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

On domination for (non-symmetric) Dirichlet forms

Liping Li Fudan University, Shanghai, China. [email protected]  and  Jiangang Ying Fudan University, Shanghai, China. [email protected]
Abstract.

The primary aim of this article is to investigate the domination relationship between two L2L^{2}-semigroups using probabilistic methods. According to Ouhabaz’s domination criterion, the domination of semigroups can be transformed into relationships involving the corresponding Dirichlet forms. Our principal result establishes the equivalence between the domination of Dirichlet forms and the killing transformation of the associated Markov processes, which generalizes and completes the results in [37] and [38]. Based on this equivalence, we provide a representation of the dominated Dirichlet form using the bivariate Revuz measure associated with the killing transformation and further characterize the sandwiched Dirichlet form within the broader Dirichlet form framework. In particular, our findings apply to the characterization of operators sandwiched between the Dirichlet Laplacian and the Neumann Laplacian. For the local boundary case, we eliminate all technical conditions identified in the literature [4] and deliver a complete representation of all sandwiched operators governed by a Robin boundary condition determined by a specific quasi-admissible measure. Additionally, our results offer a comprehensive characterization of related operators in the non-local Robin boundary case, specifically resolving an open problem posed in the literature [25].

Key words and phrases:
Domination of semigroups, Ouhabaz’s domination criterion, Laplace operators, Robin boundary conditions, Non-local boundary conditions, Dirichlet forms, Markov processes, Killing transformation, Multiplicative functionals, Bivariate Revuz measures, Silverstein extensions
2020 Mathematics Subject Classification:
Primary 31C25, 60J46, 60J57, 60J50, 47D07.
The first named author is a member of LMNS, Fudan University. He is partially supported by NSFC (No. 11931004 and 12371144).

1. Introduction

Consider an open set Ωn\Omega\subset{\mathbb{R}}^{n}. Let ΔD\Delta^{D} and ΔN\Delta^{N} denote the Laplace operators on L2(Ω)L^{2}(\Omega) subject to the Dirichlet boundary condition u|Ω=0u|_{\partial\Omega}=0 and the Neumann boundary condition un|Ω=0\frac{\partial u}{\partial n}|_{\partial\Omega}=0, respectively. Here, L2(Ω)L^{2}(\Omega) denotes the space of all square-integrable functions on Ω\Omega, Ω\partial\Omega represents the boundary of Ω\Omega, and un\frac{\partial u}{\partial n} denotes the exterior normal derivative on the boundary (whose existence requires certain regularity assumptions). Denote by (etΔD)t0(e^{t\Delta^{D}})_{t\geq 0} and (etΔN)t0(e^{t\Delta^{N}})_{t\geq 0} the strongly continuous semigroups generated by ΔD\Delta^{D} and ΔN\Delta^{N}, respectively. In [4], Arendt and Warma investigated the question of which self-adjoint operator generates a semigroup (Tt)t0(T_{t})_{t\geq 0} that is “sandwiched” between (etΔD)t0(e^{t\Delta^{D}})_{t\geq 0} and (etΔN)t0(e^{t\Delta^{N}})_{t\geq 0} in the sense that

etΔDfTtfetΔNf,t0,fpL2(Ω),e^{t\Delta^{D}}f\leq T_{t}f\leq e^{t\Delta^{N}}f,\quad t\geq 0,f\in\mathrm{p}L^{2}(\Omega), (1.1)

where pL2(Ω)\mathrm{p}L^{2}(\Omega) represents the family of all positive functions in L2(Ω)L^{2}(\Omega). Their main result demonstrates that, under certain necessary conditions, (Tt)t0(T_{t})_{t\geq 0} corresponds to the Laplace operator subject to a Robin boundary condition determined by a certain admissible measure μ\mu on Ω\partial\Omega (see Example 7.6). This operator and its semigroup are denoted by Δμ\Delta_{\mu} and etΔμe^{t\Delta_{\mu}}, respectively. Specifically, when μ\mu is absolutely continuous with respect to the surface measure on Ω\partial\Omega with density β\beta, Δμ\Delta_{\mu} corresponds to the classical Robin boundary condition:

βu+un=0on Ω.\beta u+\frac{\partial u}{\partial n}=0\quad\text{on }\partial\Omega.

Arendt and Warma’s research builds upon Ouhabaz’s work [24], which provides a quadratic form characterization of the domination relation in (1.1). Specifically, if (a1,𝒟(a1))(a^{1},\mathcal{D}(a^{1})) and (a2,𝒟(a2))(a^{2},\mathcal{D}(a^{2})) are the closed forms associated with the strongly continuous semigroups (Tt1)t0(T^{1}_{t})_{t\geq 0} and (Tt2)t0(T^{2}_{t})_{t\geq 0}, and both semigroups are positive (i.e., for any non-negative fL2(Ω)f\in L^{2}(\Omega), Tt1f0T^{1}_{t}f\geq 0 and Tt2f0T^{2}_{t}f\geq 0), then

Tt1fTt2f,fpL2(Ω)T^{1}_{t}f\leq T^{2}_{t}f,\quad\forall f\in\mathrm{p}L^{2}(\Omega) (1.2)

is equivalent to the conditions that 𝒟(a1)\mathcal{D}(a^{1}) is an ideal of 𝒟(a2)\mathcal{D}(a^{2}) and that

a1(u,v)a2(u,v),u,vp𝒟(a1),a^{1}(u,v)\geq a^{2}(u,v),\quad u,v\in\mathrm{p}\mathcal{D}(a^{1}),

where p𝒟(a1)\mathrm{p}\mathcal{D}(a^{1}) denotes the set of non-negative elements in 𝒟(a1)\mathcal{D}(a^{1}) (see Lemma 2.1 for details).

In the context of (1.1), since the semigroups on both sides are positive and satisfy the (sub-)Markovian property (i.e., for any fL2(Ω)f\in L^{2}(\Omega) with 0f10\leq f\leq 1, it holds that 0etΔDf,etΔNf10\leq e^{t\Delta^{D}}f,e^{t\Delta^{N}}f\leq 1), it follows that the sandwiched semigroup is also positive and Markovian. According to the foundational work of Beurling and Deny [6], semigroups with these properties correspond to closed forms known as Dirichlet forms.

Over the past two decades, the study of sandwiched semigroups and quadratic forms has been advanced and applied from various perspectives. To illustrate, we highlight a selection of contributions. While much of the research continues to focus on Laplace operators (or elliptic differential operators), the scope of boundary conditions has expanded to include non-local settings, as explored in [2, 25, 35]. Additionally, the focus has broadened to encompass general Dirichlet forms, with notable works including [5, 18, 22, 26, 29]. Recent years have also seen growing interest in extending these results to nonlinear semigroups and nonlinear Dirichlet forms, as exemplified by [11]. Beyond these theoretical advances, related research has found applications in areas such as partial differential equations, fractal structures, and other fields, as demonstrated in [8, 17, 28].

It is important to note that, under conditions of regularity or quasi-regularity, the relationship between Dirichlet forms and Markov processes in probability theory is well established (see [9, 15, 20]). However, to our knowledge, aside from the mention of the probabilistic counterparts of Dirichlet forms in [1], the aforementioned studies utilizing Ouhabaz’s domination criterion have primarily adopted an analytical perspective, largely overlooking the substantial potential of probabilistic methods, despite most discussions occurring within the framework of regularity. Researchers familiar with Markov processes can readily relate the domination relationship of L2L^{2}-semigroups from (1.2) to the subordination between two Markov transition semigroups: (Qt)t0(Q_{t})_{t\geq 0} is said to be subordinated to (Pt)t0(P_{t})_{t\geq 0} if

QtfPtf,fp.Q_{t}f\leq P_{t}f,\quad\forall f\in\mathrm{p}\mathscr{B}. (1.3)

This represents a classic topic in the theory of general Markov processes (see [7, III] and [30, VII]). In simpler terms, the subordination stated in (1.3) corresponds to the killing transformation of Markov processes. The Markov process corresponding to (Qt)t0(Q_{t})_{t\geq 0}, often referred to as the subprocess, is always derived from the Markov process corresponding to (Pt)t0(P_{t})_{t\geq 0} through a mechanism governed by some multiplicative functional that terminates sample paths. The Dirichlet form characterization of general killing transformations has also been addressed in prior studies (such as [38, 37]). Nevertheless, it is crucial to acknowledge that there exists a significant disparity between domination in the sense of almost everywhere from (1.1) and subordination in the pointwise sense from (1.3). Fortunately, the theory of regular Dirichlet forms established by Fukushima, along with the quasi-regular Dirichlet form framework developed by Ma et al. (see [20]), is specifically designed to bridge this gap. Therefore, at least theoretically, there exists a probabilistic pathway to investigate domination in (1.2).

The primary objective of this paper is precisely to establish the equivalence between the domination (1.2) of L2L^{2}-semigroups in analysis and the killing transformation of Markov processes in probability. Our principal result demonstrates the following intuitive fact: each dominated Dirichlet form is derived from two successive killing transformations. The first step involves terminating the Markov process when it exits a specific finely open set, which defines the actual state space of the subprocess corresponding to the dominated Dirichlet form. The second step consists of an additional killing transformation dictated by another multiplicative functional within the finely open set, which does not further alter the state space of the subprocess. At the same time, these two steps of killing transformations correspond to the analytical decomposition of domination described in Theorem 2.4: Every domination can be expressed as a composition of a Silverstein extension and strong subordination.

Extending beyond Ouhabaz’s domination criterion, our probabilistic characterization offers a representation of the dominated Dirichlet form using the bivariate Revuz measure of the multiplicative functional (which can be naturally extended to represent the sandwiched Dirichlet form, see Theorem 6.2). Specifically, for any (non-symmetric) Dirichlet forms (a1,𝒟(a1))(a^{1},\mathcal{D}(a^{1})) and (a2,𝒟(a2))(a^{2},\mathcal{D}(a^{2})) satisfying (1.2), there exists a positive measure σ\sigma on Ω¯×Ω¯\overline{\Omega}\times\overline{\Omega}, which can be expressed as detailed in (3.8), such that

𝒟(a1)=𝒟(a2)L2(Ω¯,σ¯),\displaystyle\mathcal{D}(a^{1})=\mathcal{D}(a^{2})\cap L^{2}(\overline{\Omega},\bar{\sigma}), (1.4)
a1(u,v)=a2(u,v)+σ(uv),u,v𝒟(a1),\displaystyle a^{1}(u,v)=a^{2}(u,v)+\sigma(u\otimes v),\quad u,v\in\mathcal{D}(a^{1}),

where σ¯\bar{\sigma} denotes the marginal measure of σ\sigma. It is noteworthy that [5] has achieved a similar representation for regular and symmetric Dirichlet forms using analytical methods.

Our investigation is situated within the general framework of non-symmetric Dirichlet forms, with the essential condition being the quasi-regularity of the dominated Dirichlet form (see §5). In other words, in the representation provided in (1.4), the quasi-regularity of (a1,𝒟(a1))(a_{1},\mathcal{D}(a_{1})) is necessary, whereas (a2,𝒟(a2))(a_{2},\mathcal{D}(a_{2})) does not need to satisfy the quasi-regularity condition. Consequently, the findings presented in this paper are applicable to the majority of situations discussed in the existing literature, including works such as [1, 5, 18, 22, 26, 29], among others. In particular, we re-examine the Laplacian-related sandwiched Dirichlet forms in §7. Theorem 7.3 eliminates all assumptions imposed by Arendt and Warma [4] concerning the characterization of sandwiched semigroups in (1.1): In fact, every semigroup (Tt)t0(T_{t})_{t\geq 0} satisfying (1.1) is determined by some quasi-admissible measure μ\mu associated with the Robin Laplacian. (The definition of quasi-admissibility can be found in Definition 7.1.) Moreover, comprehensive solutions can also be provided for Laplacian-related problems in the non-local Robin boundary case. Notably, our results address an open problem posed in the literature [25], as discussed in §7.3.

Finally, we provide a brief description of commonly used symbols. The concepts and notations associated with Dirichlet forms align with those presented in [9, 15, 20], among others. For a function uu in a Dirichlet space, u~\tilde{u} denotes its quasi-continuous modification by default. Given a topological space EE, (E)\mathscr{B}(E) represents the collection of Borel measurable sets or Borel measurable functions, and (E)\mathscr{B}^{*}(E) denotes the collection of all universally measurable sets or universally measurable functions. For a class of functions 𝒞\mathscr{C}, p𝒞\mathrm{p}\mathscr{C} and b𝒞\mathrm{b}\mathscr{C} denote the families of non-negative elements and bounded elements, respectively, within 𝒞\mathscr{C}.

2. Decomposition of domination

Let EE be a measurable space, and let mm be a σ\sigma-finite measure on EE. The norms of L2(E,m)L^{2}(E,m) and L(E,m)L^{\infty}(E,m) are denoted by 2\|\cdot\|_{2} and \|\cdot\|_{\infty}, respectively, while the inner product of L2(E,m)L^{2}(E,m) is denoted by (,)m(\cdot,\cdot)_{m}. A (non-symmetric) Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(E,m)L^{2}(E,m) is defined as a coercive closed form that satisfies the Markovian property, as described in, e.g., [20, I, Definition 4.5]. The 1{\mathscr{E}}_{1}-norm on {\mathscr{F}} is denoted by 1\|\cdot\|_{{\mathscr{E}}_{1}}, where f1:=((f,f)+(f,f)m)1/2\|f\|_{{\mathscr{E}}_{1}}:=\left({\mathscr{E}}(f,f)+(f,f)_{m}\right)^{1/2} for all ff\in{\mathscr{F}}. The Dirichlet form is called symmetric if, in addition, (f,g)=(g,f){\mathscr{E}}(f,g)={\mathscr{E}}(g,f) for all f,gf,g\in{\mathscr{F}}.

For a given Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(E,m)L^{2}(E,m), it is well known that b:=L(E,m)\mathrm{b}{\mathscr{F}}:={\mathscr{F}}\cap L^{\infty}(E,m) is a subalgebra of L(E,m)L^{\infty}(E,m) (see [20, I, Corollary 4.15]), and {\mathscr{F}} forms a sublattice of L2(E,m)L^{2}(E,m) (see [20, I, Proposition 4.11]).

If U,VU,V are subalgebras of L(E,m)L^{\infty}(E,m), we say that UU is an algebraic ideal in VV if fUf\in U and gVg\in V imply fgUfg\in U. For sublattices U,VU,V of L2(E,m)L^{2}(E,m), UU is called an order ideal in VV if fU,gVf\in U,g\in V, and |g||f||g|\leq|f| imply gUg\in U.

For two Dirichlet forms (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}), we say that {\mathscr{E}} dominates {\mathscr{E}}^{\prime} (or equivalently, {\mathscr{E}}^{\prime} is dominated by {\mathscr{E}}) and write (,)(,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})\preceq({\mathscr{E}},{\mathscr{F}}) or {\mathscr{E}}^{\prime}\preceq{\mathscr{E}} if either of the conditions in the following lemma is satisfied. The second condition in this lemma is commonly referred to as the Ouhabaz domination criterion (see [24]) for the domination of semigroups.

Lemma 2.1.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two Dirichlet forms on L2(E,m)L^{2}(E,m), whose associated semigroups (as described in [20, I, Remark 2.9(ii)]) are denoted by (Tt)t0(T_{t})_{t\geq 0} and (Tt)t0(T^{\prime}_{t})_{t\geq 0}, respectively. The following assertions are equivalent:

  • (1)

    TtfTtfT^{\prime}_{t}f\leq T_{t}f for all t0t\geq 0 and all non-negative fL2(E,m)f\in L^{2}(E,m).

  • (2)

    {\mathscr{F}}^{\prime}\subset{\mathscr{F}}, b:=L(E,m)\mathrm{b}{\mathscr{F}}^{\prime}:={\mathscr{F}}^{\prime}\cap L^{\infty}(E,m) is an algebraic ideal in b\mathrm{b}{\mathscr{F}}, and

    (f,g)(f,g){\mathscr{E}}^{\prime}(f,g)\geq{\mathscr{E}}(f,g) (2.1)

    for all non-negative f,gf,g\in{\mathscr{F}}^{\prime}.

  • (3)

    {\mathscr{F}}^{\prime}\subset{\mathscr{F}}, {\mathscr{F}}^{\prime} is an order ideal in {\mathscr{F}}, and (2.1) holds for all non-negative f,gf,g\in{\mathscr{F}}^{\prime}.

Proof.

The equivalence between the first and third assertions was established in [22, Corollary 4.3]. The proof of (3)\Rightarrow(2) is straightforward. It suffices to demonstrate that, under the assumptions of the third assertion, if fbf\in\mathrm{b}{\mathscr{F}}^{\prime} and gbg\in\mathrm{b}{\mathscr{F}}, then fgbfg\in\mathrm{b}{\mathscr{F}}^{\prime}. This follows immediately, as |fg|g|f||fg|\leq\|g\|_{\infty}|f| and {\mathscr{F}}^{\prime} is an order ideal in {\mathscr{F}}.

The proof of (2)\Rightarrow(3) can be completed by repeating the argument in [29, Lemma 2.2], where only the symmetric case is considered. For the convenience of the readers, we restate the necessary details. Let ff\in{\mathscr{F}} and gg\in{\mathscr{F}}^{\prime} with |f||g||f|\leq|g|. Our goal is to show that ff\in{\mathscr{F}}^{\prime}. Since {\mathscr{F}} and {\mathscr{F}}^{\prime} are lattices, and by [20, I, Proposition 4.17], we may assume without loss of generality that 0fg0\leq f\leq g and ff and gg are bounded. Let A:={(x,y2:0xy}A:=\{(x,y\in\mathbb{R}^{2}:0\leq x\leq y\} and ε>0\varepsilon>0. Consider the function

Cε:A,Cε(x,y):=xyy+ε.C_{\varepsilon}:A\rightarrow\mathbb{R},\quad C_{\varepsilon}(x,y):=\frac{xy}{y+\varepsilon}.

It satisfies the following inequality

|Cε(x1,y1)Cε(x2,y2)||x1x2|+|y1y2|\left|C_{\varepsilon}(x_{1},y_{1})-C_{\varepsilon}(x_{2},y_{2})\right|\leq|x_{1}-x_{2}|+|y_{1}-y_{2}| (2.2)

for (x1,y1),(x2,y2)A(x_{1},y_{1}),(x_{2},y_{2})\in A. The function Hε(y):=yy+εH_{\varepsilon}(y):=\frac{y}{y+\varepsilon} is Lipschitz continuous with Hε<1ε\|H^{\prime}_{\varepsilon}\|_{\infty}<\frac{1}{\varepsilon} and Hε(0)=0H_{\varepsilon}(0)=0. By [20, I, Proposition 4.11], we have Hε(g)bH_{\varepsilon}(g)\in\mathrm{b}{\mathscr{F}}^{\prime}. Therefore, by the second assertion, hε:=Cε(f,g)=fHε(g)bh_{\varepsilon}:=C_{\varepsilon}(f,g)=f\cdot H_{\varepsilon}(g)\in\mathrm{b}{\mathscr{F}}^{\prime}. It is evident that hεfh_{\varepsilon}\rightarrow f in L2(E,m)L^{2}(E,m) as ε0\varepsilon\downarrow 0. We now aim to show that

supε>0(hε,hε)<,\sup_{\varepsilon>0}{\mathscr{E}}^{\prime}(h_{\varepsilon},h_{\varepsilon})<\infty,

which, by the Banach-Saks theorem, will imply that ff\in{\mathscr{F}}^{\prime}. Indeed, it follows from 0hεfg0\leq h_{\varepsilon}\leq f\leq g and (2.1) that

(g,g)\displaystyle{\mathscr{E}}^{\prime}(g,g) =(hε+(ghε),hε+(ghε))\displaystyle={\mathscr{E}}^{\prime}\left(h_{\varepsilon}+(g-h_{\varepsilon}),h_{\varepsilon}+(g-h_{\varepsilon})\right)
(hε,hε)+(g,g)(hε,hε).\displaystyle\geq{\mathscr{E}}^{\prime}(h_{\varepsilon},h_{\varepsilon})+{\mathscr{E}}(g,g)-{\mathscr{E}}(h_{\varepsilon},h_{\varepsilon}).

Using the property (2.2) and applying [20, I, Proposition 4.11] to (,)({\mathscr{E}},{\mathscr{F}}), we get

(hε,hε)1/2(f,f)1/2+(g,g)1/2.{\mathscr{E}}(h_{\varepsilon},h_{\varepsilon})^{1/2}\leq{\mathscr{E}}(f,f)^{1/2}+{\mathscr{E}}(g,g)^{1/2}.

Thus, we have

supε>0(hε,hε)\displaystyle\sup_{\varepsilon>0}{\mathscr{E}}^{\prime}(h_{\varepsilon},h_{\varepsilon}) supε>0(hε,hε)+(g,g)(g,g)\displaystyle\leq\sup_{\varepsilon>0}{\mathscr{E}}(h_{\varepsilon},h_{\varepsilon})+{\mathscr{E}}^{\prime}(g,g)-{\mathscr{E}}(g,g)
((f,f)1/2+(g,g)1/2)2+(g,g)(g,g)<.\displaystyle\leq\left({\mathscr{E}}(f,f)^{1/2}+{\mathscr{E}}(g,g)^{1/2}\right)^{2}+{\mathscr{E}}^{\prime}(g,g)-{\mathscr{E}}(g,g)<\infty.

This completes the proof. ∎

We introduce two additional notions.

Definition 2.2.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two Dirichlet forms on L2(E,m)L^{2}(E,m).

  • (1)

    We say that {\mathscr{E}}^{\prime} is subordinate to {\mathscr{E}} if {\mathscr{F}}^{\prime}\subset{\mathscr{F}} and (2.1) holds for all non-negative f,gf,g\in{\mathscr{F}}^{\prime}. Furthermore, {\mathscr{E}}^{\prime} is strongly subordinate to {\mathscr{E}} if, in addition, {\mathscr{F}}^{\prime} is dense in {\mathscr{F}} with respect to the 1{\mathscr{E}}_{1}-norm.

  • (2)

    We say that {\mathscr{E}} is a Silverstein extension of {\mathscr{E}}^{\prime} if b\mathrm{b}{\mathscr{F}}^{\prime} is an algebraic ideal in b\mathrm{b}{\mathscr{F}} and ={\mathscr{E}}={\mathscr{E}}^{\prime} on b×b\mathrm{b}{\mathscr{F}}^{\prime}\times\mathrm{b}{\mathscr{F}}^{\prime}.

Remark 2.3.

The notions of (strong) subordination were discussed in [38], where the author, under the additional assumption of quasi-regularity, examined the relationship between the (strong) subordination of Dirichlet forms and the killing transformation of the associated Markov processes. The second concept, introduced by Silverstein (see [32, 33]), is standard in the theory of (symmetric) Dirichlet forms. For more detailed information, readers are referred to [9, §6.6]. However, it is important to emphasize that our discussion is not restricted to symmetric Dirichlet forms.

Clearly, {\mathscr{E}}^{\prime}\preceq{\mathscr{E}} if and only if {\mathscr{E}}^{\prime} is subordinate to {\mathscr{E}} and b\mathrm{b}{\mathscr{F}}^{\prime} is an algebraic ideal in b\mathrm{b}{\mathscr{F}} (or equivalently, {\mathscr{F}}^{\prime} is an order ideal in {\mathscr{F}}). By the proof of Lemma 2.1, {\mathscr{E}} is a Silverstein extension of {\mathscr{E}}^{\prime} if and only if {\mathscr{F}}^{\prime} is an order ideal in {\mathscr{F}} and ={\mathscr{E}}={\mathscr{E}}^{\prime} on b×b\mathrm{b}{\mathscr{F}}^{\prime}\times\mathrm{b}{\mathscr{F}}^{\prime}. In particular, if {\mathscr{E}} is a Silverstein extension of {\mathscr{E}}^{\prime}, then {\mathscr{E}}^{\prime}\preceq{\mathscr{E}} and {\mathscr{E}}^{\prime} is subordinate to {\mathscr{E}}.

We now present a simple yet intriguing result, which demonstrates that domination can be uniquely decomposed into a combination of strong subordination and Silverstein extension.

Theorem 2.4.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two Dirichlet forms on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. Denote ~\tilde{{\mathscr{F}}} as the closure of {\mathscr{F}}^{\prime} in {\mathscr{F}} with respect to the 1{\mathscr{E}}_{1}-norm, and define

~(f,g):=(f,g),f,g~.\tilde{{\mathscr{E}}}(f,g):={\mathscr{E}}(f,g),\quad f,g\in\tilde{{\mathscr{F}}}.

Then, {\mathscr{E}}^{\prime} is strongly subordinate to ~\tilde{{\mathscr{E}}}, and {\mathscr{E}} is a Silverstein extension of ~\tilde{{\mathscr{E}}}. Furthermore, if (~1,~1)(\tilde{\mathscr{E}}^{1},\tilde{\mathscr{F}}^{1}) is another Dirichlet form on L2(E,m)L^{2}(E,m) that satisfies these two conditions, then (~1,~1)=(~,~)(\tilde{\mathscr{E}}^{1},\tilde{\mathscr{F}}^{1})=(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}).

Proof.

The first assertion that {\mathscr{E}}^{\prime} is strongly subordinate to ~\tilde{{\mathscr{E}}} follows directly from the definition of (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). We will now demonstrate that {\mathscr{E}} is a Silverstein extension of ~\tilde{{\mathscr{E}}}. Consider fb~:=~L(E,m)f\in\mathrm{b}\tilde{{\mathscr{F}}}:=\tilde{{\mathscr{F}}}\cap L^{\infty}(E,m) and gbg\in\mathrm{b}{\mathscr{F}}. Our objective is to establish that fgb~fg\in\mathrm{b}\tilde{{\mathscr{F}}}. Indeed, we may take a sequence fnbf_{n}\in\mathrm{b}{\mathscr{F}}^{\prime} such that 1(fnf,fnf)0{\mathscr{E}}_{1}(f_{n}-f,f_{n}-f)\rightarrow 0, supn1fn<\sup_{n\geq 1}\|f_{n}\|_{\infty}<\infty, and fnf_{n} converges to ff, mm-a.e. Since b\mathrm{b}{\mathscr{F}}^{\prime} is an algebraic ideal in b\mathrm{b}{\mathscr{F}}, it follows that fngbb~f_{n}g\in\mathrm{b}{\mathscr{F}}^{\prime}\subset\mathrm{b}\tilde{{\mathscr{F}}}. Furthermore, according to [15, Theorem 1.4.2], we have

supn1~1(fng,fng)1/2supn1fn1(g,g)+gsupn11(fn,fn)<.\sup_{n\geq 1}\tilde{{\mathscr{E}}}_{1}(f_{n}g,f_{n}g)^{1/2}\leq\sup_{n\geq 1}\|f_{n}\|_{\infty}\cdot\sqrt{{\mathscr{E}}_{1}(g,g)}+\|g\|_{\infty}\sup_{n\geq 1}\sqrt{{\mathscr{E}}_{1}(f_{n},f_{n})}<\infty.

By the Banach-Saks theorem, there exists a subsequence {fnkg}\{f_{n_{k}}g\} of {fng}\{f_{n}g\} and h~h\in\tilde{{\mathscr{F}}} such that

1Nk=1Nfnkgh,N\frac{1}{N}\sum_{k=1}^{N}f_{n_{k}}g\rightarrow h,\quad N\rightarrow\infty

with respect to the ~1\tilde{{\mathscr{E}}}_{1}-norm. Note that fnff_{n}\rightarrow f, mm-a.e. Therefore, h=fgh=fg, mm-a.e., confirming that fgb~fg\in\mathrm{b}\tilde{{\mathscr{F}}}.

