albanese and D. Jornet. Dissipative operators and additive perturbations in locally convex spaces. Math. Nachr., 289(8-9):920–949, 2016. doi:10.1002/mana.201500150.
  • Budde [2021a] C. Budde. Positive Miyadera–Voigt perturbations of bi-continuous semigroups. Positivity, 25(3):1107–1129, 2021a. doi:10.1007/s11117-020-00806-1.
  • Budde [2021b] C. Budde. Positive Desch–Schappacher perturbations of bi-continuous semigroups on AM-spaces. Acta Sci. Math. (Szeged), 87(3-4):571–594, 2021b. doi:10.14232/actasm-021-914-5.
  • Budde and Farkas [2019] C. Budde and B. Farkas. Intermediate and extrapolated spaces for bi-continuous operator semigroups. J. Evol. Equ., 19(2):321–359, 2019. doi:10.1007/s00028-018-0477-8.
  • Budde and Wegner [2022 (Online first] C. Budde and S.-A. Wegner. A Lumer–Phillips type generation theorem for bi-continuous semigroups. Z. Anal. Anwend., pages 1–16, 2022 (Online first). doi:10.4171/ZAA/1695.
  • Conway [1967] J.B. Conway. The strict topology and compactness in the space of measures. II. Trans. Amer. Math. Soc., 126(3):474–486, 1967. doi:10.1090/S0002-9947-1967-0206685-2.
  • Cooper [1978] J.B. Cooper. Saks spaces and applications to functional analysis. North-Holland Math. Stud. 28. North-Holland, Amsterdam, 1978.
  • Engel and Nagel [2000] K.-J. Engel and R. Nagel. One-parameter semigroups for linear evolution equations. Grad. Texts in Math. 194. Springer, New York, 2000. doi:10.1007/b97696.
  • Es-Sarhir and Farkas [2006] A. Es-Sarhir and B. Farkas. Perturbation for a class of transition semigroups on the Hölder space Cb,locθ(H)C_{b,\operatorname{loc}}^{\theta}(H). J. Math. Anal. Appl., 315(2):666–685, 2006. doi:10.1016/j.jmaa.2005.04.024.
  • Fabian et al. [2011] 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.
  • Farkas [2003] B. Farkas. Perturbations of bi-continuous semigroups. PhD thesis, Eötvös Loránd University, Budapest, 2003.
  • Farkas [2004] B. Farkas. Perturbations of bi-continuous semigroups with applications to transition semigroups on Cb(H)C_{b}(H). Semigroup Forum, 68(1):87–107, 2004. doi:10.1007/s00233-002-0024-2.
  • Farkas [2011] B. Farkas. Adjoint bi-continuous semigroups and semigroups on the space of measures. Czech. Math. J., 61(2):309–322, 2011. doi:10.1007/s10587-011-0076-0.
  • Feller [1952] W. Feller. The parabolic differential equations and the associated semi-groups of transformations. Ann. of Math. (2), 55(3):468–519, 1952. doi:10.2307/1969644.
  • Glück et al. [2022] J. Glück, R. Nagel, and I. Remizov (eds.). Open problems in operator semigroup theory (OPSO 2021 and 2022), Part 2, 2022. https://nnov.hse.ru/data/2022/02/18/1747622322/Open_problems_in_OPSO_theory_part_2.pdf#page=4&zoom=auto,-466,795 (June 01, 2022).
  • Hille [1948] E. Hille. Functional analysis and semi-groups. Colloquium Publications 31. AMS, Providence, RI, 1948.
  • Kalton [1973] N.J. Kalton. Mackey duals and almost shrinking bases. Math. Proc. Cambridge Philos. Soc., 74(1):73–81, 1973. doi:10.1017/S0305004100047812.
  • Kruse and Schwenninger [2022a] K. Kruse and F.L. Schwenninger. On equicontinuity and tightness of bi-continuous semigroups. J. Math. Anal. Appl., 509(2):1–27, 2022a. doi:10.1016/j.jmaa.2021.125985.
  • Kruse and Schwenninger [2022b] K. Kruse and F.L. Schwenninger. Sun dual theory for bi-continuous semigroups, 2022b. arXiv preprint https://arxiv.org/abs/2203.12765v2.
  • Kruse and Seifert [2022] K. Kruse and C. Seifert. Final state observability and cost-uniform approximate null-controllability for bi-continuous semigroups, 2022. arXiv preprint https://arxiv.org/abs/2206.00562v1.
  • Kruse et al. [2021] K. Kruse, J. Meichsner, and C. Seifert. Subordination for sequentially equicontinuous equibounded C0C_{0}-semigroups. J. Evol. Equ., 21(2):2665–2690, 2021. doi:10.1007/s00028-021-00700-7.
  • Kühnemund [2001] F. Kühnemund. Bi-continuous semigroups on spaces with two topologies: Theory and applications. PhD thesis, Eberhard-Karls-Universität Tübingen, 2001.
  • Kühnemund [2003] F. Kühnemund. A Hille–Yosida theorem for bi-continuous semigroups. Semigroup Forum, 67(2):205–225, 2003. doi:10.1007/s00233-002-5000-3.
  • Kunze [2009] M. Kunze. Continuity and equicontinuity of semigroups on norming dual pairs. Semigroup Forum, 79(3):540–560, 2009. doi:10.1007/s00233-009-9174-9.
  • Lumer and Phillips [1961] G. Lumer and R.S. Phillips. Dissipative operators in a Banach space. Pacific J. Math., 11(2):679–698, 1961. doi:10.2140/pjm.1961.11.679.
  • Meise and Vogt [1997] R. Meise and D. Vogt. Introduction to functional analysis. Oxf. Grad. Texts Math. 2. Clarendon Press, Oxford, 1997.
  • Michael [1973] E.A. Michael. On kk-spaces, kRk_{R}-spaces and k(X)k(X). Pacific J. Math., 47(2):487–498, 1973. doi:10.2140/pjm.1973.47.487.
  • Miyadera [1952] I. Miyadera. Generation of a strongly continuous semi-group operators. Tohoku Math. J. (2), 4(2):109–114, 1952. doi:10.2748/tmj/1178245412.
  • Phillips [1952] R.S. Phillips. On the generation of semigroups of linear operators. Pacific J. Math., 2(3):343–369, 1952. doi:10.2140/pjm.1952.2.343.
  • Schlüchtermann and Wheeler [1988] G. Schlüchtermann and R.F. Wheeler. On strongly WCG Banach spaces. Math. Z., 199(3):387–398, 1988. doi:10.1007/BF01159786.
  • Snipes [1973] R.F. Snipes. C-sequential and S-bornological topological vector spaces. Math. Ann., 202(4):273–283, 1973. doi:10.1007/BF01433457.
  • Wilansky [1981] A. Wilansky. Mazur spaces. Int. J. Math. Math. Sci., 4(1):39–53, 1981. doi:10.1155/S0161171281000021.
  • Wiweger [1961] A. Wiweger. Linear spaces with mixed topology. Studia Math., 20(1):47–68, 1961. doi:10.4064/sm-20-1-47-68.
  • Yosida [1948] K. Yosida. On the differentiability and the representation of one-parameter semi-group of linear operators. J. Math. Soc. Japan, 1(1):15–21, 1948. doi:10.2969/jmsj/00110015.