On domination for (non-symmetric) Dirichlet forms
Abstract.
The primary aim of this article is to investigate the domination relationship between two -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 extensions2020 Mathematics Subject Classification:
Primary 31C25, 60J46, 60J57, 60J50, 47D07.1. Introduction
Consider an open set . Let and denote the Laplace operators on subject to the Dirichlet boundary condition and the Neumann boundary condition , respectively. Here, denotes the space of all square-integrable functions on , represents the boundary of , and denotes the exterior normal derivative on the boundary (whose existence requires certain regularity assumptions). Denote by and the strongly continuous semigroups generated by and , respectively. In [4], Arendt and Warma investigated the question of which self-adjoint operator generates a semigroup that is “sandwiched” between and in the sense that
(1.1) |
where represents the family of all positive functions in . Their main result demonstrates that, under certain necessary conditions, corresponds to the Laplace operator subject to a Robin boundary condition determined by a certain admissible measure on (see Example 7.6). This operator and its semigroup are denoted by and , respectively. Specifically, when is absolutely continuous with respect to the surface measure on with density , corresponds to the classical Robin boundary condition:
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 and are the closed forms associated with the strongly continuous semigroups and , and both semigroups are positive (i.e., for any non-negative , and ), then
(1.2) |
is equivalent to the conditions that is an ideal of and that
where denotes the set of non-negative elements in (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 with , it holds that ), 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 -semigroups from (1.2) to the subordination between two Markov transition semigroups: is said to be subordinated to if
(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 , often referred to as the subprocess, is always derived from the Markov process corresponding to 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 -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 and satisfying (1.2), there exists a positive measure on , which can be expressed as detailed in (3.8), such that
(1.4) | ||||
where denotes the marginal measure of . 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 is necessary, whereas 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 satisfying (1.1) is determined by some quasi-admissible measure 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 in a Dirichlet space, denotes its quasi-continuous modification by default. Given a topological space , represents the collection of Borel measurable sets or Borel measurable functions, and denotes the collection of all universally measurable sets or universally measurable functions. For a class of functions , and denote the families of non-negative elements and bounded elements, respectively, within .
2. Decomposition of domination
Let be a measurable space, and let be a -finite measure on . The norms of and are denoted by and , respectively, while the inner product of is denoted by . A (non-symmetric) Dirichlet form on is defined as a coercive closed form that satisfies the Markovian property, as described in, e.g., [20, I, Definition 4.5]. The -norm on is denoted by , where for all . The Dirichlet form is called symmetric if, in addition, for all .
For a given Dirichlet form on , it is well known that is a subalgebra of (see [20, I, Corollary 4.15]), and forms a sublattice of (see [20, I, Proposition 4.11]).
If are subalgebras of , we say that is an algebraic ideal in if and imply . For sublattices of , is called an order ideal in if , and imply .
For two Dirichlet forms and , we say that dominates (or equivalently, is dominated by ) and write or 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 and be two Dirichlet forms on , whose associated semigroups (as described in [20, I, Remark 2.9(ii)]) are denoted by and , respectively. The following assertions are equivalent:
-
(1)
for all and all non-negative .
-
(2)
, is an algebraic ideal in , and
(2.1) for all non-negative .
-
(3)
, is an order ideal in , and (2.1) holds for all non-negative .
Proof.
The equivalence between the first and third assertions was established in [22, Corollary 4.3]. The proof of (3)(2) is straightforward. It suffices to demonstrate that, under the assumptions of the third assertion, if and , then . This follows immediately, as and is an order ideal in .
The proof of (2)(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 and with . Our goal is to show that . Since and are lattices, and by [20, I, Proposition 4.17], we may assume without loss of generality that and and are bounded. Let and . Consider the function
It satisfies the following inequality
(2.2) |
for . The function is Lipschitz continuous with and . By [20, I, Proposition 4.11], we have . Therefore, by the second assertion, . It is evident that in as . We now aim to show that
which, by the Banach-Saks theorem, will imply that . Indeed, it follows from and (2.1) that
Using the property (2.2) and applying [20, I, Proposition 4.11] to , we get
Thus, we have
This completes the proof. ∎
We introduce two additional notions.