Let (~1,~1)(\tilde{{\mathscr{E}}}^{1},\tilde{{\mathscr{F}}}^{1}) be another Dirichlet form that satisfies the two specified conditions. Since {\mathscr{E}} is a Silverstein extension of ~1\tilde{{\mathscr{E}}}^{1}, it follows that ~1(f,g)=(f,g)\tilde{{\mathscr{E}}}^{1}(f,g)={\mathscr{E}}(f,g) for all f,g~1f,g\in\tilde{{\mathscr{F}}}^{1}. The strong subordination of {\mathscr{E}} to ~1\tilde{{\mathscr{E}}}^{1} implies that {\mathscr{F}}^{\prime} is dense in ~1\tilde{{\mathscr{F}}}^{1} with respect to the 1{\mathscr{E}}_{1}-norm. Consequently, we have ~=~1\tilde{{\mathscr{F}}}=\tilde{{\mathscr{F}}}^{1} and ~(f,g)=(f,g)=~1(f,g)\tilde{{\mathscr{E}}}(f,g)={\mathscr{E}}(f,g)=\tilde{{\mathscr{E}}}^{1}(f,g) for all f,g~=~1f,g\in\tilde{{\mathscr{F}}}=\tilde{{\mathscr{F}}}^{1}. This completes the proof. ∎

3. Killing is domination

The relationship between Dirichlet forms and Markov processes is commonly known as contingent upon the condition of quasi-regularity. In Appendix A, we present an overview of the content relevant to this condition and the properties of the corresponding Markov processes.

Let EE be a Lusin topological space and mm a σ\sigma-finite measure on EE with support supp[m]=E\text{supp}[m]=E. We consider a quasi-regular Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(E,m)L^{2}(E,m). Let X=(Ω,,t,Xt,θt,x)X=\left(\Omega,\mathcal{F},\mathcal{F}_{t},X_{t},\theta_{t},\mathbb{P}_{x}\right) with lifetime ζ\zeta be an mm-tight Borel right process properly associated with (,)({\mathscr{E}},{\mathscr{F}}). The cemetery is denoted by Δ\Delta. We may assume that Ω\Omega contains a distinguished point [Δ][\Delta] such that Xt([Δ])=ΔX_{t}([\Delta])=\Delta for all t0t\geq 0. The goal of this section is to demonstrate that the killing transformation on XX induces the domination of (,)({\mathscr{E}},{\mathscr{F}}).

3.1. Part process

We first investigate a specific killing transformation that results in a part process defined on a finely open set.

Let GEG\subset E be a finely open set, which is nearly Borel measurable since XX is Borel measurable, and define B:=EGB:=E\setminus G. We denote DB:=inf{t0:XtB}D_{B}:=\inf\{t\geq 0:X_{t}\in B\}, and for ωΩ\omega\in\Omega, we define

XtG(ω):={Xt(ω),0t<DB(ω)ζ(ω),Δ,tDB(ω)ζ(ω),ζG(ω):=DB(ω)ζ(ω),X^{G}_{t}(\omega):=\left\{\begin{aligned} &X_{t}(\omega),\quad&0\leq t<D_{B}(\omega)\wedge\zeta(\omega),\\ &\Delta,\quad&t\geq D_{B}(\omega)\wedge\zeta(\omega),\end{aligned}\right.\qquad\zeta^{G}(\omega):=D_{B}(\omega)\wedge\zeta(\omega),

and

θtG(ω):={θt(ω),t<ζG(ω),[Δ],tζG(ω).\theta^{G}_{t}(\omega):=\left\{\begin{aligned} &\theta_{t}(\omega),\quad&t<\zeta^{G}(\omega),\\ &[\Delta],\quad&t\geq\zeta^{G}(\omega).\end{aligned}\right.

According to [30, (12.24)],

XG:=(Ω,,t,XtG,θtG,x)X^{G}:=(\Omega,\mathcal{F},\mathcal{F}_{t},X^{G}_{t},\theta^{G}_{t},\mathbb{P}_{x}) (3.1)

is a right process on the Radon space (G,(G))(G,\mathscr{B}^{*}(G)) with lifetime ζG\zeta^{G}. This process is commonly referred to as the part process of XX on GG, and its Dirichlet form, as expressed in (3.2), is called the part Dirichlet form of (,)({\mathscr{E}},{\mathscr{F}}) on GG.

Lemma 3.1.

The right process XGX^{G} is m|Gm|_{G}-tight and is properly associated with the quasi-regular Dirichlet form given by

G:={f:f~=0,-q.e. on B},\displaystyle{\mathscr{F}}^{G}:=\{f\in{\mathscr{F}}:\tilde{f}=0,{\mathscr{E}}\text{-q.e. on }B\}, (3.2)
G(f,g):=(f,g),f,gG\displaystyle{\mathscr{E}}^{G}(f,g):={\mathscr{E}}(f,g),\quad f,g\in{\mathscr{F}}^{G}

on L2(G,m|G)L^{2}(G,m|_{G}), where f~\tilde{f} denotes the {\mathscr{E}}-quasi-continuous mm-version of ff.

Proof.

This characterization has already been established in [12, Theorem 5.10]. However, in the discussion of that theorem, tightness is established in the space (3.5), which is slightly larger than GG. Quasi-regularity is a direct consequence of tightness, as stated in [12, Theorem 3.22]. In the following, we will adopt a similar approach to that discussed in [9, Theorem 3.3.8] to prove the tightness of XGX^{G} within GG.

Define u(x):=1xeTBu(x):=1-\mathbb{P}_{x}e^{-T_{B}} for xEx\in E. Then uu is finely continuous, and hence {\mathscr{E}}-quasi-continuous by Lemma A.2 (3). It follows that {u>0}=EBr=G(BBr)\{u>0\}=E\setminus B^{r}=G\cup(B\setminus B^{r}), where BrB^{r} denotes the set of all regular points for BB. Note that N:=BBrN:=B\setminus B^{r} is semipolar, and hence also {\mathscr{E}}-polar. Consider an {\mathscr{E}}-nest {Kn}\{K_{n}\} consisting of compact sets such that n1KnEN\bigcup_{n\geq 1}K_{n}\subset E\setminus N and uC({Kn})u\in C(\left\{K_{n}\right\}). Let Fn:={u1/n}F_{n}:=\{u\geq 1/n\} and KnG:=KnFnK^{G}_{n}:=K_{n}\cap F_{n}. It is straightforward to verify that KnGK^{G}_{n} is a compact subset of GG. Define TnG:=inf{t>0:XtGGKnG}T^{G}_{n}:=\inf\{t>0:X^{G}_{t}\in G\setminus K^{G}_{n}\}. Note that TnGTGKnTGFnT^{G}_{n}\geq T_{G\setminus K_{n}}\wedge T_{G\setminus F_{n}}. For mm-a.e. xGx\in G, we have

{limnTnG<ζG}\displaystyle\left\{\lim_{n\rightarrow\infty}T^{G}_{n}<\zeta^{G}\right\} {limnTGKn<TBζ}{limnTGFn<TBζ}\displaystyle\subset\left\{\lim_{n\rightarrow\infty}T_{G\setminus K_{n}}<T_{B}\wedge\zeta\right\}\cup\left\{\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}<T_{B}\wedge\zeta\right\}
{limnTEKn<ζ}{limnTGFn<TBζ}.\displaystyle\subset\left\{\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}<\zeta\right\}\cup\left\{\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}<T_{B}\wedge\zeta\right\}.

By the mm-tightness of XX, it suffices to demonstrate that for mm-a.e. xGx\in G,

x(limnTGFn<TBζ)=0.\mathbb{P}_{x}\left(\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}<T_{B}\wedge\zeta\right)=0.

In fact, it follows from the quasi-left-continuity of XX and [12, (4.8)] that

x(limnTGFn<TBζ)=x(u(XT)=limnu(XTGFn),T<TBζ),\mathbb{P}_{x}\left(\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}<T_{B}\wedge\zeta\right)=\mathbb{P}_{x}\left(u(X_{T})=\lim_{n\rightarrow\infty}u(X_{T_{G\setminus F_{n}}}),T<T_{B}\wedge\zeta\right),

where T:=limnTGFnT:=\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}. Since XTGFn(GFn)r{u1/n}X_{T_{G\setminus F_{n}}}\in\left(G\setminus F_{n}\right)^{r}\subset\{u\leq 1/n\}, we obtain

x(limnTGFn<TBζ)=x(u(XT)=0,T<TBζ)=0.\mathbb{P}_{x}\left(\lim_{n\rightarrow\infty}T_{G\setminus F_{n}}<T_{B}\wedge\zeta\right)=\mathbb{P}_{x}\left(u(X_{T})=0,T<T_{B}\wedge\zeta\right)=0.

This completes the proof. ∎

It is important to examine the relationship between G{\mathscr{E}}^{G}-quasi-notion and {\mathscr{E}}-quasi-notion. Let {Fn:n1}\{F_{n}:n\geq 1\} be an {\mathscr{E}}-nest, and define FnG:=FnGF^{G}_{n}:=F_{n}\cap G for n1n\geq 1. Since

TGFnGG:=inf{t>0:Xt(EFn)G}TEFn,T^{G}_{G\setminus F^{G}_{n}}:=\inf\{t>0:X_{t}\in(E\setminus F_{n})\cap G\}\geq T_{E\setminus F_{n}},

it follows from Lemma A.2 (2) that {FnG:n1}\{F^{G}_{n}:n\geq 1\} is an G{\mathscr{E}}^{G}-nest. In particular, the restriction of an {\mathscr{E}}-quasi-continuous function to GG is G{\mathscr{E}}^{G}-quasi-continuous, and NGN\cap G is G{\mathscr{E}}^{G}-polar if NN is {\mathscr{E}}-polar.

When (,)({\mathscr{E}},{\mathscr{F}}) is a regular and symmetric Dirichlet form, it has been established in [9, Theorem 3.3.8 (iii)] that {\mathscr{E}}-polar sets are equivalent to G{\mathscr{E}}^{G}-polar sets through the use of 11-capacity. By employing a standard transfer method via quasi-homeomorphism, this result also holds in the quasi-regular case.

Corollary 3.2.

A set NGN\subset G is G{\mathscr{E}}^{G}-polar if and only if it is {\mathscr{E}}-polar.

Proof.

Since {\mathscr{E}}-polarity and G{\mathscr{E}}^{G}-polarity are only related to the symmetric parts of {\mathscr{E}} and G{\mathscr{E}}^{G}, we only need to consider the special case where (,)({\mathscr{E}},{\mathscr{F}}) is a symmetric Dirichlet form. It is well known that (,)({\mathscr{E}},{\mathscr{F}}) is quasi-homeomorphic to another regular Dirichlet form (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}) on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) via a quasi-homeomorphism jj; see, e.g., [9, Theorem 1.4.3]. Let {Fn:n1}\{F_{n}:n\geq 1\} and {F^n:n1}\{\widehat{F}_{n}:n\geq 1\} be the {\mathscr{E}}-nest and ^\widehat{{\mathscr{E}}}-nest associated with this quasi-homeomorphism, as described in Definition A.4. By Lemma A.2 (4), we may assume without loss of generality that GFnG\cap F_{n} is an open subset of FnF_{n} for all n1n\geq 1. Define

G^:=j(G)(n1F^n).\widehat{G}:=j(G)\cap\left(\bigcup_{n\geq 1}\widehat{F}_{n}\right).

Since j:FnF^nj:F_{n}\rightarrow\widehat{F}_{n} is a topological homeomorphism, it follows that G^F^n=j(GFn)\widehat{G}\cap\widehat{F}_{n}=j(G\cap F_{n}) is an open subset of F^n\widehat{F}_{n}. In particular, G^\widehat{G} is an ^\widehat{{\mathscr{E}}}-quasi-open set. As established in [9, Theorem 3.3.8 (iii)], the ^G^\widehat{{\mathscr{E}}}^{\widehat{G}}-polarity is equivalent to the ^\widehat{{\mathscr{E}}}-polarity. Note that ^\widehat{{\mathscr{E}}}-polarity is also equivalent to the {\mathscr{E}}-polarity, as stated in, e.g., [9, Exercise 1.4.2 (ii)]. It remains to demonstrate that the G{\mathscr{E}}^{G}-polarity is equivalent to the ^G^\widehat{{\mathscr{E}}}^{\widehat{G}}-polarity.

We aim to prove that (G,G)({\mathscr{E}}^{G},{\mathscr{F}}^{G}) is quasi-homeomorphic to (^G^,^G^)(\widehat{{\mathscr{E}}}^{\widehat{G}},\widehat{{\mathscr{F}}}^{\widehat{G}}) via jG:=j|G(n1Fn)j_{G}:=j|_{G\cap\left(\cup_{n\geq 1}F_{n}\right)}. It has been indicated that {FnG:n1}\{F_{n}\cap G:n\geq 1\} forms an G{\mathscr{E}}^{G}-nest. Similarly, {F^nG^:n1}\{\widehat{F}_{n}\cap\widehat{G}:n\geq 1\} constitutes an ^G^\widehat{{\mathscr{E}}}^{\widehat{G}}-nest. It is evident that jGj_{G} is measurable and serves as a topological homeomorphism from FnGF_{n}\cap G to F^nG^\widehat{F}_{n}\cap\widehat{G}. Note that m^|G^=mj1|G^=m|GjG1\widehat{m}|_{\widehat{G}}=m\circ j^{-1}|_{\widehat{G}}=m|_{G}\circ j_{G}^{-1}. It suffices to demonstrate that

jGL2(G^,m^|G^)=L2(G,m|G).j^{*}_{G}L^{2}(\widehat{G},\widehat{m}|_{\widehat{G}})=L^{2}(G,m|_{G}). (3.3)

(Consequently, (^G^,^G^)=jG(G,G)(\widehat{{\mathscr{E}}}^{\widehat{G}},\widehat{{\mathscr{F}}}^{\widehat{G}})=j_{G}({\mathscr{E}}^{G},{\mathscr{F}}^{G}) can be shown through the equivalence between {\mathscr{E}}-quasi-notion and ^\widehat{{\mathscr{E}}}-quasi-notion.) Consider fL2(G,m|G)f\in L^{2}(G,m|_{G}), which can be treated as a function in L2(E,m)L^{2}(E,m) through zero extension. Since jj^{*} is onto, there exists a function f^L2(E^,m^)\widehat{f}\in L^{2}(\widehat{E},\widehat{m}) such that f=f^jf=\widehat{f}\circ j. It is straightforward to show that f^=0\widehat{f}=0, m^\widehat{m}-a.e. on E^G^\widehat{E}\setminus\widehat{G}. Thus, f^L2(G^,m^)\widehat{f}\in L^{2}(\widehat{G},\widehat{m}), and f=f^jGf=\widehat{f}\circ j_{G}, which establishes (3.3). This completes the proof. ∎

We present another fact that will be utilized subsequently.

Corollary 3.3.

Let G1GG_{1}\subset G. Then, G1G_{1} is {\mathscr{E}}-quasi-open if and only if it is G{\mathscr{E}}^{G}-quasi-open.

Proof.

By Lemma A.2 (4) and Corollary 3.2, it suffices to demonstrate that G1G_{1} is finely open with respect to XX if and only if it is finely open with respect to XGX^{G}. Since GG is finely open with respect to XX, this assertion is evident from the definition of finely open sets (see, e.g., [7, II, (4.1)]). ∎

3.2. Multiplicative functionals and killing transformation

As a significant type of transformation for Markov processes (see [7, III]), the general killing transformation is defined through the use of decreasing multiplicative functionals. It is important to note that the formulation in this subsection is applicable not only to Borel right processes but also to general right processes on a Radon topological space; see, e.g., [30, Chapter VII].

Definition 3.4.

A real-valued process M=(Mt)t0M=(M_{t})_{t\geq 0} is called a multiplicative functional (MF, for short) of XX if it satisfies the following conditions:

  • (1)

    The map tMt(ω)t\mapsto M_{t}(\omega) is decreasing, right continuous and takes values in [0,1][0,1] for all ωΩ\omega\in\Omega;

  • (2)

    MM is adapted, i.e., MttM_{t}\in\mathcal{F}_{t} for every t0t\geq 0;

  • (3)

    Mt+s(ω)=Mt(ω)Ms(θtω)M_{t+s}(\omega)=M_{t}(\omega)M_{s}(\theta_{t}\omega) holds for all s,t0s,t\geq 0 and ωΩ\omega\in\Omega.

Remark 3.5.

In the conventional definition of an MF, as presented in [7, III, Definition 1.1], the third condition is expressed in an a.s. sense. That is, for t,s0t,s\geq 0 and any xEx\in E, the set {ωΩ:Mt+s(ω)Mt(ω)Ms(θtω)}\{\omega\in\Omega:M_{t+s}(\omega)\neq M_{t}(\omega)M_{s}(\theta_{t}\omega)\} is a x\mathbb{P}_{x}-null set. This type of definition is referred to as a weak MF in [30, §54]. In contrast, an MF that satisfies s,t0{ωΩ:Mt+s(ω)Mt(ω)Ms(θtω)}=\bigcup_{s,t\geq 0}\{\omega\in\Omega:M_{t+s}(\omega)\neq M_{t}(\omega)M_{s}(\theta_{t}\omega)\}=\emptyset is called a perfect MF. In fact, since a weak MF MM always admits a perfect regularization M¯\bar{M} (see [30, (55.19)]), it follows that MM is equivalent to a perfect MF M¯1[0,DEEM)\bar{M}\cdot 1_{[0,D_{E\setminus E_{M}})} in the sense that

Mt=M¯t1[0,DEEM)(t),t0,a.s.M_{t}=\bar{M}_{t}\cdot 1_{[0,D_{E\setminus E_{M}})}(t),\quad\forall t\geq 0,\;\text{a.s.}

provided that the set EME_{M}, defined below, is nearly Borel measurable.

For an MF MM, we define EM:={xE:x(M0=1)=1}E_{M}:=\{x\in E:\mathbb{P}_{x}(M_{0}=1)=1\}, the set of permanent points of MM, and SM:=inf{t>0:Mt=0}S_{M}:=\inf\{t>0:M_{t}=0\}, the lifetime of MM. There is no loss of generality in assuming that Mt=0M_{t}=0 for tζt\geq\zeta since we can always replace MM with another equivalent right continuous MF, given by Mt1[0,ζ)(t)M_{t}1_{[0,\zeta)}(t). Particularly, Mt([Δ])=0M_{t}([\Delta])=0 for all t0t\geq 0. In addition, note that if EME_{M} is nearly Borel measurable and DEEM:=inf{t0:XtEEM}<D_{E\setminus E_{M}}:=\inf\{t\geq 0:X_{t}\in E\setminus E_{M}\}<\infty, then MDEEM=0M_{D_{E\setminus E_{M}}}=0 and, in particular, SMDEEMS_{M}\leq D_{E\setminus E_{M}}; see [30, (57.2)].

Let MF(X)\text{MF}(X) denote the family of all MFs MM of XX such that EME_{M} is finely open and Mζ=0M_{\zeta}=0. Further, we define

MF+(X):={MMF(X):EM=E},MF++(X):={MMF(X)+:SM=ζ}.\text{MF}_{+}(X):=\{M\in\text{MF}(X):E_{M}=E\},\quad\text{MF}_{++}(X):=\{M\in\text{MF}(X)_{+}:S_{M}=\zeta\}.

Since a fine open set is nearly Borel measurable, every MMF(X)M\in\text{MF}(X) is a right MF in the sense of [30, (57.1)]. In this context, we can define a probability measure x\mathbb{Q}_{x} on (Ω,0)(\Omega,\mathcal{F}^{0}) for each xEMx\in E_{M} by

x(Z):=x(0,]Zktd(Mt),Zb0,\mathbb{Q}_{x}(Z):=\mathbb{P}_{x}\int_{(0,\infty]}Z\circ k_{t}d(-M_{t}),\quad Z\in\text{b}\mathcal{F}^{0}, (3.4)

where M:=0M_{\infty}:=0, (kt)t0(k_{t})_{t\geq 0} are the killing operators on Ω\Omega defined by ktω(s):=ω(s)k_{t}\omega(s):=\omega(s) if t>st>s and ktω(s)=Δk_{t}\omega(s)=\Delta if tst\leq s, and 0\mathcal{F}^{0} is the σ\sigma-algebra generated by the maps f(Xt)f(X_{t}) with t0t\geq 0 and f(E)f\in\mathscr{B}(E). According to, e.g., [30, (61.5)], the restriction of (Ω,Xt,x)(\Omega,X_{t},\mathbb{Q}_{x}) to EME_{M} is a right process on EME_{M}, referred to as the (MM-)subprocess of XX and denoted by XMX^{M} or (X,M)(X,M). A more intuitive construction for subprocesses can also be found in [7, III§3]. Clearly, the transition semigroup (Qt)t0(Q_{t})_{t\geq 0} and the resolvent (Vα)α>0(V_{\alpha})_{\alpha>0} of XMX^{M} are given by

Qtf(x)=x(f(Xt)Mt),\displaystyle Q_{t}f(x)=\mathbb{P}_{x}\left(f(X_{t})M_{t}\right),
Vαf(x)=x0eαtf(Xt)Mt𝑑t\displaystyle V_{\alpha}f(x)=\mathbb{P}_{x}\int_{0}^{\infty}e^{-\alpha t}f(X_{t})M_{t}dt

for f(EM)f\in\mathscr{B}(E_{M}) and xEMx\in E_{M}. Note that Qtf(EM)Q_{t}f\in\mathscr{B}^{*}(E_{M}), but it is not necessarily Borel measurable even if EME_{M} is Borel measurable.

Remark 3.6.

An MF MM is termed exact if for any t>0t>0 and every sequence tn0t_{n}\downarrow 0, it holds that MttnθtnMtM_{t-t_{n}}\circ\theta_{t_{n}}\rightarrow M_{t}, a.s., as nn\rightarrow\infty. If MM is exact, then EME_{M} is finely open, implying that MMF(X)M\in\text{MF}(X); see [30, (56.10)]. Furthermore, every MMF+(X)M\in\text{MF}_{+}(X) is exact; see [7, III, Corollary 4.10].

It should be noted that in [37], MF(X)\text{MF}(X) represents all exact MFs of XX, which differs from its usage in this paper. In fact, for any finely closed set BB, 1[0,DBζ)1_{[0,D_{B}\wedge\zeta)} is an MF of XX, with the set of its permanent points being the fine open set G=EBG=E\setminus B. Namely, 1[0,DBζ)MF(X)1_{[0,D_{B}\wedge\zeta)}\in\text{MF}(X) in our context. However, 1[0,DBζ)1_{[0,D_{B}\wedge\zeta)} is not exact, as limt0(t+DBθt)=TB\lim_{t\downarrow 0}(t+D_{B}\circ\theta_{t})=T_{B} (not DBD_{B}; see [30, (10.4)]).

It is also important to note that another MF 1[0,TBζ)1_{[0,T_{B}\wedge\zeta)} defined by the first hitting time TBT_{B} is typically exact; however, its set of permanent points is

EG:={xE:x(TB=0)=0}=G(BBr),E_{G}:=\left\{x\in E:\mathbb{P}_{x}(T_{B}=0)=0\right\}=G\cup(B\setminus B^{r}), (3.5)

which is not necessarily equal to GG. Killing XX by 1[0,TBζ)1_{[0,T_{B}\wedge\zeta)} can also be understood as killing the part process XEGX^{E_{G}} using 1[0,TBζ)MF+(XEG)1_{[0,T_{B}\wedge\zeta)}\in\text{MF}_{+}(X^{E_{G}}). This killing transformation yields a right process, denoted by XTBX^{T_{B}}, on EGE_{G}, as examined in [12, Theorem 5.10]. It is noteworthy that EGGE_{G}\setminus G is {\mathscr{E}}-polar, and hence m|EGm|_{E_{G}}-inessential for XTBX^{T_{B}}. In particular, the restriction of XTBX^{T_{B}} to GG is precisely the part process XGX^{G}.

We present a simple observation for future reference. Recall that XGX^{G}, as expressed in (3.1), denotes the part process of XX on GG when GG is finely open.

Lemma 3.7.

If MMF(X)M\in\text{MF}(X), then MMF+(XEM)M\in\text{MF}_{+}(X^{E_{M}}), and the subprocess (X,M)(X,M) is the same as the subprocess (XEM,M)(X^{E_{M}},M). Conversely, if GG is finely open and MMF+(XG)M\in\text{MF}_{+}(X^{G}), then M1[0,DEG)MF(X)M\cdot 1_{[0,D_{E\setminus G})}\in\text{MF}(X), whose set of permanent points is GG, and the subprocess (XG,M)(X^{G},M) is the same as (X,M1[0,DEG))(X,M\cdot 1_{[0,D_{E\setminus G})}).

Proof.

It is straightforward to verify the conditions of Definition 3.4 to obtain that if MMF(X)M\in\text{MF}(X), then M1[0,DEEM)M\cdot 1_{[0,D_{E\setminus E_{M}})} is an MF of XEMX^{E_{M}}. The set of its permanent points is clearly EME_{M}, since DEEM>0D_{E\setminus E_{M}}>0, x\mathbb{P}_{x}-a.s. for all xEMx\in E_{M}. Note that Mζ=0M_{\zeta}=0 and MDEEM=0M_{D_{E\setminus E_{M}}}=0. Thus, MζEM=MζDEEM=0M_{\zeta^{E_{M}}}=M_{\zeta\wedge D_{E\setminus E_{M}}}=0. This implies that M=M1[0,DEEM)MF+(XEM)M=M\cdot 1_{[0,D_{E\setminus E_{M}})}\in\text{MF}_{+}(X^{E_{M}}). To establish the identification between (X,M)(X,M) and (XEM,M)(X^{E_{M}},M), we observe that

k^t:=ktDEEM,t0\hat{k}_{t}:=k_{t\wedge D_{E\setminus E_{M}}},\quad t\geq 0

are the killing operators of XEMX^{E_{M}}. The probability measure ^x\hat{\mathbb{Q}}_{x} of (XEM,M)(X^{E_{M}},M) on (Ω,0)(\Omega,\mathcal{F}^{0}) for xEMx\in E_{M} is given by

^x(Z)=x(0,]Zk^td(Mt),Zb0.\hat{\mathbb{Q}}_{x}(Z)=\mathbb{P}_{x}\int_{(0,\infty]}Z\circ\hat{k}_{t}d(-M_{t}),\quad Z\in\mathrm{b}\mathcal{F}^{0}.

It follows from MDEEM=0M_{D_{E\setminus E_{M}}}=0 that

^x(Z)\displaystyle\hat{\mathbb{Q}}_{x}(Z) =x(0,DEEM]Zk^td(Mt)=x(0,DEEM]Zktd(Mt)\displaystyle=\mathbb{P}_{x}\int_{(0,D_{E\setminus E_{M}}]}Z\circ\hat{k}_{t}d(-M_{t})=\mathbb{P}_{x}\int_{(0,D_{E\setminus E_{M}}]}Z\circ k_{t}d(-M_{t})
=x(0,]Zktd(Mt)=x(Z).\displaystyle=\mathbb{P}_{x}\int_{(0,\infty]}Z\circ k_{t}d(-M_{t})=\mathbb{Q}_{x}(Z).

According to [30, (61.5)], we can obtain the desired conclusion.

The converse part can be demonstrated in a similar manner. ∎

Remark 3.8.

An (t)(\mathcal{F}_{t})-stopping time TT is called a (perfect) terminal time if the set {ωΩ:t+T(θtω)T(ω),t<T(ω)}\{\omega\in\Omega:t+T(\theta_{t}\omega)\neq T(\omega),\exists t<T(\omega)\} is empty; see [30, (12.1)]. Typical examples of terminal times include the first hitting time TBT_{B} and the first entrance time DBD_{B} for any nearly Borel set BB. Consider a terminal time TT such that 0<Tζ0<T\leq\zeta. We have 1[0,T)MF+(X)1_{[0,T)}\in\text{MF}_{+}(X), and we denote by XTX^{T} (with lifetime ζT=T\zeta^{T}=T) the subprocess of XX that is killed by 1[0,T)1_{[0,T)}; see [30, (12.23)]. By employing a similar argument, it is straightforward to verify the following facts:

  • (1)

    If MM is an MF of XX, then M1[0,T)M\cdot 1_{[0,T)} is an MF of XTX^{T} with the same set of permanent points.