Definition 2.2.
Let and be two Dirichlet forms on .
-
(1)
We say that is subordinate to if and (2.1) holds for all non-negative . Furthermore, is strongly subordinate to if, in addition, is dense in with respect to the -norm.
-
(2)
We say that is a Silverstein extension of if is an algebraic ideal in and on .
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, if and only if is subordinate to and is an algebraic ideal in (or equivalently, is an order ideal in ). By the proof of Lemma 2.1, is a Silverstein extension of if and only if is an order ideal in and on . In particular, if is a Silverstein extension of , then and is subordinate to .
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 and be two Dirichlet forms on such that . Denote as the closure of in with respect to the -norm, and define
Then, is strongly subordinate to , and is a Silverstein extension of . Furthermore, if is another Dirichlet form on that satisfies these two conditions, then .
Proof.
The first assertion that is strongly subordinate to follows directly from the definition of . We will now demonstrate that is a Silverstein extension of . Consider and . Our objective is to establish that . Indeed, we may take a sequence such that , , and converges to , -a.e. Since is an algebraic ideal in , it follows that . Furthermore, according to [15, Theorem 1.4.2], we have
By the Banach-Saks theorem, there exists a subsequence of and such that
with respect to the -norm. Note that , -a.e. Therefore, , -a.e., confirming that .
Let be another Dirichlet form that satisfies the two specified conditions. Since is a Silverstein extension of , it follows that for all . The strong subordination of to implies that is dense in with respect to the -norm. Consequently, we have and for all . 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 be a Lusin topological space and a -finite measure on with support . We consider a quasi-regular Dirichlet form on . Let with lifetime be an -tight Borel right process properly associated with . The cemetery is denoted by . We may assume that contains a distinguished point such that for all . The goal of this section is to demonstrate that the killing transformation on induces the domination of .
3.1. Part process
We first investigate a specific killing transformation that results in a part process defined on a finely open set.
Let be a finely open set, which is nearly Borel measurable since is Borel measurable, and define . We denote , and for , we define
and
According to [30, (12.24)],
(3.1) |
is a right process on the Radon space with lifetime . This process is commonly referred to as the part process of on , and its Dirichlet form, as expressed in (3.2), is called the part Dirichlet form of on .
Lemma 3.1.
The right process is -tight and is properly associated with the quasi-regular Dirichlet form given by
(3.2) | ||||
on , where denotes the -quasi-continuous -version of .
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 . 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 within .
Define for . Then is finely continuous, and hence -quasi-continuous by Lemma A.2 (3). It follows that , where denotes the set of all regular points for . Note that is semipolar, and hence also -polar. Consider an -nest consisting of compact sets such that and . Let and . It is straightforward to verify that is a compact subset of . Define . Note that . For -a.e. , we have
By the -tightness of , it suffices to demonstrate that for -a.e. ,
In fact, it follows from the quasi-left-continuity of and [12, (4.8)] that
where . Since , we obtain
This completes the proof. ∎
It is important to examine the relationship between -quasi-notion and -quasi-notion. Let be an -nest, and define for . Since
it follows from Lemma A.2 (2) that is an -nest. In particular, the restriction of an -quasi-continuous function to is -quasi-continuous, and is -polar if is -polar.
When is a regular and symmetric Dirichlet form, it has been established in [9, Theorem 3.3.8 (iii)] that -polar sets are equivalent to -polar sets through the use of -capacity. By employing a standard transfer method via quasi-homeomorphism, this result also holds in the quasi-regular case.
Corollary 3.2.
A set is -polar if and only if it is -polar.
Proof.
Since -polarity and -polarity are only related to the symmetric parts of and , we only need to consider the special case where is a symmetric Dirichlet form. It is well known that is quasi-homeomorphic to another regular Dirichlet form on via a quasi-homeomorphism ; see, e.g., [9, Theorem 1.4.3]. Let and be the -nest and -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 is an open subset of for all . Define
Since is a topological homeomorphism, it follows that is an open subset of . In particular, is an -quasi-open set. As established in [9, Theorem 3.3.8 (iii)], the -polarity is equivalent to the -polarity. Note that -polarity is also equivalent to the -polarity, as stated in, e.g., [9, Exercise 1.4.2 (ii)]. It remains to demonstrate that the -polarity is equivalent to the -polarity.
We aim to prove that is quasi-homeomorphic to via . It has been indicated that forms an -nest. Similarly, constitutes an -nest. It is evident that is measurable and serves as a topological homeomorphism from to . Note that . It suffices to demonstrate that
(3.3) |
(Consequently, can be shown through the equivalence between -quasi-notion and -quasi-notion.) Consider , which can be treated as a function in through zero extension. Since is onto, there exists a function such that . It is straightforward to show that , -a.e. on . Thus, , and , which establishes (3.3). This completes the proof. ∎
We present another fact that will be utilized subsequently.
Corollary 3.3.
Let . Then, is -quasi-open if and only if it is -quasi-open.
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 is called a multiplicative functional (MF, for short) of if it satisfies the following conditions:
-
(1)
The map is decreasing, right continuous and takes values in for all ;
-
(2)
is adapted, i.e., for every ;
-
(3)
holds for all and .
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 and any , the set is a -null set. This type of definition is referred to as a weak MF in [30, §54]. In contrast, an MF that satisfies is called a perfect MF. In fact, since a weak MF always admits a perfect regularization (see [30, (55.19)]), it follows that is equivalent to a perfect MF in the sense that
provided that the set , defined below, is nearly Borel measurable.
For an MF , we define , the set of permanent points of , and , the lifetime of . There is no loss of generality in assuming that for since we can always replace with another equivalent right continuous MF, given by . Particularly, for all . In addition, note that if is nearly Borel measurable and , then and, in particular, ; see [30, (57.2)].
Let denote the family of all MFs of such that is finely open and . Further, we define
Since a fine open set is nearly Borel measurable, every is a right MF in the sense of [30, (57.1)]. In this context, we can define a probability measure on for each by
(3.4) |
where , are the killing operators on defined by if and if , and is the -algebra generated by the maps with and . According to, e.g., [30, (61.5)], the restriction of to is a right process on , referred to as the (-)subprocess of and denoted by or . A more intuitive construction for subprocesses can also be found in [7, III§3]. Clearly, the transition semigroup and the resolvent of are given by
for and . Note that , but it is not necessarily Borel measurable even if is Borel measurable.
Remark 3.6.
An MF is termed exact if for any and every sequence , it holds that , a.s., as . If is exact, then is finely open, implying that ; see [30, (56.10)]. Furthermore, every is exact; see [7, III, Corollary 4.10].
It should be noted that in [37], represents all exact MFs of , which differs from its usage in this paper. In fact, for any finely closed set , is an MF of , with the set of its permanent points being the fine open set . Namely, in our context. However, is not exact, as (not ; see [30, (10.4)]).
It is also important to note that another MF defined by the first hitting time is typically exact; however, its set of permanent points is
(3.5) |
which is not necessarily equal to . Killing by can also be understood as killing the part process using . This killing transformation yields a right process, denoted by , on , as examined in [12, Theorem 5.10]. It is noteworthy that is -polar, and hence -inessential for . In particular, the restriction of to is precisely the part process .
We present a simple observation for future reference. Recall that , as expressed in (3.1), denotes the part process of on when is finely open.
Lemma 3.7.
If , then , and the subprocess is the same as the subprocess . Conversely, if is finely open and , then , whose set of permanent points is , and the subprocess is the same as .
Proof.