  • (2)

    For MMF+(X)M\in\text{MF}_{+}(X) and T:=SMT:=S_{M}, it holds that MMF++(XT)M\in\text{MF}_{++}(X^{T}). Furthermore, the subprocess (X,M)(X,M) is the same as (XT,M)(X^{T},M).

Additionally, according to [37, Theorem 2.2], every MMF+(X)M\in\text{MF}_{+}(X) admits the decomposition

Mt=0<st(1Φ(Xs,Xs))exp{0ta(Xs)𝑑As}1[0,JΓζ)(t),M_{t}=\prod_{0<s\leq t}\left(1-\Phi(X_{s-},X_{s})\right)\exp\left\{-\int_{0}^{t}a(X_{s})dA_{s}\right\}1_{[0,J_{\Gamma}\wedge\zeta)}(t), (3.6)

where Φ(E×E),0Φ<1\Phi\in\mathscr{B}(E\times E),0\leq\Phi<1, Φ\Phi vanishes on the diagonal dd of E×EE\times E, ap(E)a\in\mathrm{p}\mathscr{B}(E), AA is a positive continuous additive functional of XX, Γ(E×E)\Gamma\in\mathscr{B}(E\times E) is disjoint from dd such that SM=JΓζS_{M}=J_{\Gamma}\wedge\zeta, where JΓ:=inf{t>0:(Xt,Xt)Γ}>0J_{\Gamma}:=\inf\{t>0:(X_{t-},X_{t})\in\Gamma\}>0, a.s. (Every exact terminal time admits a representation of the form JΓJ_{\Gamma}; see [31, Theorem 6.1].) Note that the product and the integral in (3.6) can diverge only at tJΓt\geq J_{\Gamma}. See also [31, Theorem 7.1].

3.3. Bivariate Revuz measures of MFs

For an MF MM of XX, we define its bivariate Revuz measure νM\nu_{M} (relative to mm) by

νM(F):=limt01tm0tF(Xs,Xs)d(Ms),Fp(E×E).\nu_{M}(F):=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\int_{0}^{t}F(X_{s-},X_{s})d(-M_{s}),\quad F\in\mathrm{p}\mathscr{B}(E\times E). (3.7)

The existence of νM\nu_{M} is referred to in [13]. (In fact, t0tF(Xs,Xs)d(Ms)t\mapsto\int_{0}^{t}F(X_{s-},X_{s})d(-M_{s}) is a raw additive functional of XMX^{M} in the sense of [13, Definition 2.3]. Hence, the limit in (3.7) exists in [0,][0,\infty], as demonstrated in [13, Proposition 2.5]). For f,g(E)f,g\in\mathscr{B}(E), we write f(x)g(y)f(x)g(y) as fgf\otimes g. Denote by ρM\rho_{M} and λM\lambda_{M} the marginal measures of νM\nu_{M}, i.e., for any fp(E)f\in\mathrm{p}\mathscr{B}(E),

ρM(f):=νM(1f),λM(f):=νM(f1).\rho_{M}(f):=\nu_{M}(1\otimes f),\quad\lambda_{M}(f):=\nu_{M}(f\otimes 1).

Note that any function defined on EE is extended to E{Δ}E\cup\{\Delta\} by setting f(Δ)=0f(\Delta)=0. Similarly, for F(E×E)F\in\mathscr{B}(E\times E), whenever xx or yy is Δ\Delta, we define F(x,y):=0F(x,y):=0. Particularly, the actual integration interval in (3.7) is [0,t][0,SM][0,ζ)[0,t]\cap[0,S_{M}]\cap[0,\zeta).

The significance of bivariate Revuze measures lies in the representation of the Dirichlet form associated with the subprocess XMX^{M}. Let m:=m|EMm^{*}:=m|_{E_{M}}. The process XMX^{M} is called nearly mm^{*}-symmetric if its transition semigroup (Qt)t0(Q_{t})_{t\geq 0} acts on L2(EM,m)L^{2}(E_{M},m^{*}) as a strongly continuous contraction semigroup, and its infinitesimal generator \mathscr{L} with domain 𝒟()L2(EM,m)\mathcal{D}(\mathscr{L})\subset L^{2}(E_{M},m^{*}) satisfies the sector condition. This condition states that there exists a finite constant K1K\geq 1 such that

|(f,g)m|K(f,f)m1/2(g,g)m1/2\left|(-\mathscr{L}f,g)_{m^{*}}\right|\leq K\cdot(-\mathscr{L}f,f)_{m^{*}}^{1/2}\cdot(-\mathscr{L}g,g)_{m^{*}}^{1/2}

for all f,g𝒟()f,g\in\mathcal{D}(\mathscr{L}). The following theorem characterizes the Dirichlet form of XMX^{M} for the case where MMF++(X)M\in\text{MF}_{++}(X). Its general form will be presented in Theorem 3.15.

Theorem 3.9.

Let MMF++(X)M\in\text{MF}_{++}(X). Then the subprocess XMX^{M} is nearly mm-symmetric, and the associated Dirichlet form on L2(E,m)L^{2}(E,m) is given by

M=L2(E,ρM+λM),\displaystyle{\mathscr{F}}^{M}={\mathscr{F}}\cap L^{2}(E,\rho_{M}+\lambda_{M}),
M=(f,g)+νM(f~g~),f,gM.\displaystyle{\mathscr{E}}^{M}={\mathscr{E}}(f,g)+\nu_{M}(\tilde{f}\otimes\tilde{g}),\quad f,g\in{\mathscr{F}}^{M}.
Proof.

See [37, Theorem 3.10]. ∎

We present a characterization of the bivariate Revuz measure νM\nu_{M} for MMF+(X)M\in\text{MF}_{+}(X), expressed as in (3.6). Let (N,H)(N,H) be the Lévy system of XX (see [30, §73]). Note that HH is a positive continuous additive functional of XX, whose Revuz measure with respect to mm is denoted by ρH\rho_{H}. (The Revuz measure of AA is denoted by ρA\rho_{A}.) As established in [37, Theorem 4.6], the following representation of νM\nu_{M} holds:

νM(dxdy)=(1Γ+1ΓcΦ)(x,y)ν(dxdy)+δy(dx)a(y)ρA(dy),\nu_{M}(dxdy)=\left(1_{\Gamma}+1_{\Gamma^{c}}\cdot\Phi\right)(x,y)\nu(dxdy)+\delta_{y}(dx)a(y)\rho_{A}(dy), (3.8)

where ν(dxdy):=N(x,dy)ρH(dy)\nu(dxdy):=N(x,dy)\rho_{H}(dy) is referred to as the canonical measure of XX (off the diagonal dd), and δy\delta_{y} denotes the Dirac measure at yy.

Remark 3.10.

If (,)({\mathscr{E}},{\mathscr{F}}) is symmetric, then the canonical measure ν\nu of XX coincides exactly with the jumping measure of (,)({\mathscr{E}},{\mathscr{F}}), as stated in [9, Theorem 4.3.3 and Proposition 6.4.1]. When (,)({\mathscr{E}},{\mathscr{F}}) is regular but non-symmetric, the jumping measure JJ of (,)({\mathscr{E}},{\mathscr{F}}) can be defined using the method outlined in [15, (3.2.7)]; see also [23, Theorem 2.6]. Furthermore,

(f,g)=2E×Edf(x)g(y)J(dxdy){\mathscr{E}}(f,g)=-2\int_{E\times E\setminus d}f(x)g(y)J(dxdy) (3.9)

holds for any f,gCc(E)f,g\in{\mathscr{F}}\cap C_{c}(E) with disjoint support; see [23, (2.18)]. On the other hand, following the argument in [15, Lemma 5.3.2], it can be verified that (3.9) also holds true with the canonical measure ν\nu of XX in place of JJ. Thus, ν\nu is also identical to JJ. If (,)({\mathscr{E}},{\mathscr{F}}) is only quasi-regular, a similar conclusion can be obtained by the transfer method using quasi-homeomorphism. Particularly, it holds that

E×Ed(f~(x)f~(y))2ν(dxdy)<\int_{E\times E\setminus d}\left(\tilde{f}(x)-\tilde{f}(y)\right)^{2}\nu(dxdy)<\infty (3.10)

for any ff\in{\mathscr{F}}.

A σ\sigma-finite positive measure μ\mu on EE is called a smooth measure (with respect to {\mathscr{E}}) if it charges no {\mathscr{E}}-polar sets and there exists an {\mathscr{E}}-nest {Kn:n1}\{K_{n}:n\geq 1\} consisting of compact sets such that μ(Kn)<\mu(K_{n})<\infty for all n1n\geq 1. Let σ\sigma be a positive measure on E×EE\times E, and let σl\sigma_{l} and σr\sigma_{r} denote its left and right marginal measures on EE, respectively. Then, σ\sigma is called a bivariate smooth measure (with respect to {\mathscr{E}}) if

σ¯:=12(σl+σr)\bar{\sigma}:=\frac{1}{2}\left(\sigma_{l}+\sigma_{r}\right)

is smooth and σ|E×Edν\sigma|_{E\times E\setminus d}\leq\nu, where ν\nu is the canonical measure of XX.

When MMF+(X)M\in\text{MF}_{+}(X), both ρM\rho_{M} and λM\lambda_{M} are smooth measures with respect to {\mathscr{E}}; see the proof of [38, Theorem 3.4]. (This proof demonstrates the existence of two q.e. strictly positive and {\mathscr{E}}-quasi-continuous functions ff and gg such that ρM(f),λM(g)<\rho_{M}(f),\lambda_{M}(g)<\infty. Taking an {\mathscr{E}}-nest consisting of compact sets {Kn}\{K_{n}\} such that f,gC({Kn})f,g\in C(\{K_{n}\}) and f,g>0f,g>0 on n1Kn\bigcup_{n\geq 1}K_{n}, it can be readily verified that ρM(Kn),λM(Kn)<\rho_{M}(K_{n}),\lambda_{M}(K_{n})<\infty for all n1n\geq 1.) According to (3.8), the bivariate Revuz measure νM\nu_{M} is also bivariate smooth. The converse also holds true; that is, the following statement is valid.

Proposition 3.11.

The measure σ\sigma on E×EE\times E is a bivariate smooth measure if and only if there exists MMF+(X)M\in\text{MF}_{+}(X) such that σ=νM\sigma=\nu_{M}, the bivariate Revuz measure of MM.

Proof.

See [38, Theorem 4.3]. (The proof relies on the conclusion of [38, Theorem3.5], which necessitates certain improvements in its own proof, as demonstrated in Theorem 4.2 of this article.) ∎

Remark 3.12.

It should be noted that if (,)({\mathscr{E}},{\mathscr{F}}) is local (see [20, V, Definition 1.1]), meaning that ν=0\nu=0, then every bivariate smooth/Revuz measure vanishes off the diagonal dd. In this case, killing by MMF+(X)M\in\text{MF}_{+}(X) reduces to perturbation by continuous additive functionals, as examined in, e.g., [20, IV, §4c]. In this context, we have

M=L2(E,ν¯M),M(f,g)=(f,g)+Ef~g~𝑑ν¯M,f,gM,{\mathscr{F}}^{M}={\mathscr{F}}\cap L^{2}(E,\bar{\nu}_{M}),\quad{\mathscr{E}}^{M}(f,g)={\mathscr{E}}(f,g)+\int_{E}\tilde{f}\tilde{g}d\bar{\nu}_{M},\;f,g\in{\mathscr{F}}^{M},

where ν¯M:=12(ρM+λM)\bar{\nu}_{M}:=\frac{1}{2}\left(\rho_{M}+\lambda_{M}\right).

3.4. Killing by general MF

Let us turn to consider the killing transformation induced by a general MMF(X)M\in\text{MF}(X), and we denote G:=EMG:=E_{M} for convenience.

In light of Lemma 3.7 and Remark 3.8 (2), the killing transformation induced by MMF(X)M\in\text{MF}(X) can be interpreted as first applying a killing by DEGD_{E\setminus G} to obtain the part process XGX^{G}, as discussed in Section 3.1. This is followed by a second killing on XGX^{G} at the terminal time SMS_{M} (>0>0, a.s.), and finally a killing transformation using MMF++(XG,SM)M\in\text{MF}_{++}(X^{G,S_{M}}), where XG,SMX^{G,S_{M}} represents the subprocss of XGX^{G} that is killed by SMS_{M}. Note that XGX^{G} is properly associated with the part Dirichlet form (G,G)({\mathscr{E}}^{G},{\mathscr{F}}^{G}), though it is not necessarily a Borel right process. However, the representation of the bivariate Revuz measure presented in (3.8) depends on the assumption that the process is a Borel right process. Therefore, we cannot directly translate the characterization of general MF into the characterization of MF in MF+(XG)\text{MF}_{+}(X^{G}) as described in (3.8). This issue was inadvertently overlooked in [38], and we will address it in this subsection to rectify the oversight.

Lemma 3.13.

Let MMF(X)M\in\text{MF}(X) and G=EMG=E_{M}. The marginal measures ρM\rho_{M} and λM\lambda_{M} do not charge either {\mathscr{E}}-polar sets or G{\mathscr{E}}^{G}-polar sets.

Proof.

By Corollary 3.2 and Lemma A.2 (1), it suffices to demonstrate that ρM(N)=λM(N)=0\rho_{M}(N)=\lambda_{M}(N)=0 for any mm-inessential Borel set NEN\subset E for XX. For any t>0t>0, we have

ms[0,t][0,ζ)1N(Xs)d(Ms)m(XsN,s[0,ζ))=m(TN<)=0.\mathbb{P}_{m}\int_{s\in[0,t]\cap[0,\zeta)}1_{N}(X_{s})d(-M_{s})\leq\mathbb{P}_{m}\left(X_{s}\in N,\exists s\in[0,\zeta)\right)=\mathbb{P}_{m}(T_{N}<\infty)=0.

Thus, we conclude that ρM(N)=0\rho_{M}(N)=0. Additionally, according to [16, (11.1)], NN is also left mm-polar in the sense that m(SN<)=0\mathbb{P}_{m}(S_{N}<\infty)=0, where SN:=inf{0<t<ζ:XtN}S_{N}:=\inf\{0<t<\zeta:X_{t-}\in N\}. Consequently, a similar argument establishes that λM(N)=0\lambda_{M}(N)=0. ∎

Remark 3.14.

For MMF(X)M\in\text{MF}(X), the restriction of νM\nu_{M} to G×GG\times G is clearly a bivariate smooth measure with respect to G{\mathscr{E}}^{G}. (The canonical measure of XGX^{G} is 1G×Gν1_{G\times G}\cdot\nu; see [37, I, Theorem 3.9].) According to Proposition 3.11, this measure serves as the bivariate Revuz measure of an MF in MF+(XG)\text{MF}_{+}(X^{G}). By Lemma 3.7, this MF corresponds precisely to MM itself.

We can now present the general form of Theorem 3.9. This result was originally examined in [37, Theorem 4.1]; however, its proof lacks rigor. It is noteworthy that ρM(EEM)=λM(EEM)=0\rho_{M}(E\setminus E_{M})=\lambda_{M}(E\setminus E_{M})=0, since SMDEEMS_{M}\leq D_{E\setminus E_{M}}.

Theorem 3.15.

Let MMF(X)M\in\text{MF}(X) and m:=m|EMm^{*}:=m|_{E_{M}}. Then, the subprocess XMX^{M} is a nearly mm^{*}-symmetric right process on EME_{M}, which is properly associated with the Dirichlet form (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) given by

M=EML2(EM,ρM+λM),\displaystyle{\mathscr{F}}^{M}={\mathscr{F}}^{E_{M}}\cap L^{2}(E_{M},\rho_{M}+\lambda_{M}), (3.11)
M(f,g)=(f,g)+νM(f~g~),f,gM,\displaystyle{\mathscr{E}}^{M}(f,g)={\mathscr{E}}(f,g)+\nu_{M}(\tilde{f}\otimes\tilde{g}),\quad f,g\in{\mathscr{F}}^{M},

where EM{\mathscr{F}}^{E_{M}} is defined as in (3.2) with G=EMG=E_{M}. Furthermore, XMX^{M} is mm^{*}-tight, and (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) is quasi-regular on L2(EM,m)L^{2}(E_{M},m^{*}).

Proof.

Note that XMX^{M} is a right process on the Radon space (EM,(EM))(E_{M},\mathscr{B}^{*}(E_{M})). We assert that XX is mm^{*}-tight; consequently, the quasi-regularity of its associated Dirichlet form follows from, e.g., [12, Theorem 3.22] (provided that the sector condition is further verified). In light of Lemma 3.1, we need to consider the case where MMF+(X)M\in\text{MF}_{+}(X), where EM=EE_{M}=E and m=mm^{*}=m. Let {Kn:n1}\{K_{n}:n\geq 1\} be an increasing sequence of compact sets as specified in (A.1). Clearly, for any t0t\geq 0, we have

{ωΩ:limnTEKn(ktω)<ζ(ktω)}{ωΩ:limnTEKn(ω)<ζ(ω)}.\left\{\omega\in\Omega:\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}(k_{t}\omega)<\zeta(k_{t}\omega)\right\}\subset\left\{\omega\in\Omega:\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}(\omega)<\zeta(\omega)\right\}.

Therefore, it follows from (3.4) that

m(limnTEKn<ζ)\displaystyle\mathbb{Q}_{m}\left(\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}<\zeta\right) =m(0,]1{limnTEKn<ζ}ktd(Mt)\displaystyle=\mathbb{P}_{m}\int_{(0,\infty]}1_{\left\{\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}<\zeta\right\}}\circ k_{t}d(-M_{t})
m(limnTEKn<ζ)=0.\displaystyle\leq\mathbb{P}_{m}\left(\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}<\zeta\right)=0.

Note that (3.4) can be applied to this event because it remains valid for all ZbZ\in\mathrm{b}\mathcal{F}^{*}, where \mathcal{F}^{*} is the universal completion of 0\mathcal{F}^{0} (see [30, Page 286]). This establishes the mm^{*}-tightness of XMX^{M}.

It remains to demonstrate that XMX^{M} is nearly mm^{*}-symmetric and to obtain the expression (3.11) for its associated Dirichlet form (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}).

Consider first that MMF+(X)M\in\text{MF}_{+}(X), which allows for the decomposition given in (3.6). Let XΓX^{\Gamma} be the subprocess of XX that is killed by the terminal time JΓJ_{\Gamma} (namely, killed by the MF 1[0,JΓ)(t)1_{[0,J_{\Gamma})}(t)). As demonstrated in [37, Theorem 4.3], XΓX^{\Gamma} is mm-nearly symmetric on EE, and (3.11) holds with the bivariate Revuz measure given by

νΓ(F)=limt01tm(F(XJΓ,XJΓ);JΓt),Fp(E×E).\nu_{\Gamma}(F)=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\left(F(X_{J_{\Gamma}-},X_{J_{\Gamma}});J_{\Gamma}\leq t\right),\quad F\in\mathrm{p}\mathscr{B}(E\times E). (3.12)

By applying [12, Corollary 3.23] to XΓX^{\Gamma}, we can select an mm-inessential Borel set NN for XΓX^{\Gamma} such that the restricted Borel right process X1:=XΓ|ENX^{1}:=X^{\Gamma}|_{E\setminus N} is associated with the same Dirichlet form as that of XΓX^{\Gamma}. Notably, NN is {\mathscr{E}}-polar by [38, Corollary 3.3]. Consequently, it follows from Lemma 3.13 that

ρΓ(N)=λΓ(N)=ρM(N)=λM(N)=0,\rho_{\Gamma}(N)=\lambda_{\Gamma}(N)=\rho_{M}(N)=\lambda_{M}(N)=0, (3.13)

where ρΓ\rho_{\Gamma} and λΓ\lambda_{\Gamma} represent the right and left marginal measures of νΓ\nu_{\Gamma}, respectively. According to Remark 3.8 (2), MMF++(XΓ)M\in\text{MF}_{++}(X^{\Gamma}), which clearly indicates that MMF++(X1)M\in\text{MF}_{++}(X^{1}). We then have

Qt1(f|E1)(x):=Qtf(x)=x(f(Xt)Mt)=x(f(Xt1)Mt),fp(E),xE1.Q^{1}_{t}(f|_{E_{1}})(x):=Q_{t}f(x)=\mathbb{P}_{x}\left(f(X_{t})M_{t}\right)=\mathbb{P}_{x}\left(f(X^{1}_{t})M_{t}\right),\quad f\in\mathrm{p}\mathscr{B}(E),x\in E_{1}.

By applying Theorem 3.9 to X1X^{1} and MM, we can conclude that Qt1Q^{1}_{t} (and hence QtQ_{t}) acts on L2(E,m)L^{2}(E,m) as a strongly continuous contraction semigroup whose infinitesimal generator satisfies the sector condition. The Dirichlet form of the subprocess (X1,M)(X^{1},M) is evidently the same as that of XMX^{M}. By Theorem 3.9, it is given by

M=L2(E,ρΓ+λΓ+ρMΓ+λMΓ),\displaystyle{\mathscr{F}}^{M}={\mathscr{F}}\cap L^{2}(E,\rho_{\Gamma}+\lambda_{\Gamma}+\rho_{M}^{\Gamma}+\lambda_{M}^{\Gamma}),
M(f,g)=(f,g)+νΓ(f~g~)+νMΓ(f~g~),f,g,\displaystyle{\mathscr{E}}^{M}(f,g)={\mathscr{E}}(f,g)+\nu_{\Gamma}(\tilde{f}\otimes\tilde{g})+\nu^{\Gamma}_{M}(\tilde{f}\otimes\tilde{g}),\quad f,g\in{\mathscr{F}},

where νMΓ\nu^{\Gamma}_{M} is the bivariate Revuz measure of MM with respect to X1X^{1}. To finalize (3.11), it remains to establish

νM(F)=νΓ(F)+νMΓ(F),Fp(E1×E1).\nu_{M}(F)=\nu_{\Gamma}(F)+\nu^{\Gamma}_{M}(F),\quad F\in\mathrm{p}\mathscr{B}(E_{1}\times E_{1}). (3.14)

(Any function in p(E1×E1)\mathrm{p}\mathscr{B}(E_{1}\times E_{1}) can be treated as a function in p(E×E)\mathrm{p}\mathscr{B}(E\times E) through zero extension, as we have shown (3.13).) In fact, we have

νMΓ(F)\displaystyle\nu^{\Gamma}_{M}(F) =limt01tm0tF(Xs1,Xs1)d(Ms)\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\int_{0}^{t}F(X^{1}_{s-},X^{1}_{s})d(-M_{s})
=limt01tms[0,t](0,ζJΓ)F(Xs,Xs)d(Ms).\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\int_{s\in[0,t]\cap(0,\zeta\wedge J_{\Gamma})}F(X_{s-},X_{s})d(-M_{s}).

From (3.12), we can deduce that

νMΓ(F)+νΓ(F)=limt01tms[0,t](0,ζ)F(Xs,Xs)d(Ms)=νM(F).\nu^{\Gamma}_{M}(F)+\nu_{\Gamma}(F)=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\int_{s\in[0,t]\cap(0,\zeta)}F(X_{s-},X_{s})d(-M_{s})=\nu_{M}(F).

Therefore, (3.14) is established.

Finally, we consider the case where MMF(X)M\in\text{MF}(X). Let G=EMG=E_{M}, which is finely open with respect to XX, and let B:=EGB:=E\setminus G. As in the proof of Lemma 3.13, we may assume without loss of generality that XX is a special, Borel standard process and G(E)G\in\mathscr{B}(E). Let XGX^{G} represent the part process of XX on GG associated with the part Dirichlet form (G,G)({\mathscr{E}}^{G},{\mathscr{F}}^{G}), and let NN be an mm^{*}-inessential Borel set NGN\subset G such that the restricted process X1:=XG|GNX^{1}:=X^{G}|_{G\setminus N} is a Borel standard process on GNG\setminus N. It follows from Lemma 3.13 that

ρM(N)=λM(N)=0.\rho_{M}(N)=\lambda_{M}(N)=0. (3.15)

Note that MMF+(XG)M\in\text{MF}_{+}(X^{G}) by Lemma 3.7, whose restriction to GNG\setminus N is an MF in MF+(X1)\text{MF}_{+}(X^{1}). The transition semigroup of the subprocess (X1,M)(X^{1},M) is given by

Qt1(f|GN)(x):=x(f(Xt1)Mt)=x(f(Xt)Mt)=Qtf(x)Q^{1}_{t}(f|_{G\setminus N})(x):=\mathbb{P}_{x}\left(f(X^{1}_{t})M_{t}\right)=\mathbb{P}_{x}\left(f(X_{t})M_{t}\right)=Q_{t}f(x)

for xGNx\in G\setminus N and f(G)f\in\mathscr{B}(G). Based on our examination in the previous case, we conclude that Qt1Q^{1}_{t} (and hence QtQ_{t}) acts on L2(G,m)L^{2}(G,m^{*}) as a strongly continuous contraction semigroup, whose infinitesimal generator satisfies the sector condition. The Dirichlet form of XMX^{M} is identical to that of (X1,M)(X^{1},M), which has the following form:

=GL2(GN,ρM1+λM1),\displaystyle{\mathscr{F}}={\mathscr{F}}^{G}\cap L^{2}(G\setminus N,\rho^{1}_{M}+\lambda^{1}_{M}),
(f,g)=G(f,g)+νM1(f~g~),f,g,\displaystyle{\mathscr{E}}(f,g)={\mathscr{E}}^{G}(f,g)+\nu^{1}_{M}(\tilde{f}\otimes\tilde{g}),\quad f,g\in{\mathscr{F}},

where νM1\nu^{1}_{M} is the bivariate Revuz measure of MM with respect to X1X^{1}, and ρM1\rho^{1}_{M} and λM1\lambda^{1}_{M} represent the right and left marginal measures of νM1\nu^{1}_{M}, respectively. Given (3.15), it remains to show that

νM(F)=νM1(F),Fp((GN)×(GN)),\nu_{M}(F)=\nu^{1}_{M}(F),\quad F\in\mathrm{p}\mathscr{B}((G\setminus N)\times(G\setminus N)),

where FF can be regarded as a function on E×EE\times E via zero extension. In fact, since XDBBX_{D_{B}}\in B whenever DB<D_{B}<\infty and SMDBS_{M}\leq D_{B}, we have

νM1(F)\displaystyle\nu^{1}_{M}(F) =limt01tm0tF(Xs1,Xs1)d(Ms)\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m^{*}}\int_{0}^{t}F(X^{1}_{s-},X^{1}_{s})d(-M_{s})
=limt01tms[0,t](0,DBζ)F(Xs,Xs)d(Ms)\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m^{*}}\int_{s\in[0,t]\cap(0,D_{B}\wedge\zeta)}F(X_{s-},X_{s})d(-M_{s})
=limt01tms[0,t](0,DB](0,ζ)F(Xs,Xs)d(Ms)\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\mathbb{P}_{m}\int_{s\in[0,t]\cap(0,D_{B}]\cap(0,\zeta)}F(X_{s-},X_{s})d(-M_{s})
=νM(F).\displaystyle=\nu_{M}(F).

This completes the proof. ∎

Remark 3.16.

As explained before Corollary 3.2, the G{\mathscr{E}}^{G}-quasi-notion can be inherited from the {\mathscr{E}}-quasi-notion. We will now demonstrate that the M{\mathscr{E}}^{M}-quasi-notion and the G{\mathscr{E}}^{G}-quasi-notion are equivalent. Without loss of generality, assume that MMF+(X)M\in\text{MF}_{+}(X). Let

1:=L2(E,ρM+λM),1(f,g):=(f,g)+Ef~g~d(ρM+λM).{\mathscr{F}}^{1}:={\mathscr{F}}\cap L^{2}(E,\rho_{M}+\lambda_{M}),\quad{\mathscr{E}}^{1}(f,g):={\mathscr{E}}(f,g)+\int_{E}\tilde{f}\tilde{g}d(\rho_{M}+\lambda_{M}).

As shown in [9, Lemma 5.1], (1,1)({\mathscr{E}}^{1},{\mathscr{F}}^{1}) is a Dirichlet form on L2(E,m)L^{2}(E,m). (Recall that ρM+λM\rho_{M}+\lambda_{M} charges no {\mathscr{E}}-polar sets by Lemma 3.13.) According to [9, Theorem 5.1.4], the 1{\mathscr{E}}^{1}-nest is equivalent to the {\mathscr{E}}-nest. (Although only symmetric Dirichlet forms are considered in [9], the {\mathscr{E}}-quasi-notion pertains solely to the symmetric part of {\mathscr{E}}, thereby making the relevant discussions applicable to non-symmetric Dirichlet forms as well. See also [27, Proposition 2.3] for the examination of non-symmetric case.) Note that

M=1,M(f,f)1(f,f),fM.{\mathscr{F}}^{M}={\mathscr{F}}^{1},\quad{\mathscr{E}}^{M}(f,f)\leq{\mathscr{E}}^{1}(f,f),\;\forall f\in{\mathscr{F}}^{M}.

Thus, any 1{\mathscr{E}}^{1}-nest is also an M{\mathscr{E}}^{M}-nest. Conversely, it can be easily verified that (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) is strongly subordinate to (,)({\mathscr{E}},{\mathscr{F}}). (To demonstrate the denseness of M{\mathscr{F}}^{M} in {\mathscr{F}}, it suffices to consider the {\mathscr{E}}-nest {Kn}\{K_{n}\} consisting of compact sets for which ρM(Kn)+λM(Kn)<\rho_{M}(K_{n})+\lambda_{M}(K_{n})<\infty for each n1n\geq 1, and note that n1bKnM=n1bKn\bigcup_{n\geq 1}\mathrm{b}{\mathscr{F}}^{M}_{K_{n}}=\bigcup_{n\geq 1}\mathrm{b}{\mathscr{F}}_{K_{n}} is 1{\mathscr{E}}_{1}-dense in {\mathscr{F}}.) By [38, Corollary 3.3], any M{\mathscr{E}}^{M}-nest is also an {\mathscr{E}}-nest. Therefore, the M{\mathscr{E}}^{M}-nest is equivalent to the {\mathscr{E}}-nest. In particular, M{\mathscr{E}}^{M}-quasi-continuity (resp. M{\mathscr{E}}^{M}-polar set) is equivalent to {\mathscr{E}}-quasi-continuity (resp. {\mathscr{E}}-polar set).

In the special case where (,)({\mathscr{E}},{\mathscr{F}}) is symmetric, it is evident that (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) is symmetric if and only if νM\nu_{M} is symmetric, i.e., νM(dxdy)=νM(dydx)\nu_{M}(dxdy)=\nu_{M}(dydx). For further exploration of the symmetric case, see [39].

3.5. Killing is domination

From Theorem 3.15, we can readily derive the following fact.

Corollary 3.17.

Let MMF(X)M\in\text{MF}(X) such that m(EEM)=0m(E\setminus E_{M})=0. Then, the Dirichlet form (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) associated with the subprocess XMX^{M} is dominated by (,)({\mathscr{E}},{\mathscr{F}}). Furthermore, when considered as a Dirichlet form on L2(E,m)L^{2}(E,m), (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) remains quasi-regular.

Proof.

The domination property can be established by verifying the second condition of Lemma 2.1 using the expression in (3.11) for (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}). (This is also clear from the inequality QtfPtfQ_{t}f\leq P_{t}f for all fp(E)f\in\mathrm{p}\mathscr{B}(E)). We now proceed to demonstrate the quasi-regularity of (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) on L2(E,m)L^{2}(E,m). To maintain clarity, we denote (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) on L2(E,m)L^{2}(E,m) as (¯M,¯M)(\bar{{\mathscr{E}}}^{M},\bar{{\mathscr{F}}}^{M}). It is straightforward to verify that each compact subset of EME_{M} is also compact in EE because EME_{M} is endowed with the relative topology from EE. Consequently, an M{\mathscr{E}}^{M}-nest consisting of compact sets is also an ¯M\bar{{\mathscr{E}}}^{M}-nest. Therefore, the quasi-regularity of (¯M,¯M)(\bar{{\mathscr{E}}}^{M},\bar{{\mathscr{F}}}^{M}) is a consequence of the quasi-regularity of (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}). ∎

We have previously demonstrated in Theorem 2.4 that there exists a unique “intermediate” Dirichlet form (~,~)(\tilde{\mathscr{E}},\tilde{\mathscr{F}}) such that ~\tilde{\mathscr{E}} dominates M{\mathscr{E}}^{M}, and {\mathscr{E}} is a Silverstein extension of ~\tilde{\mathscr{E}}. According to Theorem 3.15, this intermediate Dirichlet form corresponds precisely to the part process XEMX^{E_{M}}.

4. Domination is killing: Quasi-regular case

In this section, we aim to establish the converse of Corollary 3.17. Specifically, we will demonstrate that every domination of a Dirichlet form corresponds to the killing transformation induced by an MF MMF(X)M\in\text{MF}(X) with m(EEM)=0m(E\setminus E_{M})=0.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two Dirichlet forms on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. Denote by (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) the intermediate Dirichlet form obtained in Theorem 2.4.

4.1. Strong subordination

Let us first interpret the probabilistic meaning of the strong subordination of (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). It should be pointed out that the Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) does not play a role in this subsection.

The following fact is elementary, and its proof is provided in [38, Corollary 3.3].

Lemma 4.1.

Any {\mathscr{E}}^{\prime}-nest is also an ~\tilde{{\mathscr{E}}}-nest. In particular, (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) is quasi-regular provided that (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is quasi-regular.

The following result essentially establishes that strong subordination corresponds to a killing transformation induced by an MF in MF+\text{MF}_{+}. While this characterization was discussed in [38, Theorem 3.5], the proof provided therein appears incomplete. Below, we present a new proof.

Theorem 4.2.

Assume that (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is quasi-regular on L2(E,m)L^{2}(E,m). Then there exists a Borel right process X~=(X~t)t0\tilde{X}=(\tilde{X}_{t})_{t\geq 0} properly associated with (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) and an MF MMF+(X~)M\in\text{MF}_{+}(\tilde{X}) such that the subprocess X~M\tilde{X}^{M} of X~\tilde{X}, killed by MM, is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Proof.

Let dd be a metric on EE that induces the topology of EE, and let Cd(E)C_{d}(E) denote the family of all dd-uniformly continuous functions on EE. The set bCd(E)\mathrm{b}C_{d}(E) consists of all bounded functions in Cd(E)C_{d}(E). Since EE is Lusin, there exists a countable subset 𝒞bCd(E)\mathcal{C}\subset\mathrm{b}C_{d}(E) that is dense in bCd(E)\mathrm{b}C_{d}(E) with respect to the uniform norm (see [30, A.2]). Define 𝒞+:={u+:u𝒞}\mathcal{C}^{+}:=\{u^{+}:u\in\mathcal{C}\}, where u+:=u0u^{+}:=u\vee 0. As mm is σ\sigma-finite, we can select a sequence of increasing Borel sets EnE_{n} such that n1En=E\bigcup_{n\geq 1}E_{n}=E and m(En)<m(E_{n})<\infty. Let

𝒞1+:={u+1En:u𝒞,n1}.\mathcal{C}^{+}_{1}:=\{u^{+}\cdot 1_{E_{n}}:u\in\mathcal{C},n\geq 1\}.

Clearly, 𝒞1+b(E)L2(E,m)\mathcal{C}^{+}_{1}\subset\mathrm{b}\mathscr{B}(E)\cap L^{2}(E,m) is countable. Denote by +\mathbb{Q}_{+} the set of all positive rationals.

Since (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) is quasi-regular by Lemma 4.1, there exists a Borel right process X~1\tilde{X}^{1} properly associated with (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). Namely, for any f(E)L2(E,m)f\in\mathscr{B}(E)\cap L^{2}(E,m) and α>0\alpha>0, U~α1f\tilde{U}^{1}_{\alpha}f is an ~\tilde{{\mathscr{E}}}-quasi-continuous mm-version of G~αf\tilde{G}_{\alpha}f, where U~α1\tilde{U}^{1}_{\alpha} is the resolvent of X~1\tilde{X}^{1}, and G~α\tilde{G}_{\alpha} is the L2L^{2}-resolvent of (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). See [20, IV, Proposition 2.8]. Similarly, there exists a Borel right process YY such that its resolvent Vα1fV^{1}_{\alpha}f is an {\mathscr{E}}^{\prime}-quasi-continuous mm-version of GαfG^{\prime}_{\alpha}f for any fb(E)L2(E,m)f\in\mathrm{b}\mathscr{B}(E)\cap L^{2}(E,m) and α>0\alpha>0, where GαG^{\prime}_{\alpha} is the L2L^{2}-resolvent of (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Since strong subordination implies domination, it follows from Lemma 2.1 (1) that

Vα1fU~α1f,m-a.e.V^{1}_{\alpha}f\leq\tilde{U}^{1}_{\alpha}f,\quad m\text{-a.e.} (4.1)

for all positive fb(E)L2(E,m)f\in\mathrm{b}\mathscr{B}(E)\cap L^{2}(E,m) and α>0\alpha>0. Note that Vα1fV^{1}_{\alpha}f is ~\tilde{{\mathscr{E}}}-quasi-continuous by Lemma 4.1. Thus, there exists an ~\tilde{{\mathscr{E}}}-nest {Fn:n1}\{F_{n}:n\geq 1\} such that

U~α1f,Vα1fC({Fn}),f𝒞1+,α+.\tilde{U}^{1}_{\alpha}f,V^{1}_{\alpha}f\in C(\{F_{n}\}),\quad\forall f\in\mathcal{C}^{+}_{1},\alpha\in\mathbb{Q}_{+}.

In particular, it follows from (4.1) that

Vα1f(x)U~α1f(x),xn1Fn,f𝒞1+,α+.V^{1}_{\alpha}f(x)\leq\tilde{U}^{1}_{\alpha}f(x),\quad\forall x\in\bigcup_{n\geq 1}F_{n},f\in\mathcal{C}^{+}_{1},\alpha\in\mathbb{Q}_{+}. (4.2)

Let us now construct an increasing sequence of mm-inessential Borel sets {Nk:k1}\{N_{k}:k\geq 1\} for X~1\tilde{X}^{1}. We begin by selecting an mm-inessential Borel set N1N_{1} for X~1\tilde{X}^{1} that contains En1FnE\setminus\bigcup_{n\geq 1}F_{n}. Since Vα11N1=0V^{1}_{\alpha}1_{N_{1}}=0, mm-a.e., and Vα11N1V^{1}_{\alpha}1_{N_{1}} is {\mathscr{E}}^{\prime}-quasi-continuous, for all α+\alpha\in\mathbb{Q}_{+}, there exists an {\mathscr{E}}^{\prime}-polar set N1N^{\prime}_{1} such that

Vα11N1(x)=0,xEN1,α+.V^{1}_{\alpha}1_{N_{1}}(x)=0,\quad\forall x\in E\setminus N^{\prime}_{1},\alpha\in\mathbb{Q}_{+}.

Note that N1N^{\prime}_{1} is also ~\tilde{{\mathscr{E}}}-polar by Lemma 4.1. Thus, we can select an mm-inessential Borel set N2N_{2} for X~1\tilde{X}^{1} such that N1N1N2N_{1}\cup N^{\prime}_{1}\subset N_{2}. For k2k\geq 2, once Nk1N^{\prime}_{k-1} and NkN_{k} have been defined, we construct NkN^{\prime}_{k} and Nk+1N_{k+1} using the following method: Select an {\mathscr{E}}^{\prime}-polar set NkN^{\prime}_{k} that contains Nk1N^{\prime}_{k-1} such that

Vα11Nk(x)=0,xENk,α+,V^{1}_{\alpha}1_{N_{k}}(x)=0,\quad\forall x\in E\setminus N^{\prime}_{k},\alpha\in\mathbb{Q}_{+}, (4.3)

and then choose an mm-inessential Borel set Nk+1N_{k+1} for X~1\tilde{X}^{1} such that NkNkNk+1N_{k}\cup N^{\prime}_{k}\subset N_{k+1}.

Define N:=k1NkN:=\bigcup_{k\geq 1}N_{k} and E0:=ENE_{0}:=E\setminus N. It is evident that E0E_{0} is a Borel invariant set for X~1\tilde{X}^{1}. Hence, the restricted process X~1|E0\tilde{X}^{1}|_{E_{0}} remains a Borel right process. Let X~\tilde{X} be the trivial extension of X~1|E0\tilde{X}^{1}|_{E_{0}} to EE, namely, X~\tilde{X} will remain at xx indefinitely if X~0=xN\tilde{X}_{0}=x\in N. Clearly, X~\tilde{X} is also a Borel right process properly associated with (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). Denote by (U~α)α>0(\tilde{U}_{\alpha})_{\alpha>0} the resolvent of X~\tilde{X}. Note that for all fb(E)f\in\mathrm{b}\mathscr{B}(E) and α>0\alpha>0,

U~αf=U~α1f1E0+1αf1N.\tilde{U}_{\alpha}f=\tilde{U}^{1}_{\alpha}f\cdot 1_{E_{0}}+\frac{1}{\alpha}f\cdot 1_{N}.

On the other hand, it follows from (4.3) that

Vα11N(x)=0,xE0,α>0.V^{1}_{\alpha}1_{N}(x)=0,\quad\forall x\in E_{0},\alpha>0.

Particularly,

Vαf:=Vα1f1E0+1αf1N,fb(E),α>0V_{\alpha}f:=V_{\alpha}^{1}f\cdot 1_{E_{0}}+\frac{1}{\alpha}f\cdot 1_{N},\quad\forall f\in\mathrm{b}\mathscr{B}(E),\alpha>0 (4.4)

provides a well-defined resolvent on EE.

We assert that

Vαf(x)U~αf(x),xE,α>0,fbp(E).V_{\alpha}f(x)\leq\tilde{U}_{\alpha}f(x),\quad\forall x\in E,\alpha>0,f\in\mathrm{bp}\mathscr{B}^{*}(E). (4.5)

It suffices to consider (and fix) xE0x\in E_{0}. From (4.2), we know that (4.5) holds for f𝒞1+f\in\mathcal{C}^{+}_{1} and α+\alpha\in\mathbb{Q}_{+}. By applying the monotone convergence theorem (for 1En1E1_{E_{n}}\uparrow 1_{E}) and the dominated convergence theorem (for the parameters +αnα>0\mathbb{Q}_{+}\ni\alpha_{n}\rightarrow\alpha>0), we can extend (4.5) to all f𝒞+f\in\mathcal{C}^{+} and all α>0\alpha>0. Fix α>0\alpha>0. Since 𝒞\mathcal{C} is dense in bCd(E)\mathrm{b}C_{d}(E) and |u+(y)v+(y)||u(y)v(y)||u^{+}(y)-v^{+}(y)|\leq|u(y)-v(y)| for all u,vbCd(E)u,v\in\mathrm{b}C_{d}(E) and yEy\in E, it follows that (4.5) holds for all positive fbCd(E)f\in\mathrm{b}C_{d}(E). Then, by utilizing [30, Proposition A2.1], (4.5) holds for all positive and bounded fC(E)f\in C(E). Note that the Lusin topological space EE is normal. Therefore, Urysohn’s lemma (see, e.g., [14, Lemma 4.15]) indicates that (4.5) holds for f=1Kf=1_{K}, where KK is an arbitrary compact set. Since the finite measure μ():=Vα(x,)+U~α(x,)\mu(\cdot):=V_{\alpha}(x,\cdot)+\tilde{U}_{\alpha}(x,\cdot) is inner regular on EE (see [30, Theorem A2.3]), we can further obtain (4.5) with f=1Bf=1_{B} for any Borel set BB. This result can be extended to all fbp(E)f\in\mathrm{b}\mathrm{p}\mathscr{B}(E) by [14, Theorem 2.10] and the monotone convergence theorem. For any fbp(E)f\in\mathrm{bp}\mathscr{B}^{*}(E), there exist f1,f2bp(E)f_{1},f_{2}\in\mathrm{b}\mathrm{p}\mathscr{B}(E) such that f1ff2f_{1}\leq f\leq f_{2} and μ(f2f1)=0\mu(f_{2}-f_{1})=0. Therefore, we can eventually conclude (4.5) for all fbp(E)f\in\mathrm{bp}\mathscr{B}^{*}(E).

For any xE0x\in E_{0}, it is straightforward to see that

αVα1E(x)=αVα11E(x)=𝔼x0et1E(Yt/α)𝑑t1E(x),α.\alpha V_{\alpha}1_{E}(x)=\alpha V^{1}_{\alpha}1_{E}(x)=\mathbb{E}_{x}\int_{0}^{\infty}e^{-t}1_{E}(Y_{t/\alpha})dt\rightarrow 1_{E}(x),\quad\alpha\rightarrow\infty.

Evidently, αVα1E(x)=1E(x)\alpha V_{\alpha}1_{E}(x)=1_{E}(x) for all xNx\in N. Hence, it follows from [7, III, Theorem 4.9] that VαV_{\alpha} is exactly subordinate to UαU_{\alpha} in the sense of [7, III, Definition 4.8]. In view of [30, Theorem (56.14)], there exists a (unique) exact MF M=(Mt)t0M=(M_{t})_{t\geq 0} of X~\tilde{X} such that

Vαf(x)=𝔼x0eαtf(X~t)Mt𝑑t,xE,fp(E).V_{\alpha}f(x)=\mathbb{E}_{x}\int_{0}^{\infty}e^{-\alpha t}f(\tilde{X}_{t})M_{t}dt,\quad\forall x\in E,f\in\mathrm{p}\mathscr{B}^{*}(E).

It is easy to verify that EM=EE_{M}=E. (If x(M0=0)=1\mathbb{P}_{x}(M_{0}=0)=1, then Vαf(x)=0V_{\alpha}f(x)=0 for all α>0\alpha>0 and fb(E)bC(E)f\in\mathrm{b}\mathscr{B}^{*}(E)\supset\mathrm{b}C(E). We must have xE0x\in E_{0} and f(x)=limααVα1f(x)=limααVαf(x)=0f(x)=\lim_{\alpha\rightarrow\infty}\alpha V^{1}_{\alpha}f(x)=\lim_{\alpha\rightarrow\infty}\alpha V_{\alpha}f(x)=0 for all fbC(E)f\in\mathrm{b}C(E), which leads to a contradiction.) Particularly, MMF+(X~)M\in\text{MF}_{+}(\tilde{X}) and VαV_{\alpha} is the resolvent of the subprocess X~M\tilde{X}^{M}.

By utilizing (4.4), it is straightforward to verify that X~M\tilde{X}^{M} is mm-nearly symmetric and that its Dirichlet form coincides with that of YY. In other words, the Dirichlet form of X~M\tilde{X}^{M} is (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). By Theorem 3.15, we can conclude that X~M\tilde{X}^{M} is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). This completes the proof. ∎

From the aforementioned theorem and Remark 3.16, it can be concluded that strong subordination guarantees the equivalence between the quasi-notions of the two Dirichlet forms.

Corollary 4.3.

An increasing sequence of closed sets is an ~\tilde{{\mathscr{E}}}-nest if and only if it is an {\mathscr{E}}^{\prime}-nest. Particularly, NEN\subset E is an ~\tilde{{\mathscr{E}}}-polar set if and only if it is an {\mathscr{E}}^{\prime}-polar set. A function ff is ~\tilde{{\mathscr{E}}}-quasi-continuous if and only if it is {\mathscr{E}}^{\prime}-quasi-continuous.

4.2. Domination is killing

We are now in a position to state the converse of Corollary 3.2 under the assumption that (,)({\mathscr{E}},{\mathscr{F}}) is quasi-regular.

Theorem 4.4.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two quasi-regular Dirichlet forms on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}, and let (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) denote the unique Dirichlet form obtained in Theorem 2.4. Then, there exists a Borel right process XX properly associated with (,)({\mathscr{E}},{\mathscr{F}}), a finely open set GG with respect to XX satisfying m(EG)=0m(E\setminus G)=0, and an MF MMF+(XG)M\in\text{MF}_{+}(X^{G}), where XGX^{G} is the part process of XX on GG, such that

  • (1)

    XGX^{G} is properly associated with (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}).

  • (2)

    The subprocess of XGX^{G} killed by MM is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Proof.

According to Theorem 2.4, (,)({\mathscr{E}},{\mathscr{F}}) is a Silverstein extension of (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). It follows that ~\tilde{{\mathscr{F}}} is a closed order ideal in {\mathscr{F}}. Therefore, the argument presented by Stollman [34] demonstrates that there exists an {\mathscr{E}}-quasi-open set G1G_{1} such that (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) is the part Dirichlet form of (,)({\mathscr{E}},{\mathscr{F}}) on G1G_{1}; see also [12, Remark 5.13]. It is evident that m(EG1)=0m(E\setminus G_{1})=0, as (~,~)(\tilde{\mathscr{E}},\tilde{\mathscr{F}}) is a Dirichlet form on L2(E,m)L^{2}(E,m).

By Lemma A.2 (4) (and its proof), we can take a Borel right process XX properly associated with (,)({\mathscr{E}},{\mathscr{F}}) and assume, without loss of generality, that G1G_{1} is a Borel, finely open set with respect to XX. It is straightforward to verify that if NG1N\subset G_{1} is an m|G1m|_{G_{1}}-inessential Borel set for the part process XG1X^{G_{1}}, then G1NG_{1}\setminus N is also finely open with respect to XX, and the restricted right process XG1|G1NX^{G_{1}}|_{G_{1}\setminus N} is precisely the part process XG1NX^{G_{1}\setminus N} of XX on G1NG_{1}\setminus N. Since XG1X^{G_{1}} is properly associated with the quasi-regular Dirichlet form (G1,G1)({\mathscr{E}}^{G_{1}},{\mathscr{F}}^{G_{1}}), as investigated in Lemma 3.1, there exists an m|G1m|_{G_{1}}-inessential Borel set N1N_{1} for XG1X^{G_{1}} such that XG1|G1N1=XG1N1X^{G_{1}}|_{G_{1}\setminus N_{1}}=X^{G_{1}\setminus N_{1}} is a Borel right process. Define G2:=G1N1G_{2}:=G_{1}\setminus N_{1}. This set is finely open with respect to XX, and XG2X^{G_{2}} is properly associated with (G2,G2)({\mathscr{E}}^{G_{2}},{\mathscr{F}}^{G_{2}}). Clearly, XG2X^{G_{2}} is also properly associated with (G1,G1)=(~,~)({\mathscr{E}}^{G_{1}},{\mathscr{F}}^{G_{1}})=(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) since any G2{\mathscr{E}}^{G_{2}}-nest consisting of compact sets is also an G1{\mathscr{E}}^{G_{1}}-nest.

Recall that (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is strongly subordinate to (~,~)=(G2,G2)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}})=({\mathscr{E}}^{G_{2}},{\mathscr{F}}^{G_{2}}). It follows from Corollary 4.3 that the ~\tilde{{\mathscr{E}}}-polar set EG2E\setminus G_{2} is also {\mathscr{E}}^{\prime}-polar. Therefore, it is straightforward to verify that (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is quasi-regular on L2(G2,m|G2)L^{2}(G_{2},m|_{G_{2}}). By following the argument presented after (4.1) in the proof of Theorem 4.2 (considering the restrictions of Vαf,U~αfV_{\alpha}f,\tilde{U}_{\alpha}f to E0E_{0} in (4.5) instead) for (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) and (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) on L2(G2,m|G2)L^{2}(G_{2},m|_{G_{2}}), we can select an m|G2m|_{G_{2}}-inessential Borel set N2N_{2} for XG2X^{G_{2}} and an MF MMF+(XG2|G2N2)M\in\text{MF}_{+}(X^{G_{2}}|_{G_{2}\setminus N_{2}}) such that the subprocess of XG2|G2N2X^{G_{2}}|_{G_{2}\setminus N_{2}} killed by MM is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on L2(G2N2,m|G2N2)L^{2}(G_{2}\setminus N_{2},m|_{G_{2}\setminus N_{2}}).

Define G:=G2N2G:=G_{2}\setminus N_{2}. Then GG is finely open with respect to XX, m(EG)=0m(E\setminus G)=0, and XG=XG2|G2N2X^{G}=X^{G_{2}}|_{G_{2}\setminus N_{2}} is properly associated with (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). In addition, MMF+(XG)M\in\text{MF}_{+}(X^{G}), and the subprocess of XGX^{G} killed by MM is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on both L2(G,m|G)L^{2}(G,m|_{G}) and L2(E,m)L^{2}(E,m). This completes the proof. ∎

Remark 4.5.

According to this proof, the fine open set GG in the theorem can be selected as a Borel set, and the part process XGX^{G} can likewise be taken as a Borel right process.

The following corollary readily follows from Lemma 3.7.

Corollary 4.6.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be two quasi-regular Dirichlet forms on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. Then, there exists a Borel right process XX properly associated with (,)({\mathscr{E}},{\mathscr{F}}) and MMF(X)M\in\text{MF}(X) with m(EEM)=0m(E\setminus E_{M})=0 such that the subprocess XMX^{M} is properly associated with (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

5. Domination is killing: General case

Now, we remove the assumption of quasi-regularity for (,)({\mathscr{E}},{\mathscr{F}}). In this setting, a right process corresponding to (,)({\mathscr{E}},{\mathscr{F}}) on EE may not exist. However, by following the approach of Silverstein [32], we can extend the space EE to establish the conclusions of Theorem 4.4 on the expanded space. See also [9, Theorem 6.6.5]. The main result is stated as follows.

Theorem 5.1.

Let (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be a quasi-regular Dirichlet form on L2(E,m)L^{2}(E,m), and let (,)({\mathscr{E}},{\mathscr{F}}) be another Dirichlet form on L2(E,m)L^{2}(E,m), not necessarily quasi-regular, such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. Denote by (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) the unique Dirichlet form obtained in Theorem 2.4. Then, there exists a locally compact separable metric space E^\widehat{E} and a measurable map j:EE^j:E\rightarrow\widehat{E} such that, by defining m^:=mj1\widehat{m}:=m\circ j^{-1}, the operator jj^{*} is a unitary map from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E,m)L^{2}(E,m), and the following hold:

  • (i)

    The image Dirichlet form (^,^):=j(,)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}):=j({\mathscr{E}},{\mathscr{F}}) is a regular Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m});

  • (ii)

    There exists an ^\widehat{{\mathscr{E}}}-quasi-open subset G^\widehat{G} of E^\widehat{E} with m^(E^G^)=0\widehat{m}(\widehat{E}\setminus\widehat{G})=0 such that jj is a quasi-homeomorphism that maps (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) to j(~,~)=(^G^,^G^)j(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}})=(\widehat{{\mathscr{E}}}^{\widehat{G}},\widehat{{\mathscr{F}}}^{\widehat{G}}), the part Dirichlet form of (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}) on G^\widehat{G};

  • (iii)

    jj is a quasi-homeomophism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to the quasi-regular image Dirichlet form j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

Furthermore, there exists a Hunt process X^\widehat{X} on E^\widehat{E}, with G^\widehat{G} being taken as a finely open set with respect to X^\widehat{X}, and an MF M^MF+(X^G^)\widehat{M}\in\text{MF}_{+}(\widehat{X}^{\widehat{G}}), where X^G^\widehat{X}^{\widehat{G}} is the part process of X^\widehat{X} on G^\widehat{G}, such that X^\widehat{X} is properly associated with j(,)j({\mathscr{E}},{\mathscr{F}}), X^G^\widehat{X}^{\widehat{G}} is properly associated with j(~,~)j(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}), and the subprocess (X^G^,M^)(\widehat{X}^{\widehat{G}},\widehat{M}) is properly associated with j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Proof.