It is straightforward to verify the conditions of Definition 3.4 to obtain that if , then is an MF of . The set of its permanent points is clearly , since , -a.s. for all . Note that and . Thus, . This implies that . To establish the identification between and , we observe that
are the killing operators of . The probability measure of on for is given by
It follows from that
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 -stopping time is called a (perfect) terminal time if the set is empty; see [30, (12.1)]. Typical examples of terminal times include the first hitting time and the first entrance time for any nearly Borel set . Consider a terminal time such that . We have , and we denote by (with lifetime ) the subprocess of that is killed by ; see [30, (12.23)]. By employing a similar argument, it is straightforward to verify the following facts:
-
(1)
If is an MF of , then is an MF of with the same set of permanent points.
-
(2)
For and , it holds that . Furthermore, the subprocess is the same as .
Additionally, according to [37, Theorem 2.2], every admits the decomposition
(3.6) |
where , vanishes on the diagonal of , , is a positive continuous additive functional of , is disjoint from such that , where , a.s. (Every exact terminal time admits a representation of the form ; see [31, Theorem 6.1].) Note that the product and the integral in (3.6) can diverge only at . See also [31, Theorem 7.1].
3.3. Bivariate Revuz measures of MFs
For an MF of , we define its bivariate Revuz measure (relative to ) by
(3.7) |
The existence of is referred to in [13]. (In fact, is a raw additive functional of in the sense of [13, Definition 2.3]. Hence, the limit in (3.7) exists in , as demonstrated in [13, Proposition 2.5]). For , we write as . Denote by and the marginal measures of , i.e., for any ,
Note that any function defined on is extended to by setting . Similarly, for , whenever or is , we define . Particularly, the actual integration interval in (3.7) is .
The significance of bivariate Revuze measures lies in the representation of the Dirichlet form associated with the subprocess . Let . The process is called nearly -symmetric if its transition semigroup acts on as a strongly continuous contraction semigroup, and its infinitesimal generator with domain satisfies the sector condition. This condition states that there exists a finite constant such that
for all . The following theorem characterizes the Dirichlet form of for the case where . Its general form will be presented in Theorem 3.15.
Theorem 3.9.
Let . Then the subprocess is nearly -symmetric, and the associated Dirichlet form on is given by
Proof.
See [37, Theorem 3.10]. ∎
We present a characterization of the bivariate Revuz measure for , expressed as in (3.6). Let be the Lévy system of (see [30, §73]). Note that is a positive continuous additive functional of , whose Revuz measure with respect to is denoted by . (The Revuz measure of is denoted by .) As established in [37, Theorem 4.6], the following representation of holds:
(3.8) |
where is referred to as the canonical measure of (off the diagonal ), and denotes the Dirac measure at .
Remark 3.10.
If is symmetric, then the canonical measure of coincides exactly with the jumping measure of , as stated in [9, Theorem 4.3.3 and Proposition 6.4.1]. When is regular but non-symmetric, the jumping measure of can be defined using the method outlined in [15, (3.2.7)]; see also [23, Theorem 2.6]. Furthermore,
(3.9) |
holds for any 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 of in place of . Thus, is also identical to . If is only quasi-regular, a similar conclusion can be obtained by the transfer method using quasi-homeomorphism. Particularly, it holds that
(3.10) |
for any .
A -finite positive measure on is called a smooth measure (with respect to ) if it charges no -polar sets and there exists an -nest consisting of compact sets such that for all . Let be a positive measure on , and let and denote its left and right marginal measures on , respectively. Then, is called a bivariate smooth measure (with respect to ) if
is smooth and , where is the canonical measure of .
When , both and are smooth measures with respect to ; see the proof of [38, Theorem 3.4]. (This proof demonstrates the existence of two q.e. strictly positive and -quasi-continuous functions and such that . Taking an -nest consisting of compact sets such that and on , it can be readily verified that for all .) According to (3.8), the bivariate Revuz measure is also bivariate smooth. The converse also holds true; that is, the following statement is valid.
Proposition 3.11.
The measure on is a bivariate smooth measure if and only if there exists such that , the bivariate Revuz measure of .
Proof.
Remark 3.12.
It should be noted that if is local (see [20, V, Definition 1.1]), meaning that , then every bivariate smooth/Revuz measure vanishes off the diagonal . In this case, killing by reduces to perturbation by continuous additive functionals, as examined in, e.g., [20, IV, §4c]. In this context, we have
where .
3.4. Killing by general MF
Let us turn to consider the killing transformation induced by a general , and we denote for convenience.
In light of Lemma 3.7 and Remark 3.8 (2), the killing transformation induced by can be interpreted as first applying a killing by to obtain the part process , as discussed in Section 3.1. This is followed by a second killing on at the terminal time (, a.s.), and finally a killing transformation using , where represents the subprocss of that is killed by . Note that is properly associated with the part Dirichlet form , 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 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 and . The marginal measures and do not charge either -polar sets or -polar sets.
Proof.
Remark 3.14.
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 , since .
Theorem 3.15.
Let and . Then, the subprocess is a nearly -symmetric right process on , which is properly associated with the Dirichlet form given by
(3.11) | ||||
where is defined as in (3.2) with . Furthermore, is -tight, and is quasi-regular on .
Proof.
Note that is a right process on the Radon space . We assert that is -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 , where and . Let be an increasing sequence of compact sets as specified in (A.1). Clearly, for any , we have
Therefore, it follows from (3.4) that
Note that (3.4) can be applied to this event because it remains valid for all , where is the universal completion of (see [30, Page 286]). This establishes the -tightness of .
It remains to demonstrate that is nearly -symmetric and to obtain the expression (3.11) for its associated Dirichlet form .
Consider first that , which allows for the decomposition given in (3.6). Let be the subprocess of that is killed by the terminal time (namely, killed by the MF ). As demonstrated in [37, Theorem 4.3], is -nearly symmetric on , and (3.11) holds with the bivariate Revuz measure given by
(3.12) |
By applying [12, Corollary 3.23] to , we can select an -inessential Borel set for such that the restricted Borel right process is associated with the same Dirichlet form as that of . Notably, is -polar by [38, Corollary 3.3]. Consequently, it follows from Lemma 3.13 that
(3.13) |
where and represent the right and left marginal measures of , respectively. According to Remark 3.8 (2), , which clearly indicates that . We then have
By applying Theorem 3.9 to and , we can conclude that (and hence ) acts on as a strongly continuous contraction semigroup whose infinitesimal generator satisfies the sector condition. The Dirichlet form of the subprocess is evidently the same as that of . By Theorem 3.9, it is given by
where is the bivariate Revuz measure of with respect to . To finalize (3.11), it remains to establish
(3.14) |
(Any function in can be treated as a function in through zero extension, as we have shown (3.13).) In fact, we have
From (3.12), we can deduce that
Therefore, (3.14) is established.
Finally, we consider the case where . Let , which is finely open with respect to , and let . As in the proof of Lemma 3.13, we may assume without loss of generality that is a special, Borel standard process and . Let represent the part process of on associated with the part Dirichlet form , and let be an -inessential Borel set such that the restricted process is a Borel standard process on . It follows from Lemma 3.13 that
(3.15) |
Note that by Lemma 3.7, whose restriction to is an MF in . The transition semigroup of the subprocess is given by
for and . Based on our examination in the previous case, we conclude that (and hence ) acts on as a strongly continuous contraction semigroup, whose infinitesimal generator satisfies the sector condition. The Dirichlet form of is identical to that of , which has the following form:
where is the bivariate Revuz measure of with respect to , and and represent the right and left marginal measures of , respectively. Given (3.15), it remains to show that
where can be regarded as a function on via zero extension. In fact, since whenever and , we have
This completes the proof. ∎
Remark 3.16.
As explained before Corollary 3.2, the -quasi-notion can be inherited from the -quasi-notion. We will now demonstrate that the -quasi-notion and the -quasi-notion are equivalent. Without loss of generality, assume that . Let
As shown in [9, Lemma 5.1], is a Dirichlet form on . (Recall that charges no -polar sets by Lemma 3.13.) According to [9, Theorem 5.1.4], the -nest is equivalent to the -nest. (Although only symmetric Dirichlet forms are considered in [9], the -quasi-notion pertains solely to the symmetric part of , 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
Thus, any -nest is also an -nest. Conversely, it can be easily verified that is strongly subordinate to . (To demonstrate the denseness of in , it suffices to consider the -nest consisting of compact sets for which for each , and note that is -dense in .) By [38, Corollary 3.3], any -nest is also an -nest. Therefore, the -nest is equivalent to the -nest. In particular, -quasi-continuity (resp. -polar set) is equivalent to -quasi-continuity (resp. -polar set).