Note that (,)({\mathscr{E}},{\mathscr{F}}) is a Silverstein extension of (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}), and according to Lemma 4.1, (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) is quasi-regular. We can then apply an argument involving the Gelfand transformation, as in [9, Theorem 6.6.5], to the symmetric parts of (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) and (,)({\mathscr{E}},{\mathscr{F}}). This yields an ~\tilde{{\mathscr{E}}}-nest {Fn:n1}\{F_{n}:n\geq 1\} consisting of compact sets, a locally compact metrizable space E^\widehat{E}, and a Borel measurable map

j:n1FnE^j:\bigcup_{n\geq 1}F_{n}\rightarrow\widehat{E}

such that j|Fnj|_{F_{n}} is a topological homeomorphism from FnF_{n} to F^n:=j(Fn)\widehat{F}_{n}:=j(F_{n}) for each n1n\geq 1, m^=mj1\widehat{m}=m\circ j^{-1} is a fully supported Radon measure on E^\widehat{E}, and jj^{*} is unitary from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E,m)L^{2}(E,m). Furthermore, (^,^):=j(,)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}):=j({\mathscr{E}},{\mathscr{F}}) is a regular Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}), and there exists an ^\widehat{{\mathscr{E}}}-quasi-open subset G^\widehat{G} of E^\widehat{E} with m^(E^G^)=0\widehat{m}(\widehat{E}\setminus\widehat{G})=0 such that jj is a quasi-homoemorphism that maps (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) to the quasi-regular Dirichlet form j(~,~)=(^G^,^G^)j(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}})=(\widehat{{\mathscr{E}}}^{\widehat{G}},\widehat{{\mathscr{F}}}^{\widehat{G}}) on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}). In particular, there exists another ~\tilde{{\mathscr{E}}}-nest {Kn:n1}\{K_{n}:n\geq 1\} such that {j(FnKn):n1}\{j(F_{n}\cap K_{n}):n\geq 1\} is an ^G^\widehat{{\mathscr{E}}}^{\widehat{G}}-nest.

Since {\mathscr{E}}^{\prime} is strongly subordinate to ~\tilde{{\mathscr{E}}}, it follows from Corollay 4.3 that the ~\tilde{{\mathscr{E}}}-nest {Fn:=FnKn:n1}\{F^{\prime}_{n}:=F_{n}\cap K_{n}:n\geq 1\} is also an {\mathscr{E}}^{\prime}-nest. Clearly, j:Fnj(Fn)j:F^{\prime}_{n}\rightarrow j(F^{\prime}_{n}) is a topological homoemorphism for each n1n\geq 1. As jj^{*} is onto L2(E,m)L^{2}(E,m), (^,^):=j(,)(\widehat{{\mathscr{E}}}^{\prime},\widehat{{\mathscr{F}}}^{\prime}):=j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is well-defined as the image Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}). To verify that jj is a quasi-homeomorphism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to (^,^)(\widehat{{\mathscr{E}}}^{\prime},\widehat{{\mathscr{F}}}^{\prime}), it remains to show that {F^n:=j(Fn):n1}\{\widehat{F}^{\prime}_{n}:=j(F^{\prime}_{n}):n\geq 1\} forms an ^\widehat{{\mathscr{E}}}^{\prime}-nest. In fact, by the definition of the image Dirichlet form, we have

^F^n\displaystyle\widehat{{\mathscr{F}}}^{\prime}_{\widehat{F}^{\prime}_{n}} ={f^L2(E^,m^):f^j,f^=0,m^-a.e. on E^j(Fn)}\displaystyle=\{\widehat{f}\in L^{2}(\widehat{E},\widehat{m}):\widehat{f}\circ j\in{\mathscr{F}}^{\prime},\widehat{f}=0,\widehat{m}\text{-a.e. on }\widehat{E}\setminus j(F^{\prime}_{n})\} (5.1)
={f^L2(E^,m^):f^j,f^j=0,m-a.e. on EFn}\displaystyle=\{\widehat{f}\in L^{2}(\widehat{E},\widehat{m}):\widehat{f}\circ j\in{\mathscr{F}}^{\prime},\widehat{f}\circ j=0,m\text{-a.e. on }E\setminus F^{\prime}_{n}\}
={f^L2(E^,m^):f^jFn}.\displaystyle=\{\widehat{f}\in L^{2}(\widehat{E},\widehat{m}):\widehat{f}\circ j\in{\mathscr{F}}^{\prime}_{F^{\prime}_{n}}\}.

As a result, n1^F^n\bigcup_{n\geq 1}\widehat{{\mathscr{F}}}^{\prime}_{\widehat{F}^{\prime}_{n}} is ^1\widehat{{\mathscr{E}}}^{\prime}_{1}-dense in ^={f^L2(E^,m^):f^j}\widehat{{\mathscr{F}}}^{\prime}=\{\widehat{f}\in L^{2}(\widehat{E},\widehat{m}):\widehat{f}\circ j\in{\mathscr{F}}^{\prime}\}. This confirms that {F^n:n1}\{\widehat{F}^{\prime}_{n}:n\geq 1\} is indeed an ^\widehat{{\mathscr{E}}}^{\prime}-nest.

For the second part of the statements, we observe that the existence of X^\widehat{X} and the fact that G^\widehat{G} can be chosen as a finely open set with respect to X^\widehat{X} follow directly from [9, Theorem 6.6.5]. It is straightforward to verify that j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is strongly subordinate to j(~,~)j(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}) and that j(,)j({\mathscr{E}},{\mathscr{F}}) is a Silverstein extension of j(~,~)j(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). By applying Theorem 4.4 to these Dirichlet forms, we can further identify an MF MMF+(X^G^)M\in\text{MF}_{+}(\widehat{X}^{\widehat{G}}) that satisfies the desired conditions. This completes the proof. ∎

Similar to Corollary 4.6, this result directly leads to the following corollary.

Corollary 5.2.

Let (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) be a quasi-regular Dirichlet form on L2(E,m)L^{2}(E,m), and let (,)({\mathscr{E}},{\mathscr{F}}) be another (not necessarily quasi-regular) Dirichlet form on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. Then there exists a locally compact separable metric space E^\widehat{E} and a measurable map j:EE^j:E\rightarrow\widehat{E} such that, by defining m^:=mj1\widehat{m}:=m\circ j^{-1}, jj^{*} is a unitary map from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E,m)L^{2}(E,m), and the following hold:

  • (1)

    The image Dirichlet form j(,)j({\mathscr{E}},{\mathscr{F}}) is a regular Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}), and jj serves as a quasi-homeomorphism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to the quasi-regular image Dirichlet form j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

  • (2)

    There exists a Hunt process X^\widehat{X} on E^\widehat{E} and an MF M^MF(X^)\widehat{M}\in\text{MF}(\widehat{X}) with m^(E^EM^)=0\widehat{m}(\widehat{E}\setminus E_{\widehat{M}})=0 such that X^\widehat{X} is properly associated with j(,)j({\mathscr{E}},{\mathscr{F}}), and the subprocess (X^,M^)(\widehat{X},\widehat{M}) is properly associated with j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Remark 5.3.

From Remark 3.12, it is clear that if {\mathscr{E}} is local, then {\mathscr{E}}^{\prime} is also local. This observation reflects the preservation of locality under the given conditions. For further discussions and related results in the regular and symmetric setting, see [2, Theorem 4.3].

6. Sandwiched Dirichlet form for domination

We consider two Dirichlet forms (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on L2(E,m)L^{2}(E,m) such that {\mathscr{E}}^{\prime}\preceq{\mathscr{E}}. The goal of this section is to explore the properties of a third Dirichlet form(𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) on L2(E,m)L^{2}(E,m) that lies between (,)({\mathscr{E}},{\mathscr{F}}) and (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). Specifically, (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) satisfies the sandwiching property:

𝒜.{\mathscr{E}}^{\prime}\preceq{\mathscr{A}}\preceq{\mathscr{E}}.

This setup enables a deeper investigation into the structure of intermediate forms and their associated processes.

6.1. General characterization

The primary assumption for this section is the quasi-regularity of (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). However, to facilitate exposition, we will also assume the quasi-regularity of (,)({\mathscr{E}},{\mathscr{F}}) and (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})). While these additional assumptions streamline the analysis, they are not strictly necessary. Indeed, through arguments analogous to those in §5, the results can be extended to the case where (,)({\mathscr{E}},{\mathscr{F}}) and (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) are not quasi-regular. The following lemma illustrates this extension and demonstrates how quasi-regularity assumptions can be relaxed.

Lemma 6.1.

Let (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) be a Dirichlet form on L2(E,m)L^{2}(E,m) satisfying 𝒜{\mathscr{E}}^{\prime}\preceq{\mathscr{A}}\preceq{\mathscr{E}}, where (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is assumed to be quasi-regular. Then, there exists a locally compact separable metric space E^\widehat{E} and a measurable map j:EE^j:E\rightarrow\widehat{E} such that, by defining m^:=mj1\widehat{m}:=m\circ j^{-1}, the operator jj^{*} is a unitary map from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E,m)L^{2}(E,m), and the following hold:

  • (1)

    jj is a quasi-homoemorphism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to the quasi-regular image Dirichlet form j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

  • (2)

    The image Dirichlet form j(𝒜,𝒟(𝒜))j({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is quasi-regular on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

  • (3)

    The image Dirichlet form j(,)j({\mathscr{E}},{\mathscr{F}}) is regular on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

Particularly, the domination relationship

j(,)j(𝒜,𝒟(𝒜))j(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})\preceq j({\mathscr{A}},\mathcal{D}({\mathscr{A}}))\preceq j({\mathscr{E}},{\mathscr{F}})

is preserved under the transformation jj.

Proof.

We begin by applying Corollary 5.2 to the Dirichlet forms (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) and (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})). This provides us with a locally compact separable metric space E^1\widehat{E}_{1} and a measurable map j1:EE^1j_{1}:E\rightarrow\widehat{E}_{1} such that, by letting m^1:=mj11\widehat{m}_{1}:=m\circ j^{-1}_{1}, j1j^{*}_{1} is a unitary map from L2(E^1,m^1)L^{2}(\widehat{E}_{1},\widehat{m}_{1}) onto L2(E,m)L^{2}(E,m), j1j_{1} acts as a quasi-homeomorphism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to the quasi-regular Dirichlet form j1(,)j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}), and j1(𝒜,𝒟(𝒜))j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is regular on L2(E^1,m^1)L^{2}(\widehat{E}_{1},\widehat{m}_{1}). Since j1j^{*}_{1} is onto L2(E,m)L^{2}(E,m), it follows that j1(,)j_{1}({\mathscr{E}},{\mathscr{F}}) defines a Dirichlet form on L2(E^1,m^1)L^{2}(\widehat{E}_{1},\widehat{m}_{1}). Additionally, the domination relationship

j1(,)j1(𝒜,𝒟(𝒜))j1(,)j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})\preceq j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}}))\preceq j_{1}({\mathscr{E}},{\mathscr{F}})

is preserved under the transformation j1j_{1}. Next, we apply Corollary 5.2 to j1(𝒜,𝒟(𝒜))j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}})) and j1(,)j_{1}({\mathscr{E}},{\mathscr{F}}). This yields another locally compact separable metric space E^\widehat{E} and a measurable map j2:E^1E^j_{2}:\widehat{E}_{1}\rightarrow\widehat{E} such that, by letting m^:=m^1j21\widehat{m}:=\widehat{m}_{1}\circ j_{2}^{-1}, j2j^{*}_{2} is a unitary map from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E^1,m^1)L^{2}(\widehat{E}_{1},\widehat{m}_{1}), j2j_{2} is a quasi-homeomorphism that maps j1(𝒜,𝒟(𝒜))j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}})) to the quasi-regular image Dirichlet form j2j1(𝒜,𝒟(𝒜))j_{2}\circ j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}})), and the image Dirichlet form j2j1(,)j_{2}\circ j_{1}({\mathscr{E}},{\mathscr{F}}) is regular on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}). Particularly, j2j1(,)j_{2}\circ j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) gives a Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}).

Define j:=j2j1j:=j_{2}\circ j_{1}, which is clearly Borel measurable from EE to E^\widehat{E}. It is evident that m^=mj1\widehat{m}=m\circ j^{-1}, jj^{*} is a unitary map from L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) onto L2(E,m)L^{2}(E,m), and

j(𝒜,𝒟(𝒜))=j2j1(𝒜,𝒟(𝒜)),j(,)=j2j1(,).j({\mathscr{A}},\mathcal{D}({\mathscr{A}}))=j_{2}\circ j_{1}({\mathscr{A}},\mathcal{D}({\mathscr{A}})),\quad j({\mathscr{E}},{\mathscr{F}})=j_{2}\circ j_{1}({\mathscr{E}},{\mathscr{F}}).

Since j1j_{1} is a quasi-homeomorphism that maps (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to j1(,)j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}), it suffices to show that j2j_{2} is a quasi-homoemorphism that maps j1(,)j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) to j(,)=j2j1(,)j({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})=j_{2}\circ j_{1}({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). To accomplish this, consider a j1𝒜j_{1}{\mathscr{A}}-nest {F^n1E^1:n1}\{\widehat{F}^{1}_{n}\subset\widehat{E}_{1}:n\geq 1\} such that j2j_{2} is a topological homeomorphism from F^n1\widehat{F}^{1}_{n} to j2(F^n1)j_{2}(\widehat{F}^{1}_{n}). Let G^1\widehat{G}_{1} be the j1𝒜j_{1}{\mathscr{A}}-quasi-open subset of E^1\widehat{E}_{1} as specified in Theorem 5.1 for (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) and (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})). According to Remark 3.16, {F^n1G^1:n1}\{\widehat{F}^{1}_{n}\cap\widehat{G}_{1}:n\geq 1\} forms a j1j_{1}{\mathscr{E}}^{\prime}-nest. Consider another j1j_{1}{\mathscr{E}}^{\prime}-nest {K^n1:n1}\{\widehat{K}^{1}_{n}:n\geq 1\} consisting of compact sets in G^1\widehat{G}_{1}, and define K^n1:=K^n1F^n1G^1\widehat{K}^{\prime 1}_{n}:=\widehat{K}^{1}_{n}\cap\widehat{F}^{1}_{n}\cap\widehat{G}_{1}. Then {K^n1:n1}\{\widehat{K}^{\prime 1}_{n}:n\geq 1\} is also a j1j_{1}{\mathscr{E}}^{\prime}-nest, and j2:K^n1K^n:=j2(K^n1)j_{2}:\widehat{K}^{\prime 1}_{n}\rightarrow\widehat{K}_{n}:=j_{2}(\widehat{K}^{\prime 1}_{n}) is a topological homoemorphism. It can be verified that {K^n:n1}\{\widehat{K}_{n}:n\geq 1\} is a jj{\mathscr{E}}^{\prime}-nest by following the argument in (5.1). This completes the proof. ∎

From this point forward, we will further assume that both (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) and (,)({\mathscr{E}},{\mathscr{F}}) are quasi-regular on L2(E,m)L^{2}(E,m). Let XX denote a Borel right process properly associated with (,)({\mathscr{E}},{\mathscr{F}}). According to Corollary 4.6, there exists an MF MMF(X)M\in\text{MF}(X) with m(EEM)=0m(E\setminus E_{M})=0 such that

(,)=(M,M).({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})=({\mathscr{E}}^{M},{\mathscr{F}}^{M}).

Recall that νM|EM×EM\nu_{M}|_{E_{M}\times E_{M}} is a bivariate smooth measure with respect to EM{\mathscr{E}}^{E_{M}}. Specifically, ν¯M:=12(ρM+λM)|EM\bar{\nu}_{M}:=\frac{1}{2}\left(\rho_{M}+\lambda_{M}\right)|_{E_{M}} is smooth with respect to EM{\mathscr{E}}^{E_{M}}, and νM|EM×EMdν|EM×EMd\nu_{M}|_{E_{M}\times E_{M}\setminus d}\leq\nu|_{E_{M}\times E_{M}\setminus d}, where ν\nu is the canonical measure of XX.

Theorem 6.2.

Let (,)({\mathscr{E}},{\mathscr{F}}) and (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) be given as above. A quasi-regular Dirichlet form (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) on L2(E,m)L^{2}(E,m) is sandwiched between {\mathscr{E}} and M{\mathscr{E}}^{M}, i.e.,

M𝒜,{\mathscr{E}}^{M}\preceq{\mathscr{A}}\preceq{\mathscr{E}},

if and only if there exists an MF MMF(X)M^{\prime}\in\text{MF}(X) with the properties

EMEM,νM|EM×EMνM|EM×EME_{M}\subset E_{M^{\prime}},\quad\nu_{M^{\prime}}|_{E_{M}\times E_{M}}\leq\nu_{M}|_{E_{M}\times E_{M}}

such that (𝒜,𝒟(𝒜))=(M,M)({\mathscr{A}},\mathcal{D}({\mathscr{A}}))=({\mathscr{E}}^{M^{\prime}},{\mathscr{F}}^{M^{\prime}}).

Proof.

To demonstrate the sufficiency, we note that Corollary 3.17 indicate that 𝒜{\mathscr{A}}\preceq{\mathscr{E}}. For convenience, let G:=EMG:=E_{M^{\prime}} and σ:=νM\sigma:=\nu_{M^{\prime}}. Consider fMf\in{\mathscr{F}}^{M}. We will show that f𝒟(𝒜)f\in\mathcal{D}({\mathscr{A}}). Since EMGE_{M}\subset G, {\mathscr{E}}-q.e., and σ|EM×EMνM|EM×EM\sigma|_{E_{M}\times E_{M}}\leq\nu_{M}|_{E_{M}\times E_{M}}, it suffices to prove that

EMf~(x)2(σ(dx,GEM)+σ(GEM,dx))<.\int_{E_{M}}\tilde{f}(x)^{2}\left(\sigma(dx,G\setminus E_{M})+\sigma(G\setminus E_{M},dx)\right)<\infty. (6.1)

In fact, it follows from fMEMf\in{\mathscr{F}}^{M}\subset{\mathscr{F}}^{E_{M}} and (3.10) that

EMf~(x)2(ν(dx,GEM)+ν(GEM,dx))G×G(f~(x)f~(y))2ν(dxdy)<.\int_{E_{M}}\tilde{f}(x)^{2}\left(\nu(dx,G\setminus E_{M})+\nu(G\setminus E_{M},dx)\right)\leq\int_{G\times G}\left(\tilde{f}(x)-\tilde{f}(y)\right)^{2}\nu(dxdy)<\infty.

Note that σ|G×Gdν|G×Gd\sigma|_{G\times G\setminus d}\leq\nu|_{G\times G\setminus d}. Thus, we can easily derive (6.1). By utilizing EMGE_{M}\subset G, {\mathscr{E}}-q.e., and M𝒟(𝒜){\mathscr{F}}^{M}\subset\mathcal{D}({\mathscr{A}}), it is straightforward to verify that bM\mathrm{b}{\mathscr{F}}^{M} forms an algebraic ideal in b𝒟(𝒜)\mathrm{b}\mathcal{D}({\mathscr{A}}). The condition σ|EM×EMνM|EM×EM\sigma|_{E_{M}\times E_{M}}\leq\nu_{M}|_{E_{M}\times E_{M}} also implies that M(f,g)𝒜(f,g){\mathscr{E}}^{M}(f,g)\geq{\mathscr{A}}(f,g) for all non-negative f,gMf,g\in{\mathscr{F}}^{M}.

Conversely, let (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) be a quasi-regular Dirichlet form such that M𝒜{\mathscr{E}}^{M}\preceq{\mathscr{A}}\preceq{\mathscr{E}}. By applying Theorem 4.4 to 𝒜{\mathscr{A}} and {\mathscr{E}}, we obtain an MF MMF(X)M^{\prime}\in\text{MF}(X) with m(EEM)=0m(E\setminus E_{M^{\prime}})=0 such that (𝒜,𝒟(𝒜))=(M,M)({\mathscr{A}},\mathcal{D}({\mathscr{A}}))=({\mathscr{E}}^{M^{\prime}},{\mathscr{F}}^{M^{\prime}}). Let G:=EMG:=E_{M^{\prime}} and σ:=νM\sigma:=\nu_{M^{\prime}}. It follows from M=EML2(EM,ν¯M)𝒟(𝒜)G{\mathscr{F}}^{M}={\mathscr{F}}^{E_{M}}\cap L^{2}(E_{M},\bar{\nu}_{M})\subset\mathcal{D}({\mathscr{A}})\subset{\mathscr{F}}^{G} that EMG{\mathscr{F}}^{E_{M}}\subset{\mathscr{F}}^{G}. (Note that M{\mathscr{F}}^{M} is 1{\mathscr{E}}_{1}-dense in EM{\mathscr{F}}^{E_{M}}.) Thus, EMGE_{M}\subset G, {\mathscr{E}}-q.e. It remains to show

σ|EM×EMνM|EM×EM.\sigma|_{E_{M}\times E_{M}}\leq\nu_{M}|_{E_{M}\times E_{M}}. (6.2)

In fact, from Corollary 3.3, we know that EME_{M} is an 𝒜{\mathscr{A}}-quasi-open set. Consider the part Dirichlet form of (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) on EME_{M}:

𝒟(𝒜EM)=EML2(EM,σ¯|EM),\displaystyle\mathcal{D}({\mathscr{A}}^{E_{M}})={\mathscr{F}}^{E_{M}}\cap L^{2}(E_{M},\bar{\sigma}|_{E_{M}}),
𝒜EM(f,g)=(f,g)+σ|EM×EM(f~g~),f,g𝒟(𝒜EM).\displaystyle{\mathscr{A}}^{E_{M}}(f,g)={\mathscr{E}}(f,g)+\sigma|_{E_{M}\times E_{M}}(\tilde{f}\otimes\tilde{g}),\quad f,g\in\mathcal{D}({\mathscr{A}}^{E_{M}}).

Applying Theorems 2.4 and 4.4 to M{\mathscr{E}}^{M} and 𝒜{\mathscr{A}}, we find that M{\mathscr{E}}^{M} is strongly subordinate to 𝒜EM{\mathscr{A}}^{E_{M}}. Moreover, it follows from Proposition 3.11 and Theorem 4.2 that there exists a bivariate smooth measure σ\sigma^{\prime} on EM×EME_{M}\times E_{M} with respect to 𝒜EM{\mathscr{A}}^{E_{M}} such that νM|EM×EM=σ|EM×EM+σ\nu_{M}|_{E_{M}\times E_{M}}=\sigma|_{E_{M}\times E_{M}}+\sigma^{\prime}. Particularly, (6.2) is established. This completes the proof. ∎

Remark 6.3.

According to Proposition 3.11, the sandwiched Dirichlet form (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) can also be expressed as

𝒟(𝒜)=GL2(G,σ¯),\displaystyle\mathcal{D}({\mathscr{A}})={\mathscr{F}}^{G}\cap L^{2}(G,\bar{\sigma}), (6.3)
𝒜(f,g)=(f,g)+σ(f~g~),f,g𝒟(𝒜),\displaystyle{\mathscr{A}}(f,g)={\mathscr{E}}(f,g)+\sigma(\tilde{f}\otimes\tilde{g}),\quad f,g\in\mathcal{D}({\mathscr{A}}),

for some {\mathscr{E}}-quasi-open set GG such that EMGE_{M}\subset G, {\mathscr{E}}-q.e., and a bivariate smooth measure σ\sigma on G×GG\times G with respect to G{\mathscr{E}}^{G} such that σ|EM×EMνM|EM×EM\sigma|_{E_{M}\times E_{M}}\leq\nu_{M}|_{E_{M}\times E_{M}}.

As noted in Remark 3.12, killing by MFs reduces to perturbation for local Dirichlet forms. Therefore, we can state the following.

Corollary 6.4.

Assume that (,)({\mathscr{E}},{\mathscr{F}}) is local. Then, the quasi-regular Dirichlet form (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is sandwiched between {\mathscr{E}} and M{\mathscr{E}}^{M} if and only if there exists an {\mathscr{E}}-quasi-open set GG with EMGE_{M}\subset G, {\mathscr{E}}-q.e., and a smooth measure μ\mu on GG with respect to G{\mathscr{E}}^{G} with μ|EMν¯M\mu|_{E_{M}}\leq\bar{\nu}_{M} such that

𝒟(𝒜)=GL2(G,μ),𝒜(f,g)=(f,g)+Gf~g~𝑑μ,f,g𝒟(𝒜).\mathcal{D}({\mathscr{A}})={\mathscr{F}}^{G}\cap L^{2}(G,\mu),\quad{\mathscr{A}}(f,g)={\mathscr{E}}(f,g)+\int_{G}\tilde{f}\tilde{g}d\mu,\;f,g\in\mathcal{D}({\mathscr{A}}).

We have decomposed domination into a composition of strong subordination and Silverstein extension in Theorem 2.4. The following corollary focuses on the case of strong subordination, providing an explicit characterization within this setting. The complementary case of Silverstein extension will be explored in detail in the subsequent two subsections.

Corollary 6.5.

Assume that (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) is strongly subordinated to (,)({\mathscr{E}},{\mathscr{F}}), which implies that EM=EE_{M}=E. Then, the quasi-regular Dirichlet form (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is sandwiched between (,)({\mathscr{E}},{\mathscr{F}}) and (M,M)({\mathscr{E}}^{M},{\mathscr{F}}^{M}) if and only if there exists a positive measure σ\sigma on E×EE\times E with σνM\sigma\leq\nu_{M} such that

𝒟(𝒜)=L2(E,σ¯),𝒜(f,g)=(f,g)+σ(f~g~),f,g𝒟(𝒜).\mathcal{D}({\mathscr{A}})={\mathscr{F}}\cap L^{2}(E,\bar{\sigma}),\quad{\mathscr{A}}(f,g)={\mathscr{E}}(f,g)+\sigma(\tilde{f}\otimes\tilde{g}),\;f,g\in\mathcal{D}({\mathscr{A}}).
Proof.

It is sufficient to observe that a positive measure σ\sigma satisfying σνM\sigma\leq\nu_{M} is a bivariate smooth measure, as is νM\nu_{M}. ∎

6.2. Sandwiched by Silverstein extension

When (,)({\mathscr{E}},{\mathscr{F}}) is a Silverstein extension of (,)=(M,M)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})=({\mathscr{E}}^{M},{\mathscr{F}}^{M}), we have

νM|EM×EM0,(,)=(EM,EM).\nu_{M}|_{E_{M}\times E_{M}}\equiv 0,\quad({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime})=({\mathscr{E}}^{E_{M}},{\mathscr{F}}^{E_{M}}).

The result in Theorem 6.2 indicates that the sandwiched Dirichlet form can be expressed as (𝒜,𝒟(𝒜))=(M,M)({\mathscr{A}},\mathcal{D}({\mathscr{A}}))=({\mathscr{E}}^{M^{\prime}},{\mathscr{F}}^{M^{\prime}}) for some MMF(X)M^{\prime}\in\text{MF}(X) such that

EMEM,-q.e.,νM|EM×EM=0.E_{M}\subset E_{M^{\prime}},\;{\mathscr{E}}\text{-q.e.},\quad\nu_{M^{\prime}}|_{E_{M}\times E_{M}}=0.

Particularly, νM\nu_{M^{\prime}} only charges ((EMEM)×EM)(EM×(EMEM))\left((E_{M^{\prime}}\setminus E_{M})\times E_{M}\right)\cup\left(E_{M}\times(E_{M^{\prime}}\setminus E_{M})\right) (off diagonal) and {(x,x)d:xEMEM}\{(x,x)\in d:x\in E_{M^{\prime}}\setminus E_{M}\} (on diagonal).

In the example below, we enumerate all sandwiched Dirichlet forms derived from the Dirichlet form of absorbing Brownian motion and one of its Silverstein extensions. This illustrates that the bivariate smooth measure σ\sigma in the expression (6.3) does not necessarily reduce to a smooth measure as stated in Corollary 6.4.

Example 6.6.

Consider the Sobolev spaces

H1([0,1]):={fL2([0,1]):f is absolutely continuous, and fL2([0,1])}H^{1}([0,1]):=\{f\in L^{2}([0,1]):f\text{ is absolutely continuous, and }f^{\prime}\in L^{2}([0,1])\}

and H01([0,1]):={fH1([0,1]):f(0)=f(1)=0}H^{1}_{0}([0,1]):=\{f\in H^{1}([0,1]):f(0)=f(1)=0\}. Let

\displaystyle{\mathscr{F}} :=H1([0,1]),\displaystyle:=H^{1}([0,1]),
(f,g)\displaystyle{\mathscr{E}}(f,g) :=12𝐃(f,g)+(f(0)f(1))(g(0)g(1)),f,g,\displaystyle:=\frac{1}{2}\mathbf{D}(f,g)+(f(0)-f(1))(g(0)-g(1)),\quad f,g\in{\mathscr{F}},

where 𝐃(f,g):=01f(x)g(x)𝑑x\mathbf{D}(f,g):=\int_{0}^{1}f^{\prime}(x)g^{\prime}(x)dx. Additionally, define

\displaystyle{\mathscr{F}}^{\prime} :=H01([0,1]),\displaystyle:=H^{1}_{0}([0,1]),
(f,g)\displaystyle{\mathscr{E}}^{\prime}(f,g) :=12𝐃(f,g),f,g.\displaystyle:=\frac{1}{2}\mathbf{D}(f,g),\quad f,g\in{\mathscr{F}}^{\prime}.

It is evident that (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is a regular, symmetric Dirichlet form on L2((0,1))L^{2}((0,1)) properly associated with the absorbing Brownian motion on (0,1)(0,1), and (,)({\mathscr{E}},{\mathscr{F}}) is a regular, symmetric Dirichlet form on L2([0,1])L^{2}([0,1]). The canonical measure corresponding to (,)({\mathscr{E}},{\mathscr{F}}) is

ν=δ{(0,1)}+δ{(1,0)}.\nu=\delta_{\{(0,1)\}}+\delta_{\{(1,0)\}}.

Since the 1{\mathscr{E}}_{1}-norm is equivalent to the 𝐃1\mathbf{D}_{1}-norm on H1([0,1])H^{1}([0,1]), it follows that every sing-point set contained in [0,1][0,1] is not {\mathscr{E}}-polar, all {\mathscr{E}}-quasi-continuous functions are continuous, and all {\mathscr{E}}-quasi-open sets are open. Moreover, (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) is quasi-regular on L2([0,1])L^{2}([0,1]), and (,)({\mathscr{E}},{\mathscr{F}}) is a Silverstein extension of (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}).

Let (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) be a quasi-regular (not necessarily symmetric) Dirichlet form on L2([0,1])L^{2}([0,1]) that is sandwiched between (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) and (,)({\mathscr{E}},{\mathscr{F}}). In its representation (6.3), the ({\mathscr{E}}-quasi-)open set GG has four possibilities:

(0,1),[0,1),(0,1],[0,1].(0,1),\quad[0,1),\quad(0,1],\quad[0,1].

The first case, where G=(0,1)G=(0,1), is straightforward to address: We must have σ0\sigma\equiv 0 in (6.3). In other words, the sandwiched Dirichlet form is identical to (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}). The second and third cases are analogous, so it suffices to discuss the second case, where G=[0,1)G=[0,1). The bivariate smooth measure σ\sigma on [0,1)×[0,1)[0,1)\times[0,1) satisfies

σ|(0,1)×(0,1)0,σ|([0,1)×[0,1))dν|([0,1)×[0,1))d.\sigma|_{(0,1)\times(0,1)}\equiv 0,\quad\sigma|_{\left([0,1)\times[0,1)\right)\setminus d}\leq\nu|_{\left([0,1)\times[0,1)\right)\setminus d}.

From these conditions, we conclude that σ\sigma only charges {(0,0)}\{(0,0)\}. Particularly,

𝒟(𝒜)\displaystyle\mathcal{D}({\mathscr{A}}) ={fH1([0,1]):f(1)=0},\displaystyle=\{f\in H^{1}([0,1]):f(1)=0\},
𝒜(f,g)\displaystyle{\mathscr{A}}(f,g) =12𝐃(f,g)+cf(0)g(0),f,g𝒟(𝒜),\displaystyle=\frac{1}{2}\mathbf{D}(f,g)+c\cdot f(0)g(0),\quad f,g\in\mathcal{D}({\mathscr{A}}),

for some constant c:=1+σ({(0,0)})1c:=1+\sigma(\{(0,0)\})\geq 1.

It remains to consider the final case where G=[0,1]G=[0,1]. Similarly, σ\sigma only charges {(0,0),(0,1),(1,0),(1,1)}\{(0,0),(0,1),(1,0),(1,1)\}, with β0:=σ({(0,1)})1\beta_{0}:=\sigma(\{(0,1)\})\leq 1 and β1:=σ({(1,0)})1\beta_{1}:=\sigma(\{(1,0)\})\leq 1. Specifically, we have

𝒟(𝒜)\displaystyle\mathcal{D}({\mathscr{A}}) =H1([0,1]),\displaystyle=H^{1}([0,1]),
𝒜(f,g)\displaystyle{\mathscr{A}}(f,g) =(f,g)+α0f(0)g(0)+α1f(1)g(1)+β0f(0)g(1)+β1f(1)g(0),\displaystyle={\mathscr{E}}(f,g)+\alpha_{0}f(0)g(0)+\alpha_{1}f(1)g(1)+\beta_{0}f(0)g(1)+\beta_{1}f(1)g(0),

for f,gH1([0,1])f,g\in H^{1}([0,1]), where α0:=σ({(0,0)})0\alpha_{0}:=\sigma(\{(0,0)\})\geq 0 and α1:=σ({(1,1)})0\alpha_{1}:=\sigma(\{(1,1)\})\geq 0. Note that (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is not necessarily symmetric. It is symmetric if and only if β0=β1\beta_{0}=\beta_{1}(=:β=:\beta), in which case (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) admits the Beurling-Deny decomposition

𝒜(f,g)\displaystyle{\mathscr{A}}(f,g) =12𝐃(f,g)+(1β)(f(0)f(1))(g(0)g(1))\displaystyle=\frac{1}{2}\mathbf{D}(f,g)+(1-\beta)\cdot(f(0)-f(1))(g(0)-g(1))
+(α0+β)f(0)g(0)+(α1+β)f(1)g(1),f,gH1([0,1]),\displaystyle\qquad+(\alpha_{0}+\beta)f(0)g(0)+(\alpha_{1}+\beta)f(1)g(1),\quad f,g\in H^{1}([0,1]),

where 0β10\leq\beta\leq 1.

6.3. Sandwiched by reflected Dirichlet space

We close this section by considering a typical case examined in the existing literatures such as [5, 18, 29]. Assume further that both (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) and (,)({\mathscr{E}},{\mathscr{F}}) are symmetric, and (,)({\mathscr{E}},{\mathscr{F}}) is the active reflected Dirichlet space of (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) in the sense of [9, Definition 6.4.4]. In order to maintain consistency with the notation employed in [9], we will (only in this subsection) denote (,)({\mathscr{E}}^{\prime},{\mathscr{F}}^{\prime}) as (,)({\mathscr{E}},{\mathscr{F}}) and refer to (,)({\mathscr{E}},{\mathscr{F}}) as (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}). It is important to note that (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}) is a Silverstein extension of (,)({\mathscr{E}},{\mathscr{F}}); see [9, Theorem 6.6.3].

Let MM be the MF corresponding to the domination ref{\mathscr{E}}\preceq{\mathscr{E}}^{\mathrm{ref}}. Then, (,)({\mathscr{E}},{\mathscr{F}}) is the part Dirichlet form of (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}) on EME_{M}. The set EEME\setminus E_{M} is precisely the “boundary” (of EME_{M}), as considered in [18, Definition 4.2]. The following result demonstrates that the Dirichlet form sandwiched between (,)({\mathscr{E}},{\mathscr{F}}) and its active reflected Dirichlet space can be expressed as a perturbation of the part Dirichlet form of (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}) on some ref{\mathscr{E}}^{\mathrm{ref}}-quasi-open set GG (EM\supset E_{M}, ref{\mathscr{E}}^{\mathrm{ref}}-q.e.), thereby recovering the main result, Theorem 4.6, of [18].

Corollary 6.7.

The quasi-regular (not necessarily symmetric) Dirichlet form (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is sandwiched between (,)({\mathscr{E}},{\mathscr{F}}) and (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}) if and only if there exists an ref{\mathscr{E}}^{\mathrm{ref}}-quasi-open set GG with EMGE_{M}\subset G, ref{\mathscr{E}}^{\mathrm{ref}}-q.e., and a smooth measure μ\mu with respect to ref,G{\mathscr{E}}^{\mathrm{ref},G} with μ(GEM)=0\mu(G\setminus E_{M})=0 such that

𝒟(𝒜)=aref,GL2(G,μ),\displaystyle\mathcal{D}({\mathscr{A}})={\mathscr{F}}^{\mathrm{ref},G}_{a}\cap L^{2}(G,\mu),
𝒜(f,g)=ref(f,g)+Gf~g~𝑑μ,f,g𝒟(𝒜),\displaystyle{\mathscr{A}}(f,g)={\mathscr{E}}^{\mathrm{ref}}(f,g)+\int_{G}\tilde{f}\tilde{g}d\mu,\quad f,g\in\mathcal{D}({\mathscr{A}}),

where (ref,G,aref,G)({\mathscr{E}}^{\mathrm{ref},G},{\mathscr{F}}^{\mathrm{ref},G}_{a}) is the part Dirichlet form of (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}) on GG. Particularly, (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) is symmetric.

Proof.

It suffices to prove the necessity. For convenience, we take every function to be its ref{\mathscr{E}}^{\mathrm{ref}}-quasi-continuous mm-version. According to, e.g., [9, Proposition 6.4.1], (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\text{ref}}_{a}) admits the Beurling-Deny decomposition as follows: For fareff\in{\mathscr{F}}^{\mathrm{ref}}_{a},

ref(f,f)=12μ~fc(E)+12E×Ed(f(x)f(y))2J~(dxdy)+Ef(x)2κ~(dx),{\mathscr{E}}^{\mathrm{ref}}(f,f)=\frac{1}{2}\tilde{\mu}^{c}_{\langle f\rangle}(E)+\frac{1}{2}\int_{E\times E\setminus d}(f(x)-f(y))^{2}\tilde{J}(dxdy)+\int_{E}f(x)^{2}\tilde{\kappa}(dx),

where μ~fc\tilde{\mu}^{c}_{\langle f\rangle} denotes the energy measure defined in [9, (4.3.8)]. For ff\in{\mathscr{F}}, we have

(f,f)\displaystyle{\mathscr{E}}(f,f) =ref(f,f)\displaystyle={\mathscr{E}}^{\mathrm{ref}}(f,f)
=12μ~fc(E)+12EM×EMd(f(x)f(y))2J(dxdy)+EMf(x)2κ(dx),\displaystyle=\frac{1}{2}\tilde{\mu}^{c}_{\langle f\rangle}(E)+\frac{1}{2}\int_{E_{M}\times E_{M}\setminus d}(f(x)-f(y))^{2}J(dxdy)+\int_{E_{M}}f(x)^{2}\kappa(dx),

where

J:=J~|EM×EMd,κ:=(κ~+J~(dx,EEM))|EM.J:=\tilde{J}|_{E_{M}\times E_{M}\setminus d},\quad\kappa:=\left(\tilde{\kappa}+\tilde{J}(dx,E\setminus E_{M})\right)\big{|}_{E_{M}}.

Thus, JJ is the jumping measure of (,)({\mathscr{E}},{\mathscr{F}}) and κ\kappa serves as its killing measure. Utilizing [9, (4.3.34)], we further conclude that

μfc=μ~fc|EM,f,\mu^{c}_{\langle f\rangle}=\tilde{\mu}^{c}_{\langle f\rangle}\big{|}_{E_{M}},\quad f\in{\mathscr{F}}, (6.4)

where μfc\mu_{\langle f\rangle}^{c} denotes the energy measure of ff\in{\mathscr{F}} for the Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}). Let loc\overset{\circ}{{\mathscr{F}}}_{\mathrm{loc}} denote the local Dirichlet space of {\mathscr{F}}, as defined in [9, (4.3.31)]. The measure μfc\mu^{c}_{\langle f\rangle} is well-defined for each flocf\in\overset{\circ}{{\mathscr{F}}}_{\mathrm{loc}} using [9, Theorem 4.3.10]. According to the definition of the reflected Dirichlet space (see [9, Definition 6.4.4]), we have

aref={flocL2(E,m):^(f,f)<},{\mathscr{F}}^{\mathrm{ref}}_{a}=\{f\in\overset{\circ}{{\mathscr{F}}}_{\mathrm{loc}}\cap L^{2}(E,m):\widehat{{\mathscr{E}}}(f,f)<\infty\},

where for flocf\in\overset{\circ}{{\mathscr{F}}}_{\mathrm{loc}},

^(f,f):=12μfc(EM)+12EM×EMd(f(x)f(y))2J(dxdy)+EMf(x)2κ(dx),\widehat{{\mathscr{E}}}(f,f):=\frac{1}{2}\mu^{c}_{\langle f\rangle}(E_{M})+\frac{1}{2}\int_{E_{M}\times E_{M}\setminus d}(f(x)-f(y))^{2}J(dxdy)+\int_{E_{M}}f(x)^{2}\kappa(dx),

and ref(f,f)=^(f,f){\mathscr{E}}^{\mathrm{ref}}(f,f)=\widehat{{\mathscr{E}}}(f,f) for fareff\in{\mathscr{F}}^{\mathrm{ref}}_{a}. We will prove that

^(c)(f,g)\displaystyle\widehat{{\mathscr{E}}}^{(c)}(f,g) :=12μf,gc(EM)\displaystyle:=\frac{1}{2}\mu^{c}_{\langle f,g\rangle}(E_{M}) (6.5)
=18(μf+gc(EM)μfgc(EM)),f,garefloc\displaystyle=\frac{1}{8}\left(\mu^{c}_{\langle f+g\rangle}(E_{M})-\mu^{c}_{\langle f-g\rangle}(E_{M})\right),\quad f,g\in{\mathscr{F}}^{\mathrm{ref}}_{a}\subset\overset{\circ}{{\mathscr{F}}}_{\mathrm{loc}}

is strongly local in the sense of [9, Proposition 6.4.1 (i)]. Once this is established, the uniqueness of Beurling-Deny decomposition for (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\text{ref}}_{a}) implies that

J~(EM,EEM)=J~(EEM,EM)=J~((EEM)×(EEM)d)=0.\tilde{J}(E_{M},E\setminus E_{M})=\tilde{J}(E\setminus E_{M},E_{M})=\tilde{J}\left((E\setminus E_{M})\times(E\setminus E_{M})\setminus d\right)=0.

Note that J~\tilde{J} is the canonical measure of (ref,aref)({\mathscr{E}}^{\mathrm{ref}},{\mathscr{F}}^{\mathrm{ref}}_{a}). Consequently, the bivariate smooth measure σ\sigma in the expression (6.3) for (𝒜,𝒟(𝒜))({\mathscr{A}},\mathcal{D}({\mathscr{A}})) reduces to a smooth measure μ\mu such that μ(GEM)=0\mu(G\setminus E_{M})=0, since σ|EM×EM=0\sigma|_{E_{M}\times E_{M}}=0.

It remains to demonstrate that (6.5) is strongly local. Let FF denote the ref{\mathscr{E}}^{\mathrm{ref}}-quasi-support of ff, and assume that gg is constant in an ref{\mathscr{E}}^{\mathrm{ref}}-quasi-open set BB with FBF\subset B. Consider a sequence of increasing {\mathscr{E}}-quasi-open sets {Gn:n1}\{G_{n}:n\geq 1\} such that n1Gn=EM\bigcup_{n\geq 1}G_{n}=E_{M}, {\mathscr{E}}-q.e., along with two sequences of functions {fn:n1}\{f_{n}:n\geq 1\} and {gn:n1}\{g_{n}:n\geq 1\} in {\mathscr{F}} such that fn=ff_{n}=f and gn=gg_{n}=g, {\mathscr{E}}-q.e. on GnG_{n} for each n1n\geq 1. We aim to prove that

μfn,gnc(Gn)=0,n1.\mu^{c}_{\langle f_{n},g_{n}\rangle}(G_{n})=0,\quad n\geq 1. (6.6)

Corollary 3.3 indicates that GnFG_{n}\setminus F is ref{\mathscr{E}}^{\mathrm{ref}}-quasi-open. Given that μfnc|Gn=μfc|Gn\mu^{c}_{\langle f_{n}\rangle}|_{G_{n}}=\mu^{c}_{\langle f\rangle}|_{G_{n}}, it follows from (6.4) and [9, Proposition 4.3.1 (ii)] that

μfnc(GnF)=μ~fc(GnF)=0.\mu^{c}_{\langle f_{n}\rangle}(G_{n}\setminus F)=\tilde{\mu}^{c}_{\langle f\rangle}(G_{n}\setminus F)=0.

Thus, we have

|μfn,gnc(GnF)|2μfnc(GnF)μgnc(GnF)=0.\left|\mu^{c}_{\langle f_{n},g_{n}\rangle}(G_{n}\setminus F)\right|^{2}\leq\mu^{c}_{\langle f_{n}\rangle}(G_{n}\setminus F)\mu^{c}_{\langle g_{n}\rangle}(G_{n}\setminus F)=0.

Since gg is constant on GnBG_{n}\cap B (GnF\supset G_{n}\cap F), we can similarly obtain that μgnc(GnB)=μ~gc(GnB)=0\mu^{c}_{\langle g_{n}\rangle}(G_{n}\cap B)=\tilde{\mu}^{c}_{\langle g\rangle}(G_{n}\cap B)=0, which further implies μfn,gnc(GnF)=0\mu^{c}_{\langle f_{n},g_{n}\rangle}(G_{n}\cap F)=0. Therefore, (6.6) is established. This completes the proof. ∎

7. Application to the Laplacian with Robin boundary conditions

Let Ωn\Omega\subset{\mathbb{R}}^{n} (with n1n\geq 1) be a non-empty open set with the boundary Γ:=Ω¯Ω\Gamma:=\overline{\Omega}\setminus\Omega. Consider the Sobolev space

H1(Ω):={uL2(Ω):DjuL2(Ω),j=1,2,,n}H^{1}(\Omega):=\{u\in L^{2}(\Omega):D_{j}u\in L^{2}(\Omega),j=1,2,\cdots,n\}

with the norm

uH1(Ω)2:=uL2(Ω)2+j=1nDjuL2(Ω)2,\|u\|^{2}_{H^{1}(\Omega)}:=\|u\|^{2}_{L^{2}(\Omega)}+\sum_{j=1}^{n}\|D_{j}u\|^{2}_{L^{2}(\Omega)},

where Dju=uxjD_{j}u=\frac{\partial u}{\partial x_{j}} represents the distributional derivative. For u,vH1(Ω)u,v\in H^{1}(\Omega), define

𝐃(u,v):=j=1nΩDju(x)Djv(x)𝑑x.\mathbf{D}(u,v):=\sum_{j=1}^{n}\int_{\Omega}D_{j}u(x)D_{j}v(x)dx.

Moreover, we let

H~1(Ω):=H1(Ω)C(Ω¯)¯H1(Ω),\tilde{H}^{1}(\Omega):=\overline{H^{1}(\Omega)\cap C(\overline{\Omega})}^{H^{1}(\Omega)},

where C(Ω¯)C(\overline{\Omega}) denotes the space of all continuous real-valued functions on Ω¯\overline{\Omega}, and

H01(Ω):=Cc(Ω)¯H1(Ω),H^{1}_{0}(\Omega):=\overline{C_{c}^{\infty}(\Omega)}^{H^{1}(\Omega)},

where Cc(Ω)C_{c}^{\infty}(\Omega) denotes the space of all infinitely differential functions on Ω\Omega with compact support.

It is well known that (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) is a regular, symmetric Dirichlet form on L2(Ω¯)L^{2}(\overline{\Omega}), while (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) is a regular, symmetric Dirichlet form on L2(Ω)L^{2}(\Omega); see, e.g., [3, §2]. Denote by X¯\overline{X} and X0X^{0} the Hunt processes on Ω¯\overline{\Omega} and Ω{\Omega} corresponding to these forms, respectively. For convenience, let Cap be the 11-capacity of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)), and if the form symbol preceding the quasi-notion is omitted, it defaults to the quasi-notion corresponding to (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)). For instance, u~H~1(Ω)\tilde{u}\in\tilde{H}^{1}(\Omega) denotes the quasi-continuous version of uu with respect to (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)), and the statement u~=0\tilde{u}=0 q.e. on Γ\Gamma asserts that u~\tilde{u} is identically equal to 0 on Γ\Gamma outside some polar set with respect to (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)).

7.1. Local Robin boundary

In this subsection, we aim to investigate the (not necessarily symmetric) Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(Ω)L^{2}(\Omega) (=L2(Ω¯)=L^{2}(\overline{\Omega})) that is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)), i.e.,

(𝐃,H01(Ω))(,)(𝐃,H~1(Ω)).(\mathbf{D},H^{1}_{0}(\Omega))\preceq({\mathscr{E}},{\mathscr{F}})\preceq(\mathbf{D},\tilde{H}^{1}(\Omega)). (7.1)

To avoid trivial cases, we always assume that H01(Ω)H~1(Ω)H^{1}_{0}(\Omega)\neq\tilde{H}^{1}(\Omega), meaning that Γ\Gamma is not polar (see, e.g., [3, Proposition 2.5]).

Definition 7.1.

A positive Borel measure μ\mu on ΓμΓ\Gamma_{\mu}\subset\Gamma is called quasi-admissible if it satisfies the following conditions:

  • (1)

    Bμ:=ΓΓμB_{\mu}:=\Gamma\setminus\Gamma_{\mu} is quasi-closed.

  • (2)

    μ\mu, regarded as a measure on Ω¯Bμ\overline{\Omega}\setminus B_{\mu} with μ(Ω):=0\mu(\Omega):=0, is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on Gμ:=Ω¯BμG_{\mu}:=\overline{\Omega}\setminus B_{\mu}.

Remark 7.2.

It is important to emphasize that the defining set Γμ\Gamma_{\mu} of μ\mu plays a crucial role in determining admissibility. For two admissible measures, even if their trivial extensions on Ω¯\overline{\Omega} are identical, the Dirichlet forms defined by these measures in (7.2) will differ if their defining sets are not q.e. equal.

For any quasi-admissible measure μ\mu on Γμ\Gamma_{\mu}, we define a quadratic form on L2(Ω¯)L^{2}(\overline{\Omega}) as follows:

μ:={u~H~1(Ω):u~=0 q.e. on Bμ,Γμu~(x)2μ(dx)<},\displaystyle{\mathscr{F}}^{\mu}:=\left\{\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }B_{\mu},\int_{\Gamma_{\mu}}\tilde{u}(x)^{2}\mu(dx)<\infty\right\}, (7.2)
μ(u~,v~):=𝐃(u~,v~)+Γμu~(x)v~(x)μ(dx),u~,v~μ.\displaystyle{\mathscr{E}}^{\mu}(\tilde{u},\tilde{v}):=\mathbf{D}(\tilde{u},\tilde{v})+\int_{\Gamma_{\mu}}\tilde{u}(x)\tilde{v}(x)\mu(dx),\quad\tilde{u},\tilde{v}\in{\mathscr{F}}^{\mu}.

These forms encapsulate all Dirichlet forms that are sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)), as demonstrated in the following result. It should be emphasized that the quasi-regularity of (,)({\mathscr{E}},{\mathscr{F}}) is not essential to this characterization, as we explained in Lemma 6.1.

Theorem 7.3.

A quasi-regular (not necessarily symmetric) Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(Ω¯)L^{2}(\overline{\Omega}) is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) if and only if there exists a quasi-admissible measure μ\mu such that (,)=(μ,μ)({\mathscr{E}},{\mathscr{F}})=({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}).

Proof.

For a quasi-admissible measure μ\mu, (μ,μ)({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}) is a quasi-regular Dirichlet form on L2(Gμ)L^{2}(G_{\mu}) according to, e.g., [9, Theorems 3.3.8 and 5.1.5]. It is also quasi-regular on L2(Ω¯)L^{2}(\overline{\Omega}) for the same reason explained in the proof of Corollary 3.17. Conversely, the opposing statement is a consequence of Corollary 6.4. ∎

If (,)({\mathscr{E}},{\mathscr{F}}) is merely a coercive closed form that satisfies (7.1), then the Ouhabaz’s domination criterion (the first statement of Lemma 2.1) indicates that it must satisfy the Markovian property, meaning that (,)({\mathscr{E}},{\mathscr{F}}) is automatically a Dirichlet form. Particularly, we have the following.

Corollary 7.4.

Let (,)({\mathscr{E}},{\mathscr{F}}) be a coercive closed form on L2(Ω)L^{2}(\Omega) that satisfies (7.1). Then (,)({\mathscr{E}},{\mathscr{F}}) is a local and symmetric Dirichlet form.

Proof.

It is sufficient to note that the transformation jj in Lemma 6.1 keeps the local property of Dirichlet forms. ∎

Remark 7.5.

In [4], the authors presented Example 4.5 to demonstrate that the locality condition in [4, Theorem 4.1] is indispensable. However, upon closer inspection, the example fails to substantiate this claim because the associated L2L^{2}-semigroup (Tt)t0(T_{t})_{t\geq 0} in this example is not positive; therefore, the Ouhabaz’s domination criterion is not applicable in this context. In fact, the locality condition in [4, Theorem 4.1] can indeed be omitted, as we will explain in Example 7.6 that the admissible measures in [4] are actually quasi-admissible. Readers are also referred to [2, §3] for this correction.

Let us examine the admissible measure, which has been investigated in various works, including [2, 3, 4]. It will be demonstrated that all admissible measures are, in fact, quasi-admissible.

Example 7.6.

For a positive Borel measure μ\mu on Γ\Gamma, let

Γμ:={zΓ:r>0 such that μ(ΓB(z,r))<},\Gamma_{\mu}:=\{z\in\Gamma:\exists r>0\text{ such that }\mu(\Gamma\cap B(z,r))<\infty\},

where B(z,r):={zn:|z|<r}B(z,r):=\{z\in{\mathbb{R}}^{n}:|z|<r\}. In other words, Γμ\Gamma_{\mu} is the part of Γ\Gamma on which μ\mu is locally finite. The measure μ\mu is called admissible if Cap(A)=0\text{Cap}(A)=0 implies μ(A)=0\mu(A)=0 for any Borel set AΓμA\subset\Gamma_{\mu}. The main result, Theorem 4.1, of [4] states that if μ\mu is admissible, then

𝒟(aμ):={u~H~1(Ω):u~=0 q.e. on ΓΓμ,Γμu~(x)2μ(dx)<},\displaystyle\mathcal{D}(a_{\mu}):=\left\{\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }\Gamma\setminus\Gamma_{\mu},\int_{\Gamma_{\mu}}\tilde{u}(x)^{2}\mu(dx)<\infty\right\}, (7.3)
aμ(u~,v~):=𝐃(u~,v~)+Γμu~(x)v~(x)μ(dx),u~,v~𝒟(aμ)\displaystyle a_{\mu}(\tilde{u},\tilde{v}):=\mathbf{D}(\tilde{u},\tilde{v})+\int_{\Gamma_{\mu}}\tilde{u}(x)\tilde{v}(x)\mu(dx),\quad\tilde{u},\tilde{v}\in\mathcal{D}(a_{\mu})

is a Dirichlet form that satisfies (7.1).

In fact, Y:=ΩΓμY:=\Omega\cup\Gamma_{\mu} is clearly an open subset of Ω¯\overline{\Omega}, and μ\mu is a Radon measure on YY. If μ\mu is admissible, then μ\mu is a Radon smooth measure with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on YY. According to [9, Theorems 3.3.9 and 5.1.6], (aμ,𝒟(aμ))(a_{\mu},\mathcal{D}(a_{\mu})) is a regular Dirichlet form on L2(Y)L^{2}(Y). Particularly, Bμ:=ΓΓμ=Ω¯YB_{\mu}:=\Gamma\setminus\Gamma_{\mu}=\overline{\Omega}\setminus Y is closed and, consequently, quasi-closed. Therefore, we conclude that μ\mu, when restricted to Γμ\Gamma_{\mu}, is quasi-admissible.

Remark 7.7.