In the special case where is symmetric, it is evident that is symmetric if and only if is symmetric, i.e., . 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 such that . Then, the Dirichlet form associated with the subprocess is dominated by . Furthermore, when considered as a Dirichlet form on , 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 . (This is also clear from the inequality for all ). We now proceed to demonstrate the quasi-regularity of on . To maintain clarity, we denote on as . It is straightforward to verify that each compact subset of is also compact in because is endowed with the relative topology from . Consequently, an -nest consisting of compact sets is also an -nest. Therefore, the quasi-regularity of is a consequence of the quasi-regularity of . ∎
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 with .
Let and be two Dirichlet forms on such that . Denote by 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 to . It should be pointed out that the Dirichlet form 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 -nest is also an -nest. In particular, is quasi-regular provided that is quasi-regular.
The following result essentially establishes that strong subordination corresponds to a killing transformation induced by an MF in . 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 is quasi-regular on . Then there exists a Borel right process properly associated with and an MF such that the subprocess of , killed by , is properly associated with .
Proof.
Let be a metric on that induces the topology of , and let denote the family of all -uniformly continuous functions on . The set consists of all bounded functions in . Since is Lusin, there exists a countable subset that is dense in with respect to the uniform norm (see [30, A.2]). Define , where . As is -finite, we can select a sequence of increasing Borel sets such that and . Let
Clearly, is countable. Denote by the set of all positive rationals.
Since is quasi-regular by Lemma 4.1, there exists a Borel right process properly associated with . Namely, for any and , is an -quasi-continuous -version of , where is the resolvent of , and is the -resolvent of . See [20, IV, Proposition 2.8]. Similarly, there exists a Borel right process such that its resolvent is an -quasi-continuous -version of for any and , where is the -resolvent of .
Since strong subordination implies domination, it follows from Lemma 2.1 (1) that
(4.1) |
for all positive and . Note that is -quasi-continuous by Lemma 4.1. Thus, there exists an -nest such that
In particular, it follows from (4.1) that
(4.2) |
Let us now construct an increasing sequence of -inessential Borel sets for . We begin by selecting an -inessential Borel set for that contains . Since , -a.e., and is -quasi-continuous, for all , there exists an -polar set such that
Note that is also -polar by Lemma 4.1. Thus, we can select an -inessential Borel set for such that . For , once and have been defined, we construct and using the following method: Select an -polar set that contains such that
(4.3) |
and then choose an -inessential Borel set for such that .
Define and . It is evident that is a Borel invariant set for . Hence, the restricted process remains a Borel right process. Let be the trivial extension of to , namely, will remain at indefinitely if . Clearly, is also a Borel right process properly associated with . Denote by the resolvent of . Note that for all and ,
On the other hand, it follows from (4.3) that
Particularly,
(4.4) |
provides a well-defined resolvent on .
We assert that
(4.5) |
It suffices to consider (and fix) . From (4.2), we know that (4.5) holds for and . By applying the monotone convergence theorem (for ) and the dominated convergence theorem (for the parameters ), we can extend (4.5) to all and all . Fix . Since is dense in and for all and , it follows that (4.5) holds for all positive . Then, by utilizing [30, Proposition A2.1], (4.5) holds for all positive and bounded . Note that the Lusin topological space is normal. Therefore, Urysohn’s lemma (see, e.g., [14, Lemma 4.15]) indicates that (4.5) holds for , where is an arbitrary compact set. Since the finite measure is inner regular on (see [30, Theorem A2.3]), we can further obtain (4.5) with for any Borel set . This result can be extended to all by [14, Theorem 2.10] and the monotone convergence theorem. For any , there exist such that and . Therefore, we can eventually conclude (4.5) for all .