Consider the case n2n\geq 2, where each single-point subset of Γ\Gamma is a polar set. If μ\mu is a smooth measure with supp[μ]=Γ\text{supp}[\mu]=\Gamma, it is always possible to construct another smooth measure μ\mu^{\prime}, which is equivalent to μ\mu, such that μ\mu^{\prime} is nowhere Radon on Γ\Gamma in the sense that μ(U)=\mu^{\prime}(U)=\infty for any non-empty open subset UU of Γ\Gamma (see [20, IV, Theorem 4.7]). Note that μ\mu^{\prime} with Γμ=Γ\Gamma_{\mu^{\prime}}=\Gamma is quasi-admissible, and (μ,μ)({\mathscr{E}}^{\mu^{\prime}},{\mathscr{F}}^{\mu^{\prime}}), defined as (7.2), is a quasi-regular Dirichlet form on L2(Ω¯)L^{2}(\overline{\Omega}), which differs from (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)). However, the admissibility argument in this example leads to (aμ,𝒟(aμ))=(𝐃,H01(Ω))(a_{\mu^{\prime}},\mathcal{D}(a_{\mu^{\prime}}))=(\mathbf{D},H^{1}_{0}(\Omega)) because the part of Γ\Gamma, on which μ\mu^{\prime} is locally finite, is empty.

We close this subsection by introducing a method for deriving a quasi-admissible measure from a positive Borel measure on Γ\Gamma that does not charge any polar sets.

Let μ\mu be a positive measure on (Γ,(Γ))(\Gamma,\mathscr{B}(\Gamma)), which can be extended to Ω¯\overline{\Omega} in the way that μ(Ω):=0\mu(\Omega):=0. Assume that μ\mu charges no polar sets, i.e., Cap(A)=0\text{Cap}(A)=0 implies μ(A)=0\mu(A)=0 for any subset AΓA\subset\Gamma. Under these conditions, the integral Γu~2𝑑μ\int_{\Gamma}\tilde{u}^{2}d\mu (\leq\infty) is well defined for any quasi-continuous function u~H~1(Ω)\tilde{u}\in\tilde{H}^{1}(\Omega). Consider a sequence {un}H~1(Ω)L2(Γ,μ)\{u_{n}\}\subset\tilde{H}^{1}(\Omega)\cap L^{2}(\Gamma,\mu), which is dense in H~1(Ω)L2(Γ,μ)\tilde{H}^{1}(\Omega)\cap L^{2}(\Gamma,\mu) with respect to the norm H1(Ω)\|\cdot\|_{H^{1}(\Omega)}. We define the set

Bμ:=n=1{u~n=0}.B_{\mu}:=\bigcap_{n=1}^{\infty}\{\tilde{u}_{n}=0\}.

Note that BμΓB_{\mu}\subset\Gamma, q.e., and BμB_{\mu} is independent of the choice of the sequence {un}\{u_{n}\}, as can be demonstrated by the following fact.

Lemma 7.8.

Let μ\mu be a positive measure on (Γ,(Γ))(\Gamma,\mathscr{B}(\Gamma)) such that Cap(A)=0\text{Cap}(A)=0 implies μ(A)=0\mu(A)=0 for any subset AΓA\subset\Gamma. Then BμB_{\mu} is a quasi-closed set such that u~=0\tilde{u}=0, q.e. on BμB_{\mu} for any uH~1(Ω)L2(Γ,μ)u\in\tilde{H}^{1}(\Omega)\cap L^{2}(\Gamma,\mu). Furthermore, the restriction of μ\mu to Γμ:=ΓBμ\Gamma_{\mu}:=\Gamma\setminus B_{\mu} is quasi-admissible.

Proof.

For any uH~1(Ω)L2(Γ,μ)u\in\tilde{H}^{1}(\Omega)\cap L^{2}(\Gamma,\mu), there exists a subsequence {unk}\{u_{n_{k}}\} such that u~nku~\tilde{u}_{n_{k}}\rightarrow\tilde{u}, q.e. Thus, u~=0\tilde{u}=0, q.e. on BμB_{\mu}. For the second part, it suffices to show that μ\mu, as a measure on Ω¯Bμ\overline{\Omega}\setminus B_{\mu}, is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on Ω¯Bμ\overline{\Omega}\setminus B_{\mu}. This fact has been established in [27, Proposition 2.13]. ∎

7.2. Non-local Robin boundary

Now we turn to examine the Laplacian with non-local Robin boundary condition, which has been discussed in [25] and the references cited therein.

Definition 7.9.

Let κ\kappa be a positive Borel measure on Γκ,θ(Γ)\Gamma_{\kappa,\theta}(\subset\Gamma), and let θ\theta be a positive, symmetric measure on (Γκ,θ×Γκ,θ)d(\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta})\setminus d, where dd denotes the diagonal of Γ×Γ\Gamma\times\Gamma. The pair (κ,θ)(\kappa,\theta) is called quasi-admissible if the following conditions hold:

  • (1)

    Bκ,θ:=ΓΓκ,θB_{\kappa,\theta}:=\Gamma\setminus\Gamma_{\kappa,\theta} is quasi-closed.

  • (2)

    κ+θ¯\kappa+\bar{\theta} is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on Ω¯Bκ,θ\overline{\Omega}\setminus B_{\kappa,\theta}, where θ¯(dx):=θ(dx,Γκ,θ)\bar{\theta}(dx):=\theta(dx,\Gamma_{\kappa,\theta}).

Remark 7.10.

The measure κ\kappa with Γκ:=Γκ,θ\Gamma_{\kappa}:=\Gamma_{\kappa,\theta} is quasi-admissible in the sense of Definition 7.1. If θ=0\theta=0, then (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) as presented in (7.4) is identical to (κ,κ)({\mathscr{E}}^{\kappa},{\mathscr{F}}^{\kappa}) defined in (7.2) with μ=κ\mu=\kappa.

For any quasi-admissible pair (κ,θ)(\kappa,\theta), we define the symmetric quadratic form on L2(Ω¯)L^{2}(\overline{\Omega}) as follows:

κ,θ:={u~H~1(Ω):u~=0 q.e. on Bκ,θ,\displaystyle{\mathscr{F}}^{\kappa,\theta}:=\bigg{\{}\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }B_{\kappa,\theta}, (7.4)
Γκ,θ×Γκ,θd(u~(x)u~(y))2θ(dxdy)+Γκ,θu~(x)2κ(dx)<},\displaystyle\qquad\qquad\quad\int_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta(dxdy)+\int_{\Gamma_{\kappa,\theta}}\tilde{u}(x)^{2}\kappa(dx)<\infty\bigg{\}},
κ,θ(u~,v~):=𝐃(u~,v~)+Γκ,θ×Γκ,θd(u~(x)u~(y))(v~(x)v~(y))θ(dxdy)\displaystyle{\mathscr{E}}^{\kappa,\theta}(\tilde{u},\tilde{v}):=\mathbf{D}(\tilde{u},\tilde{v})+\int_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d}(\tilde{u}(x)-\tilde{u}(y))(\tilde{v}(x)-\tilde{v}(y))\theta(dxdy)
+Γκ,θu~(x)v~(x)κ(dx),u~,v~κ,θ.\displaystyle\qquad\qquad\quad+\int_{\Gamma_{\kappa,\theta}}\tilde{u}(x)\tilde{v}(x)\kappa(dx),\quad\tilde{u},\tilde{v}\in{\mathscr{F}}^{\kappa,\theta}.

The following fact is elementary.

Lemma 7.11.

Let (κ,θ)(\kappa,\theta) be a quasi-admissible pair. Then (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) is a quasi-regular and symmetric Dirichlet form on L2(Ω¯)L^{2}(\overline{\Omega}).

Proof.

Since H01(Ω)κ,θH^{1}_{0}(\Omega)\subset{\mathscr{F}}^{\kappa,\theta}, it follows that κ,θ{\mathscr{F}}^{\kappa,\theta} is dense in L2(Ω¯)L^{2}(\overline{\Omega}). Clearly, (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) is a symmetric, positive form that satisfies the Markovian property. To show that it is closed, consider an 1κ,θ{\mathscr{E}}^{\kappa,\theta}_{1}-Cauchy sequence {u~n}κ,θ\{\tilde{u}_{n}\}\subset{\mathscr{F}}^{\kappa,\theta}. This sequence is also Cauchy in both H~1(Ω)\tilde{H}^{1}(\Omega) and L2(Γκ,θ,κ)L^{2}(\Gamma_{\kappa,\theta},\kappa), and moreover, fn(x,y):=u~n(x)u~n(y)f_{n}(x,y):=\tilde{u}_{n}(x)-\tilde{u}_{n}(y) is Cauchy in L2(Γκ,θ×Γκ,θd,θ)L^{2}(\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d,\theta). Hence, there exist u~H~1(Ω)\tilde{u}\in\tilde{H}^{1}(\Omega), vL2(Γκ,θ,κ)v\in L^{2}(\Gamma_{\kappa,\theta},\kappa), and fL2(Γκ,θ×Γκ,θd,θ)f\in L^{2}(\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d,\theta) such that u~n\tilde{u}_{n} converges to u~\tilde{u} and vv with respect to the H1(Ω)H^{1}(\Omega) and L2(Γκ,θ,κ)L^{2}(\Gamma_{\kappa,\theta},\kappa)-norms, respectively, and fnff_{n}\rightarrow f with respect to the L2(Γκ,θ×Γκ,θd,θ)L^{2}(\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d,\theta)-norm. By taking a subsequence if necessary, we can assume that u~nu~\tilde{u}_{n}\rightarrow\tilde{u}, q.e., (which indicates that u~=\tilde{u}= q.e. on Bκ,θB_{\kappa,\theta},) u~nv\tilde{u}_{n}\rightarrow v, κ\kappa-a.e., and fnff_{n}\rightarrow f, θ\theta-a.e. Since κ+θ¯\kappa+\bar{\theta} charges no polar sets, we obtain that u~=v\tilde{u}=v, κ\kappa-a.e., and f(x,y)=u~(x)u~(y)f(x,y)=\tilde{u}(x)-\tilde{u}(y) for θ\theta-a.e. (x,y)(x,y). Particularly, u~κ,θ\tilde{u}\in{\mathscr{F}}^{\kappa,\theta} and 1κ,θ(u~nu~,u~nu~)0{\mathscr{E}}^{\kappa,\theta}_{1}(\tilde{u}_{n}-\tilde{u},\tilde{u}_{n}-\tilde{u})\rightarrow 0. This establishes the closedness of (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}).

It remains to demonstrate that (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) is quasi-regular on L2(Ω¯)L^{2}(\overline{\Omega}). Let μ:=κ+4θ¯\mu:=\kappa+4\bar{\theta}, which is quasi-admissible in the sense of Definition 7.1 with Γμ=Γκ,θ\Gamma_{\mu}=\Gamma_{\kappa,\theta}, and define (μ,μ)({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}) as in (7.2). Notably, we have

μκ,θ,κ,θ(u~,u~)μ(u~,u~),u~μ.{\mathscr{F}}^{\mu}\subset{\mathscr{F}}^{\kappa,\theta},\quad{\mathscr{E}}^{\kappa,\theta}(\tilde{u},\tilde{u})\leq{\mathscr{E}}^{\mu}(\tilde{u},\tilde{u}),\;\forall\tilde{u}\in{\mathscr{F}}^{\mu}.

This implies that any μ{\mathscr{E}}^{\mu}-nest is also an κ,θ{\mathscr{E}}^{\kappa,\theta}-nest. Similarly, consider the part Dirichlet form (𝒜,𝒢)({\mathscr{A}},{\mathscr{G}}) of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on G:=Ω¯Bκ,θG:=\overline{\Omega}\setminus B_{\kappa,\theta}. From

κ,θ𝒢,𝒜(u~,u~)κ,θ(u~,u~),u~κ,θ,{\mathscr{F}}^{\kappa,\theta}\subset{\mathscr{G}},\quad{\mathscr{A}}(\tilde{u},\tilde{u})\leq{\mathscr{E}}^{\kappa,\theta}(\tilde{u},\tilde{u}),\;\forall\tilde{u}\in{\mathscr{F}}^{\kappa,\theta},

it follows that any κ,θ{\mathscr{E}}^{\kappa,\theta}-nest is also an 𝒜{\mathscr{A}}-nest. According to [9, Theorem 5.1.4], any 𝒜{\mathscr{A}}-nest is an μ{\mathscr{E}}^{\mu}-nest. Thus, the quasi-notions for κ,θ,μ{\mathscr{E}}^{\kappa,\theta},{\mathscr{E}}^{\mu} and 𝒜{\mathscr{A}} are equivalent. Therefore, the quasi-regularity of (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) follows directly from the quasi-regularity of (𝒜,𝒢)({\mathscr{A}},{\mathscr{G}}) and (μ,μ)({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}). This completes the proof. ∎

Analogously, the admissible pair considered in [25] is quasi-admissible, as explained in the following example.

Example 7.12.

Consider a positive Borel measure κ\kappa on Γ\Gamma and a symmetric, positive Borel measure θ\theta on Γ×Γd\Gamma\times\Gamma\setminus d. Let Γκ,θ\Gamma_{\kappa,\theta} be the part of Γ\Gamma on which κ\kappa and θ^\hat{\theta} are locally finite, i.e.,

Γκ,θ:={zΓ:r>0 such that κ(ΓB(z,r))+θ^(ΓB(z,r))<},\Gamma_{\kappa,\theta}:=\{z\in\Gamma:\exists r>0\text{ such that }\kappa(\Gamma\cap B(z,r))+\hat{\theta}(\Gamma\cap B(z,r))<\infty\}, (7.5)

where θ^(dx):=θ(dx,Γ)\hat{\theta}(dx):=\theta(dx,\Gamma). The pair (κ,θ)(\kappa,\theta) is called admissible if κ(A)+θ^(A)=0\kappa(A)+\hat{\theta}(A)=0 for any polar set AΓκ,θA\subset\Gamma_{\kappa,\theta}. For such a pair, the key findings from [25, Theorems 3.2 and 3.3] indicate that

𝒟(𝒜κ,θ):={uH1(Ω)Cc(Ω¯):u=0 on ΓΓκ,θ,\displaystyle\mathcal{D}({\mathscr{A}}^{\kappa,\theta}):=\bigg{\{}u\in H^{1}(\Omega)\cap C_{c}(\overline{\Omega}):u=0\text{ on }\Gamma\setminus\Gamma_{\kappa,\theta},
Γ×Γd(u(x)u(y))2θ(dxdy)+Γu(x)2κ(dx)<},\displaystyle\qquad\qquad\quad\int_{\Gamma\times\Gamma\setminus d}({u}(x)-u(y))^{2}\theta(dxdy)+\int_{\Gamma}u(x)^{2}\kappa(dx)<\infty\bigg{\}},
𝒜κ,θ(u,v):=𝐃(u,v)+Γ×Γd(u(x)u(y))(v(x)v(y))θ(dxdy)\displaystyle{\mathscr{A}}^{\kappa,\theta}(u,v):=\mathbf{D}(u,v)+\int_{\Gamma\times\Gamma\setminus d}({u}(x)-u(y))({v}(x)-v(y))\theta(dxdy)
+Γu(x)v(x)κ(dx),u,v𝒟(𝒜κ,θ)\displaystyle\qquad\qquad\quad+\int_{\Gamma}u(x)v(x)\kappa(dx),\quad u,v\in\mathcal{D}({\mathscr{A}}^{\kappa,\theta})

is closable, and its closure is given by

𝒟(𝒜κ,θ)¯={u~H~1(Ω):u~=0 q.e. on ΓΓκ,θ,\displaystyle\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}=\bigg{\{}\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }\Gamma\setminus\Gamma_{\kappa,\theta},
Γ×Γd(u~(x)u~(y))2θ(dxdy)+Γu~(x)2κ(dx)<}.\displaystyle\qquad\qquad\quad\int_{\Gamma\times\Gamma\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta(dxdy)+\int_{\Gamma}\tilde{u}(x)^{2}\kappa(dx)<\infty\bigg{\}}.

In fact, Y:=ΩΓκ,θY:=\Omega\cup\Gamma_{\kappa,\theta} is clearly an open subset of Ω¯\overline{\Omega}, and κ+θ^\kappa+\hat{\theta} is a Radon measure on YY. Thus, κ+θ^\kappa+\hat{\theta} is a Radon smooth measure on YY with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on YY. It is noteworthy that 𝒟(𝒜κ,θ)C0(Y)\mathcal{D}({\mathscr{A}}^{\kappa,\theta})\subset C_{0}(Y), where C0(Y)C_{0}(Y) represents the family of all continuous functions on YY that vanish at infinity (see [14, page 132]), and 𝒟(𝒜κ,θ)\mathcal{D}({\mathscr{A}}^{\kappa,\theta}) is dense in C0(Y)C_{0}(Y) with respect to the uniform norm by the Stone-Weierstrass theorem. Particularly, (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}) is a regular, symmetric Dirichlet form on L2(Y)L^{2}(Y), which admits the Beurling-Deny decomposition for u~,v~𝒟(𝒜κ,θ)¯\tilde{u},\tilde{v}\in\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})} as follows:

𝒜κ,θ(u~,v~)\displaystyle{\mathscr{A}}^{\kappa,\theta}(\tilde{u},\tilde{v}) =𝐃(u~,v~)+Γκ,θ×Γκ,θd(u~(x)u~(y))(v~(x)v~(y))θ(dxdy)\displaystyle=\mathbf{D}(\tilde{u},\tilde{v})+\int_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d}(\tilde{u}(x)-\tilde{u}(y))(\tilde{v}(x)-\tilde{v}(y))\theta(dxdy)
+Γκ,θu~(x)v~(x)(κ+2θ^1)(dx),\displaystyle\qquad\qquad\quad+\int_{\Gamma_{\kappa,\theta}}\tilde{u}(x)\tilde{v}(x)\left(\kappa+2\hat{\theta}_{1}\right)(dx),

where θ^1(dx):=θ(dx,ΓΓκ,θ)\hat{\theta}_{1}(dx):={\theta}(dx,\Gamma\setminus\Gamma_{\kappa,\theta}).

It is straightforward to verify that κ:=(κ+2θ^1)|Γκ,θ\kappa^{\prime}:=(\kappa+2\hat{\theta}_{1})|_{\Gamma_{\kappa,\theta}} and θ=θ|Γκ,θ×Γκ,θd\theta^{\prime}=\theta|_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d} comprise a quasi-admissible pair with Γκ,θ=Γκ,θ\Gamma_{\kappa^{\prime},\theta^{\prime}}=\Gamma_{\kappa,\theta}, and (κ,θ,κ,θ)({\mathscr{E}}^{\kappa^{\prime},\theta^{\prime}},{\mathscr{F}}^{\kappa^{\prime},\theta^{\prime}}), as defined in (7.4), is identical to (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}).

Clearly, (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) is dominated by (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}), while (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) is not dominated by (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) due to Corollary 7.4. The following result characterizes all symmetric Dirichlet forms that are sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}).

Theorem 7.13.

Let (κ,θ)(\kappa,\theta) be a quasi-admissible pair. A quasi-regular and symmetric Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(Ω¯)L^{2}(\overline{\Omega}) is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) if and only if there exists another quasi-admissible pair (κ,θ)(\kappa^{\prime},\theta^{\prime}) with the properties

Γκ,θΓκ,θ,θθ|Γκ,θ×Γκ,θd,\displaystyle\Gamma_{\kappa^{\prime},\theta^{\prime}}\subset\Gamma_{\kappa,\theta},\quad\theta^{\prime}\leq\theta|_{\Gamma_{\kappa^{\prime},\theta^{\prime}}\times\Gamma_{\kappa^{\prime},\theta^{\prime}}\setminus d}, (7.6)
κκ|Γκ,θ+2(θθ)(dx,Γκ,θ)|Γκ,θ+2θ(dx,Γκ,θΓκ,θ)|Γκ,θ\displaystyle\kappa^{\prime}\geq\kappa|_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}+2(\theta-\theta^{\prime})(dx,\Gamma_{\kappa^{\prime},\theta^{\prime}})|_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}+2\theta(dx,\Gamma_{\kappa,\theta}\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}})|_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}

such that (,)=(κ,θ,κ,θ)({\mathscr{E}},{\mathscr{F}})=({\mathscr{E}}^{\kappa^{\prime},\theta^{\prime}},{\mathscr{F}}^{\kappa^{\prime},\theta^{\prime}}).

Proof.

The sufficiency can be verified straightforwardly; we will now demonstrate the necessity. According to Theorem 6.2 and Remark 6.3, there exists an κ,θ{\mathscr{E}}^{\kappa,\theta}-quasi-open set GG with ΩGGκ,θ:=ΩΓκ,θ\Omega\subset G\subset G_{\kappa,\theta}:=\Omega\cup\Gamma_{\kappa,\theta} along with a (symmetric) bivariate smooth measure σ\sigma on G×GG\times G with respect to the part Dirichlet form (1,1)\left({\mathscr{E}}^{1},{\mathscr{F}}^{1}\right) of (κ,θ,κ,θ)\left({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}\right) on GG such that σ(Ω×Ω)=0\sigma(\Omega\times\Omega)=0 and

=1L2(G,σ¯),\displaystyle{\mathscr{F}}={\mathscr{F}}^{1}\cap L^{2}(G,\bar{\sigma}),
(u,v)=1(u,v)+σ(u~v~),u,v.\displaystyle{\mathscr{E}}(u,v)={\mathscr{E}}^{1}(u,v)+\sigma(\tilde{u}\otimes\tilde{v}),\quad u,v\in{\mathscr{F}}.

Recall that the bivariate smooth measure σ\sigma satisfies that σ¯(dx)=σ(dx,G)\bar{\sigma}(dx)=\sigma(dx,G) is smooth with respect to 1{\mathscr{E}}^{1} and σ|G×Gd2θ|G×Gd\sigma|_{G\times G\setminus d}\leq 2\theta|_{G\times G\setminus d}. Consequently, σ(Ω×(GΩ))=σ((GΩ)×Ω)=0\sigma(\Omega\times(G\setminus\Omega))=\sigma((G\setminus\Omega)\times\Omega)=0, which implies σ¯(Ω)=0\bar{\sigma}(\Omega)=0.

A straightforward computation shows that for u,v1L2(G,σ¯)u,v\in{\mathscr{F}}^{1}\cap L^{2}(G,\bar{\sigma}),

1(u,v)+σ(u~v~)=𝐃(u,v)+GΩu~(x)v~(x)(κ+2θ(,Γκ,θG)+σ¯)(dx)\displaystyle{\mathscr{E}}^{1}(u,v)+\sigma(\tilde{u}\otimes\tilde{v})=\mathbf{D}(u,v)+\int_{G\setminus\Omega}\tilde{u}(x)\tilde{v}(x)\left(\kappa+2\theta(\cdot,\Gamma_{\kappa,\theta}\setminus G)+\bar{\sigma}\right)(dx)
+(GΩ)×(GΩ)d(u~(x)u~(y))(v~(x)v~(y))(θ12σ)(dxdy).\displaystyle\quad+\int_{(G\setminus\Omega)\times(G\setminus\Omega)\setminus d}(\tilde{u}(x)-\tilde{u}(y))(\tilde{v}(x)-\tilde{v}(y))\left(\theta-\frac{1}{2}\sigma\right)(dxdy).

Define

θ:=(θ12σ)|(GΩ)×(GΩ)d,κ:=(κ+2θ(,Γκ,θG)+σ¯)|GΩ\theta^{\prime}:=\left(\theta-\frac{1}{2}\sigma\right)|_{(G\setminus\Omega)\times(G\setminus\Omega)\setminus d},\quad\kappa^{\prime}:=\left(\kappa+2\theta(\cdot,\Gamma_{\kappa,\theta}\setminus G)+\bar{\sigma}\right)|_{G\setminus\Omega}

with

Γκ,θ:=GΩ.\Gamma_{\kappa^{\prime},\theta^{\prime}}:=G\setminus\Omega.

It is evident that they satisfy (7.6).

Let us verify that (κ,θ)(\kappa^{\prime},\theta^{\prime}) is quasi-admissible. Note that the quasi-notion of κ,θ{\mathscr{E}}^{\kappa,\theta} corresponds to that of the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on Gκ,θG_{\kappa,\theta}. As stated in Corollary 3.3, GG is quasi-open, and hence, ΓΓκ,θ=Ω¯G\Gamma\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}}=\overline{\Omega}\setminus G is quasi-closed. Since κ+θ¯\kappa+\bar{\theta} is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on Gκ,θG_{\kappa,\theta}, it follows from the remarks preceding Corollary 3.2 that the restriction of κ+θ¯\kappa+\bar{\theta} to GG is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on GG. It is notable that σ¯\bar{\sigma} is also smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on GG, as it is smooth with respect to 1{\mathscr{E}}^{1}. Therefore, we deduce that κ+θ¯\kappa^{\prime}+\bar{\theta}^{\prime} is smooth with respect to the part Dirichlet form of (𝐃,H~1(Ω))(\mathbf{D},\tilde{H}^{1}(\Omega)) on GG, where θ¯(dx):=θ(dx,Γκ,θ)\bar{\theta}^{\prime}(dx):=\theta^{\prime}(dx,\Gamma_{\kappa^{\prime},\theta^{\prime}}). This verifies the quasi-admissibility of (κ,θ)(\kappa^{\prime},\theta^{\prime}).

It remains to demonstrate that =κ,θ{\mathscr{F}}={\mathscr{F}}^{\kappa^{\prime},\theta^{\prime}}. Note that

1\displaystyle{\mathscr{F}}^{1} ={u~H~1(Ω):u~=0 q.e. on ΓΓκ,θ,\displaystyle=\bigg{\{}\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }\Gamma\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}},
Γκ,θ×Γκ,θd(u~(x)u~(y))2θ(dxdy)+Γκ,θu~(x)2κ(dx)<}\displaystyle\qquad\qquad\int_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta(dxdy)+\int_{\Gamma_{\kappa,\theta}}\tilde{u}(x)^{2}\kappa(dx)<\infty\bigg{\}}
={u~H~1(Ω):u~=0 q.e. on ΓΓκ,θ,Γκ,θ×Γκ,θd(u~(x)u~(y))2θ(dxdy)\displaystyle=\bigg{\{}\tilde{u}\in\tilde{H}^{1}(\Omega):\tilde{u}=0\text{ q.e. on }\Gamma\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}},\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}\times\Gamma_{\kappa^{\prime},\theta^{\prime}}\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta(dxdy)
+Γκ,θu~(x)2(κ(dx)+2θ(dx,Γκ,θΓκ,θ))<}.\displaystyle\qquad\qquad+\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}\tilde{u}(x)^{2}\left(\kappa(dx)+2\theta(dx,\Gamma_{\kappa,\theta}\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}})\right)<\infty\bigg{\}}.

For u~1L2(G,σ¯)\tilde{u}\in{\mathscr{F}}^{1}\cap L^{2}(G,\bar{\sigma}), we have

Γκ,θ×Γκ,θd(u~(x)u~(y))2θ(dxdy)+Γκ,θu~(x)2κ(dx)\displaystyle\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}\times\Gamma_{\kappa^{\prime},\theta^{\prime}}\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta^{\prime}(dxdy)+\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}\tilde{u}(x)^{2}\kappa^{\prime}(dx)
=\displaystyle= Γκ,θu~(x)u~(y)σ(dxdy)+Γκ,θ×Γκ,θd(u~(x)u~(y))2θ(dxdy)\displaystyle\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}\tilde{u}(x)\tilde{u}(y)\sigma(dxdy)+\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}\times\Gamma_{\kappa^{\prime},\theta^{\prime}}\setminus d}(\tilde{u}(x)-\tilde{u}(y))^{2}\theta(dxdy)
+Γκ,θu~(x)2(κ(dx)+2θ(dx,Γκ,θΓκ,θ))\displaystyle\qquad\qquad+\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}\tilde{u}(x)^{2}\left(\kappa(dx)+2\theta(dx,\Gamma_{\kappa,\theta}\setminus\Gamma_{\kappa^{\prime},\theta^{\prime}})\right)
<\displaystyle< ,\displaystyle\infty,

since |Γκ,θu~(x)u~(y)σ(dxdy)|Γκ,σu~(x)2σ¯(dx)<\left|\int_{\Gamma_{\kappa^{\prime},\theta^{\prime}}}\tilde{u}(x)\tilde{u}(y)\sigma(dxdy)\right|\leq\int_{\Gamma_{\kappa^{\prime},\sigma^{\prime}}}\tilde{u}(x)^{2}\bar{\sigma}(dx)<\infty. This establishes that =1L2(G,σ¯)κ,θ{\mathscr{F}}={\mathscr{F}}^{1}\cap L^{2}(G,\bar{\sigma})\subset{\mathscr{F}}^{\kappa^{\prime},\theta^{\prime}}. The opposite inclusion can be verified similarly. This completes the proof. ∎

One may be interested in local Dirichlet forms sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}). According to Theorem 7.13, these forms are characterized as follows.

Corollary 7.14.

Let (κ,θ)(\kappa,\theta) be a quasi-admissible pair with θ¯():=θ(,Γκ,θ)\bar{\theta}(\cdot):=\theta(\cdot,\Gamma_{\kappa,\theta}). A quasi-regular and local Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(Ω¯)L^{2}(\overline{\Omega}) is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (κ,θ,κ,θ)({\mathscr{E}}^{\kappa,\theta},{\mathscr{F}}^{\kappa,\theta}) if and only if there exists a quasi-admissible measure μ\mu with the properties

ΓμΓκ,θ,μ(κ+2θ¯)|Γμ\displaystyle\Gamma_{\mu}\subset\Gamma_{\kappa,\theta},\quad\mu\geq(\kappa+2\bar{\theta})|_{\Gamma_{\mu}}

such that (,)=(μ,μ)({\mathscr{E}},{\mathscr{F}})=({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}), as defined in (7.2).

Proof.

It suffices to note that the Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) in Theorem 7.13 is local if and only if θ=0\theta^{\prime}=0. ∎

7.3. Resolution of an open problem

We close this section with a special consideration of the admissible pair (κ,θ)(\kappa,\theta) in Example 7.12. Recall that Γκ,θ\Gamma_{\kappa,\theta} is defined as (7.5), θ^():=θ(,Γ)\hat{\theta}(\cdot):=\theta(\cdot,\Gamma), and the regular Dirichlet form (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}) corresponds to the quasi-admissible pair of measures

κ=(κ+2θ(,ΓΓκ,θ))|Γκ,θ,θ=θ|Γκ,θ×Γκ,θd\kappa^{\prime}=(\kappa+2{\theta}(\cdot,\Gamma\setminus\Gamma_{\kappa,\theta}))|_{\Gamma_{\kappa,\theta}},\quad\theta^{\prime}=\theta|_{\Gamma_{\kappa,\theta}\times\Gamma_{\kappa,\theta}\setminus d}

with the defining set Γκ,θ=Γκ,θ\Gamma_{\kappa^{\prime},\theta^{\prime}}=\Gamma_{\kappa,\theta}.

Let μ0:=κ+2θ^\mu_{0}:=\kappa+2\hat{\theta}, which is an admissible measure with Γμ0=Γκ,θ\Gamma_{\mu_{0}}=\Gamma_{\kappa,\theta} as discussed in Example 7.6. Let (aμ0,𝒟(aμ0))(a_{\mu_{0}},\mathcal{D}(a_{\mu_{0}})) be defined as (7.3) (with μ=μ0\mu=\mu_{0}). Since Γμ0=Γκ,θ\Gamma_{\mu_{0}}=\Gamma_{\kappa^{\prime},\theta^{\prime}} and μ0|Γμ0=κ+2θ¯\mu_{0}|_{\Gamma_{\mu_{0}}}=\kappa^{\prime}+2\bar{\theta}^{\prime}, where θ¯():=θ(,Γκ,θ)=θ(,Γκ,θ)\bar{\theta}^{\prime}(\cdot):=\theta^{\prime}(\cdot,\Gamma_{\kappa^{\prime},\theta^{\prime}})=\theta(\cdot,\Gamma_{\kappa,\theta}), it follows from Corollary 7.14 that (aμ0,𝒟(aμ0))(a_{\mu_{0}},\mathcal{D}(a_{\mu_{0}})) is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}). The following result establishes that it is the largest among the local Dirichlet forms sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}).

Corollary 7.15.

If (,)({\mathscr{E}},{\mathscr{F}}) is a quasi-regular and local Dirichlet form on L2(Ω¯)L^{2}(\overline{\Omega}) that is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}), then (,)({\mathscr{E}},{\mathscr{F}}) is dominated by (aμ0,𝒟(aμ0))(a_{\mu_{0}},\mathcal{D}(a_{\mu_{0}})).

Proof.

According to Corollary 7.14, it holds that (,)=(μ,μ)({\mathscr{E}},{\mathscr{F}})=({\mathscr{E}}^{\mu},{\mathscr{F}}^{\mu}) for some quasi-admissible μ\mu such that ΓμΓμ0\Gamma_{\mu}\subset\Gamma_{\mu_{0}} and μμ0|Γμ\mu\geq\mu_{0}|_{\Gamma_{\mu}}. Consequently, the conclusion follows from Corollary 6.4. ∎

Particularly, if μ\mu is an admissible measure such that (aμ,𝒟(aμ))(a_{\mu},\mathcal{D}(a_{\mu})) is sandwiched between (𝐃,H01(Ω))(\mathbf{D},H^{1}_{0}(\Omega)) and (𝒜κ,θ,𝒟(𝒜κ,θ)¯)({\mathscr{A}}^{\kappa,\theta},\overline{\mathcal{D}({\mathscr{A}}^{\kappa,\theta})}), then (aμ,𝒟(aμ))(a_{\mu},\mathcal{D}(a_{\mu})) is dominated by (aμ0,𝒟(aμ0))(a_{\mu_{0}},\mathcal{D}(a_{\mu_{0}})). This conclusion addresses the open problem posed at the end of [25].

Appendix A Quasi-regularity of Dirichlet forms

In this appendix, we recall the fundamental concepts related to the quasi-regularity of Dirichlet forms. Let EE be a Hausdorff topological space with the Borel σ\sigma-algebra (E)\mathscr{B}(E) assumed to be generated by the continuous functions on EE, and let mm be a σ\sigma-finite measure on EE with support supp[m]=E\mathrm{supp}[m]=E. Denote by (E)\mathscr{B}^{*}(E) be the universal completion of (E)\mathscr{B}(E).

A.1. Quasi-regularity

Let (,)({\mathscr{E}},{\mathscr{F}}) be a Dirichlet form on L2(E,m)L^{2}(E,m). An increasing sequence of closed subsets {Fn:n1}\{F_{n}:n\geq 1\} of EE is called an {\mathscr{E}}-nest if n1Fn\bigcup_{n\geq 1}{\mathscr{F}}_{F_{n}} is 1{\mathscr{E}}_{1}-dense in {\mathscr{F}}, where Fn:={f:f=0,m-a.e. on EFn}{\mathscr{F}}_{F_{n}}:=\{f\in{\mathscr{F}}:f=0,m\text{-a.e. on }E\setminus F_{n}\}. A set NEN\subset E is called {\mathscr{E}}-polar if there exists an {\mathscr{E}}-nest {Fn:n1}\{F_{n}:n\geq 1\} such that NE(n1Fn)N\subset E\setminus\left(\bigcup_{n\geq 1}F_{n}\right). A statement is said to hold in the sense of {\mathscr{E}}-quasi-everywhere (abbreviated as {\mathscr{E}}-q.e.) if it holds outside an {\mathscr{E}}-polar set. A function ff on EE is called {\mathscr{E}}-quasi-continuous if there exists an {\mathscr{E}}-nest {Fn:n1}\{F_{n}:n\geq 1\} such that f|Fnf|_{F_{n}} is finite and continuous on FnF_{n} for each n1n\geq 1, denoted by fC({Fn})f\in C(\{F_{n}\}). If f=gf=g, mm-a.e., and gg is {\mathscr{E}}-quasi-continuous, then gg is referred to as an {\mathscr{E}}-quasi-continuous mm-version of ff. A set GEG\subset E is termed {\mathscr{E}}-quasi-open if there exists an {\mathscr{E}}-nest {Fn:n1}\{F_{n}:n\geq 1\} such that GFnG\cap F_{n} is an open subset of FnF_{n} with respect to the relative topology for each n1n\geq 1. The complement of an {\mathscr{E}}-quasi-open set is called {\mathscr{E}}-quasi-closed. It is important to emphasize that these concepts are related solely to the symmetric part of {\mathscr{E}}.

Definition A.1.

A Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(E,m)L^{2}(E,m) is called quasi-regular if the following conditions hold:

  • (1)

    There exists an {\mathscr{E}}-nest {Kn:n1}\{K_{n}:n\geq 1\} consisting of compact sets.

  • (2)

    There exists an 1{\mathscr{E}}_{1}-dense subset of {\mathscr{F}} whose elements have {\mathscr{E}}-quasi-continuous mm-versions.

  • (3)

    There exists {fk:k1}\{f_{k}:k\geq 1\}\subset{\mathscr{F}} having {\mathscr{E}}-quasi-continuous mm-versions {f~k:k1}\{\tilde{f}_{k}:k\geq 1\} and an {\mathscr{E}}-polar set NN such that {f~k:k1}\{\tilde{f}_{k}:k\geq 1\} separates the points of ENE\setminus N.

It is called regular if EE is a locally compact separable metric space, mm is a fully supported Radon measure on EE, and Cc(E){\mathscr{F}}\cap C_{c}(E) is 1{\mathscr{E}}_{1}-dense in {\mathscr{F}} as well as uniformly dense in Cc(E)C_{c}(E). It is noteworthy that a regular Dirichlet form is always quasi-regular.

It is well known that there exists an mm-tight right process

X=(Ω,,t,Xt,θt,x)X=\left(\Omega,\mathcal{F},\mathcal{F}_{t},X_{t},\theta_{t},\mathbb{P}_{x}\right)

on EE properly associated with a quasi-regular Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) in the sense that PtfP_{t}f is an {\mathscr{E}}-quasi-continuous mm-version of TtfT_{t}f for any t0t\geq 0 and fL2(E,m)p(E)f\in L^{2}(E,m)\cap\mathrm{p}\mathscr{B}(E), where Ptf(x):=xf(Xt)P_{t}f(x):=\mathbb{P}_{x}f(X_{t}) is the transition semigroup of XX and (Tt)t0(T_{t})_{t\geq 0} is the L2L^{2}-semigroup corresponding to (,)({\mathscr{E}},{\mathscr{F}}) (see [20, IV, Theorem 3.5]). Denote by (Uα)α>0(U_{\alpha})_{\alpha>0} the resolvent of XX, defined as

Uαf(x):=x0eαtf(Xt)𝑑tU_{\alpha}f(x):=\mathbb{P}_{x}\int_{0}^{\infty}e^{-\alpha t}f(X_{t})dt

for all xEx\in E and fb(E)f\in\mathrm{b}\mathscr{B}(E). The notation related to general Markov processes can be found in, e.g., [7] and [30]. The symbol Δ\Delta represents the cemetery, and ζ:=inf{t>0:Xt=Δ}\zeta:=\inf\{t>0:X_{t}=\Delta\} denotes the lifetime of XX. Note that the mm-tightness of XX means that there exists a sequence of increasing compact sets {Kn:n1}\{K_{n}:n\geq 1\} such that

m(limnTEKn<ζ)=0,\mathbb{P}_{m}\left(\lim_{n\rightarrow\infty}T_{E\setminus K_{n}}<\zeta\right)=0, (A.1)

where m:=Em(dx)x\mathbb{P}_{m}:=\int_{E}m(dx)\mathbb{P}_{x} and TA:=inf{t>0:XtA}T_{A}:=\inf\{t>0:X_{t}\in A\} denotes the first hitting time of any nearly Borel measurable set AA with respect to XX.

Without loss of generality, we may further assume that EE is a Lusin space (see [20, IV, Remark 3.2 (iii)]), and that XX is Borel measurable in the sense that Ptf(E)P_{t}f\in\mathscr{B}(E) for all t0t\geq 0 and fb(E)f\in\mathrm{b}\mathscr{B}(E) (see, e.g., [12, Corollary 3.23]). In this context, XX is referred to as a Borel right process. As established in [36], XtX_{t-} exists in EE for t<ζt<\zeta, m\mathbb{P}_{m}-a.s. (This fact will be utilized only in § 3.3.)

A set NEN\subset E is called mm-polar (for XX) if there exists a nearly Borel measurable set N~N\tilde{N}\supset N such that m(TN~<)=0\mathbb{P}_{m}(T_{\tilde{N}}<\infty)=0. A statement holds in the sense of q.e. if it holds outside an mm-polar set. A nearly Borel measurable set AEA\subset E is said to be (XX-)invariant if for every xAx\in A, x(TEA<)=0\mathbb{P}_{x}\left(T_{E\setminus A}<\infty\right)=0. The restriction of a Borel right process XX to an invariant set is a (not necessarily Borel) right process (see [30, (12.30)]), whereas its restriction to a Borel invariant set remains a Borel right process (see, e.g., [9, Lemma A.1.27]). A subset NEN\subset E is called an mm-inessential set (for XX) if m(N)=0m(N)=0 and ENE\setminus N is XX-invariant. Obviously, an mm-inessential set is mm-polar. Note that any mm-polar set is contained in an mm-inessential Borel set (see, e.g., [9, Theorem A.2.15]). A numerical function ff defined q.e. on EE is called finely continuous q.e. if there exists an mm-inessential set NN such that f|ENf|_{E\setminus N} is nearly Borel measurable and finely continuous with respect to the restricted right process X|ENX|_{E\setminus N}. A set GEG\subset E is called q.e. finely open if there exists an mm-inessential set NN such that GNG\setminus N is a nearly Borel measurable and finely open set for X|ENX|_{E\setminus N}.

The following lemma summarizes the relationships between several {\mathscr{E}}-quasi-notions and concepts related to XX.

Lemma A.2.

Let XX be a Borel right process properly associated with the quasi-regular Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) on L2(E,m)L^{2}(E,m). Then the following statements hold:

  • (1)

    A set NEN\subset E is {\mathscr{E}}-polar, if and only if it is mm-polar.

  • (2)

    An increasing sequence of closed subsets {Fn:n1}\{F_{n}:n\geq 1\} of EE is an {\mathscr{E}}-nest if and only if

    x(limnTEFn<ζ)=0\mathbb{P}_{x}\left(\lim_{n\rightarrow\infty}T_{E\setminus F_{n}}<\zeta\right)=0 (A.2)

    for q.e. xEx\in E (equivalently, for mm-a.e. xEx\in E).

  • (3)

    If ff is {\mathscr{E}}-quasi-continuous, then ff is finely continuous q.e. Conversely, if ff\in{\mathscr{F}} is finely continuous q.e., then ff is {\mathscr{E}}-quasi-continuous.

  • (4)

    A set GEG\subset E is {\mathscr{E}}-quasi-open if and only if it is q.e. finely open.

Proof.

For the first statement, see [20, IV, Theorem 5.29(i)]. The second statement is a consequence of [20, IV, Theorem 5.4 and Proposition 5.30]. The third statement can be established by repeating the proof of [9, Theorem 3.1.7]. (This proof relies solely on the essential properties of Borel right processes and the conclusions from the previous two assertions. Additionally, it is necessary to adjust N0N_{0} in this proof to an mm-polar set N0N^{\prime}_{0} such that (A.2) holds for xN0x\notin N^{\prime}_{0}.)

The necessity of the fourth statement can be concluded by repeating the proof of [9, Theorem 3.3.3]. In particular, if GG is {\mathscr{E}}-quasi-open, there exists an mm-inessential Borel set NN such that GN(EN)G\setminus N\in\mathscr{B}(E\setminus N) is finely open with respect to X|ENX|_{E\setminus N}. Conversely, let GG be q.e. finely open, and let NN be an mm-inessential set for XX such that G1:=GNG_{1}:=G\setminus N is nearly Borel measurable and finely open with respect to X|ENX|_{E\setminus N}. Take f(E)L2(E,m)f\in\mathscr{B}(E)\cap L^{2}(E,m) that is strictly positive on EE, and define

g(x):=x0TEG1etf(Xt)𝑑t=U1f(x)x(eTEG1U1f(XTEG1)).g(x):=\mathbb{P}_{x}\int_{0}^{T_{E\setminus G_{1}}}e^{-t}f(X_{t})dt=U_{1}f(x)-\mathbb{P}_{x}\left(e^{-T_{E\setminus G_{1}}}U_{1}f(X_{T_{E\setminus G_{1}}})\right).

Note that 𝐇EG11U1f(x):=x(eTEG1U1f(XTEG1))\mathbf{H}^{1}_{E\setminus G_{1}}U_{1}f(x):=\mathbb{P}_{x}\left(e^{-T_{E\setminus G_{1}}}U_{1}f(X_{T_{E\setminus G_{1}}})\right) is 11-excessive with respect to XX (see [9, Lemma A.2.4 (ii)]), and 𝐇EG11U1fU1f\mathbf{H}^{1}_{E\setminus G_{1}}U_{1}f\leq U_{1}f\in{\mathscr{F}}. It follows from, e.g., [19, Theorem 2.6] that 𝐇EG11U1f\mathbf{H}^{1}_{E\setminus G_{1}}U_{1}f\in{\mathscr{F}}. Thus, the third statement indicates that both 𝐇EG11U1f\mathbf{H}^{1}_{E\setminus G_{1}}U_{1}f and gg are {\mathscr{E}}-quasi-continuous. It is evident that {xE:g>0}=G1(B1B1r)\{x\in E:g>0\}=G_{1}\cup(B_{1}\setminus B_{1}^{r}), where B1:=EG1B_{1}:=E\setminus G_{1} and B1rB_{1}^{r} is the set of all regular points for B1B_{1}. Note that B1B1rB_{1}\setminus B_{1}^{r} is mm-polar, and hence {\mathscr{E}}-polar. Since {g>0}\{g>0\} is {\mathscr{E}}-quasi-open, we can conclude that both G1G_{1} and GG are {\mathscr{E}}-quasi-open. This completes the proof. ∎

Remark A.3.

The third statement indicates the following fact: if ff\in{\mathscr{F}} is finely continuous q.e., then there exists an mm-inessential Borel set NN such that f|ENf|_{E\setminus N} is not only nearly Borel measurable with respect to X|ENX|_{E\setminus N} but also Borel measurable on ENE\setminus N.

A.2. Quasi-homeomorphism

Let (E^,(E^))(\widehat{E},\mathscr{B}(\widehat{E})) be a second topological space with the Borel measurable σ\sigma-algebra (E^)\mathscr{B}(\widehat{E}), and let j:(E,(E))(E^,(E^))j:(E,\mathscr{B}(E))\rightarrow(\widehat{E},\mathscr{B}(\widehat{E})) be a measurable map. Define m^:=mj1\widehat{m}:=m\circ j^{-1}, the image measure of mm under jj. Then

j:L2(E^,m^)L2(E,m),f^jf^:=f^jj^{*}:L^{2}(\widehat{E},\widehat{m})\rightarrow L^{2}(E,m),\quad\widehat{f}\mapsto j^{*}\widehat{f}:=\widehat{f}\circ j

is an isometry. If jj^{*} is onto, i.e., jL2(E^,m^)=L2(E,m)j^{*}L^{2}(\widehat{E},\widehat{m})=L^{2}(E,m), then

^:={f^L2(E^,m^):jf^},\displaystyle\widehat{{\mathscr{F}}}:=\left\{\widehat{f}\in L^{2}(\widehat{E},\widehat{m}):j^{*}\widehat{f}\in{\mathscr{F}}\right\},
^(f^,g^):=(jf^,jg^),f^,g^^\displaystyle\widehat{{\mathscr{E}}}(\widehat{f},\widehat{g}):={\mathscr{E}}(j^{*}\widehat{f},j^{*}\widehat{g}),\quad\widehat{f},\widehat{g}\in\widehat{{\mathscr{F}}}

is a Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}), referred to as the image Dirichlet form of (,)({\mathscr{E}},{\mathscr{F}}) under jj. We denote (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}) as j(,)j({\mathscr{E}},{\mathscr{F}}).

Definition A.4.

Let (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}) be another Dirichlet form on L2(E^,m^)L^{2}(\widehat{E},\widehat{m}) with supp[m^]=E^\text{supp}[\widehat{m}]=\widehat{E}. The Dirichlet form (,)({\mathscr{E}},{\mathscr{F}}) is called quasi-homeomorphic to (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}) if there exists an {\mathscr{E}}-nest {Fn:n1}\{F_{n}:n\geq 1\}, an ^\widehat{{\mathscr{E}}}-nest {F^n:n1}\{\widehat{F}_{n}:n\geq 1\} and a map j:n1Fnn1F^nj:\bigcup_{n\geq 1}F_{n}\rightarrow\bigcup_{n\geq 1}\widehat{F}_{n} such that

  • (1)

    jj is a topological homeomorphism from FnF_{n} to F^n\widehat{F}_{n} for each n1n\geq 1.

  • (2)

    m^=mj1\widehat{m}=m\circ j^{-1}.

  • (3)

    (^,^)=j(,)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}})=j({\mathscr{E}},{\mathscr{F}}), the image Dirichlet form of (,)({\mathscr{E}},{\mathscr{F}}) under jj.

Such a map jj is called a quasi-homeomorphism from (,)({\mathscr{E}},{\mathscr{F}}) to (^,^)(\widehat{{\mathscr{E}}},\widehat{{\mathscr{F}}}).

Note that a quasi-homeomorphism keeps the quasi-notions invariant; see, e.g., [10, Corollary 3.6].

References

  • [1] K. Akhlil. Probabilistic solution of the general Robin boundary value problem on arbitrary domains. Int. J. Stoch. Anal., 2012:1–17, Dec. 2012.
  • [2] K. Akhlil. Locality and domination of semigroups. Results Math., 73(2):59, June 2018.
  • [3] W. Arendt. The Laplacian with Robin boundary conditions on arbitrary domains. Potential Anal., 19(4):341–363, 2003.
  • [4] W. Arendt and M. Warma. Dirichlet and Neumann boundary conditions: What is in between? J. Evol. Equ., 3(1):119–135, Feb. 2003.
  • [5] S. Arora, R. Chill, and J.-D. Djida. Domination of semigroups generated by regular forms. Proc. Am. Math. Soc., 2024.
  • [6] A. Beurling and J. Deny. Dirichlet spaces. Proc. Natl. Acad. Sci. U. S. A., 45:208–215, 1959.
  • [7] R. M. Blumenthal and R. Getoor. Markov processes and potential theory. Pure and Applied Mathematics, Vol. 29. Academic Press, New York-London, 1968.
  • [8] B. E. Breckner and R. Chill. The Laplace operator on the Sierpinski gasket with Robin boundary conditions. Nonlinear Anal. Real World Appl., 38:245–260, Dec. 2017.
  • [9] Z.-Q. Chen and M. Fukushima. Symmetric Markov processes, time change, and boundary theory, volume 35 of London Mathematical Society Monographs Series. Princeton University Press, Princeton, NJ, 2012.
  • [10] Z.-Q. Chen, Z. M. Ma, and M. Röckner. Quasi-homeomorphisms of Dirichlet forms. Nagoya Math. J., 136(2):1–15, 1994.
  • [11] R. Chill and B. Claus. Domination of nonlinear semigroups generated by regular, local Dirichlet forms. Ann. Mat. Pura Appl. (1923 -), 2024.
  • [12] P. J. Fitzsimmons. On the quasi-regularity of semi-Dirichlet forms. Potential Anal., 15(3):151–185, 2001.
  • [13] P. J. Fitzsimmons and R. Getoor. Revuz measures and time changes. Math. Z., 199(2):233–256, 1988.
  • [14] G. B. Folland. Real analysis. Pure and Applied Mathematics (New York). John Wiley & Sons, Inc., New York, second edition, 1999.
  • [15] M. Fukushima, Y. Oshima, and M. Takeda. Dirichlet forms and symmetric Markov processes, volume 19 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, extended edition, 2011.
  • [16] R. Getoor and M. J. Sharpe. Naturality, standardness, and weak duality for Markov processes. Z. Wahrscheinlichkeitstheorie verw Gebiete, 67(1):1–62, 1984.
  • [17] J. Glück and J. Mui. Non-positivity of the heat equation with non-local Robin boundary conditions. (arXiv:2404.15114), Apr. 2024. arXiv:2404.15114.
  • [18] M. Keller, D. Lenz, M. Schmidt, M. Schwarz, and M. Wirth. Boundary representations of intermediate forms between a regular Dirichlet form and its active main part. arXiv:2301.01035.
  • [19] Z.-M. Ma, L. Overbeck, and M. Röckner. Markov processes associated with semi-Dirichlet forms. Osaka J. Math., 32(1):97–119, 1995.
  • [20] Z. M. Ma and M. Röckner. Introduction to the theory of (nonsymmetric) Dirichlet forms. Universitext. Springer-Verlag, Berlin, Berlin, Heidelberg, 1992.
  • [21] Z. M. Ma and M. Röckner. Markov processes associated with positivity preserving coercive forms. Can. J. Math., 47(4):817–840, 1995.
  • [22] A. Manavi, H. Vogt, and J. Voigt. Domination of semigroups associated with sectorial forms. J. Operat. Theor., 54(1):9–25, 2005.
  • [23] S. Mataloni. Representation formulas for non-symmetric Dirichlet forms. Z. Anal. Anwendungen, 18(4):1039–1064, Dec. 1999.
  • [24] E.-M. Ouhabaz. Invariance of closed convex sets and domination criteria for semigroups. Potential Anal., 5(6):611–625, Dec. 1996.
  • [25] N. A. Oussaid, K. Akhlil, S. B. Aadi, and M. E. Ouali. Generalized nonlocal Robin Laplacian on arbitrary domains. Arch. Math., 117(6):675–686, Dec. 2021.
  • [26] D. W. Robinson. On extensions of local Dirichlet forms. J. Evol. Equations, 17(4):1151–1182, 2016.
  • [27] M. Röckner and B. Schmuland. Quasi-regular Dirichlet forms: examples and counterexamples. Can. J. Math., 47(1):165–200, 1995.
  • [28] A. V. Santiago and M. Warma. A class of quasi-linear parabolic and elliptic equations with nonlocal Robin boundary conditions. J. Math. Anal. Appl., 372(1):120–139, Dec. 2010.
  • [29] M. Schmidt. A note on reflected Dirichlet forms. Potential Anal., 52(2):245–279, 2020.
  • [30] M. Sharpe. General theory of Markov processes, volume 133 of Pure and Applied Mathematics. Academic Press, Inc., Boston, MA, 1988.
  • [31] M. J. Sharpe. Exact multiplicative functionals in duality. Indiana Univ. Math. J., 21:27–60, 1971.
  • [32] M. L. Silverstein. Symmetric Markov processes. Springer-Verlag, Berlin-New York, 1974.
  • [33] M. L. Silverstein. Boundary theory for symmetric Markov processes. Springer-Verlag, Berlin-New York, 1976.
  • [34] P. Stollmann. Closed ideals in Dirichlet spaces. Potential Anal., 2(3):263–268, Sept. 1993.
  • [35] A. Vélez-Santiago. Solvability of linear local and nonlocal Robin problems over C(Ω¯){C}(\overline{\Omega}). J. Math. Anal. Appl., 386(2):677–698, Feb. 2012.
  • [36] J. B. Walsh. Markov processes and their functional in duality. Z. Wahrscheinlichkeitstheorie verw Gebiete, 24(3):229–246, Sept. 1972.
  • [37] J. Ying. Bivariate Revuz measures and the Feynman-Kac formula. Ann. l’Inst. Henri Poincaré, Probab. Stat., 32(2):251–287, 1996.
  • [38] J. Ying. Killing and subordination. Proc. Am. Math. Soc., 124(7):2215–2222, 1996.
  • [39] J. Ying. Strong subordination of symmetric Dirichlet forms. Chin. Ann. Math. Ser. B, 19(1):81–86, 1998.