For any , it is straightforward to see that
Evidently, for all . Hence, it follows from [7, III, Theorem 4.9] that is exactly subordinate to in the sense of [7, III, Definition 4.8]. In view of [30, Theorem (56.14)], there exists a (unique) exact MF of such that
It is easy to verify that . (If , then for all and . We must have and for all , which leads to a contradiction.) Particularly, and is the resolvent of the subprocess .
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 -nest if and only if it is an -nest. Particularly, is an -polar set if and only if it is an -polar set. A function is -quasi-continuous if and only if it is -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 is quasi-regular.
Theorem 4.4.
Let and be two quasi-regular Dirichlet forms on such that , and let denote the unique Dirichlet form obtained in Theorem 2.4. Then, there exists a Borel right process properly associated with , a finely open set with respect to satisfying , and an MF , where is the part process of on , such that
-
(1)
is properly associated with .
-
(2)
The subprocess of killed by is properly associated with .
Proof.
According to Theorem 2.4, is a Silverstein extension of . It follows that is a closed order ideal in . Therefore, the argument presented by Stollman [34] demonstrates that there exists an -quasi-open set such that is the part Dirichlet form of on ; see also [12, Remark 5.13]. It is evident that , as is a Dirichlet form on .
By Lemma A.2 (4) (and its proof), we can take a Borel right process properly associated with and assume, without loss of generality, that is a Borel, finely open set with respect to . It is straightforward to verify that if is an -inessential Borel set for the part process , then is also finely open with respect to , and the restricted right process is precisely the part process of on . Since is properly associated with the quasi-regular Dirichlet form , as investigated in Lemma 3.1, there exists an -inessential Borel set for such that is a Borel right process. Define . This set is finely open with respect to , and is properly associated with . Clearly, is also properly associated with since any -nest consisting of compact sets is also an -nest.
Recall that is strongly subordinate to . It follows from Corollary 4.3 that the -polar set is also -polar. Therefore, it is straightforward to verify that is quasi-regular on . By following the argument presented after (4.1) in the proof of Theorem 4.2 (considering the restrictions of to in (4.5) instead) for and on , we can select an -inessential Borel set for and an MF such that the subprocess of killed by is properly associated with on .
Define . Then is finely open with respect to , , and is properly associated with . In addition, , and the subprocess of killed by is properly associated with on both and . This completes the proof. ∎
Remark 4.5.
According to this proof, the fine open set in the theorem can be selected as a Borel set, and the part process can likewise be taken as a Borel right process.
The following corollary readily follows from Lemma 3.7.
Corollary 4.6.
Let and be two quasi-regular Dirichlet forms on such that . Then, there exists a Borel right process properly associated with and with such that the subprocess is properly associated with .
5. Domination is killing: General case
Now, we remove the assumption of quasi-regularity for . In this setting, a right process corresponding to on may not exist. However, by following the approach of Silverstein [32], we can extend the space 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 be a quasi-regular Dirichlet form on , and let be another Dirichlet form on , not necessarily quasi-regular, such that . Denote by the unique Dirichlet form obtained in Theorem 2.4. Then, there exists a locally compact separable metric space and a measurable map such that, by defining , the operator is a unitary map from onto , and the following hold:
-
(i)
The image Dirichlet form is a regular Dirichlet form on ;
-
(ii)
There exists an -quasi-open subset of with such that is a quasi-homeomorphism that maps to , the part Dirichlet form of on ;
-
(iii)
is a quasi-homeomophism that maps to the quasi-regular image Dirichlet form on .
Furthermore, there exists a Hunt process on , with being taken as a finely open set with respect to , and an MF , where is the part process of on , such that is properly associated with , is properly associated with , and the subprocess is properly associated with .
Proof.
Note that is a Silverstein extension of , and according to Lemma 4.1, 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 and . This yields an -nest consisting of compact sets, a locally compact metrizable space , and a Borel measurable map
such that is a topological homeomorphism from to for each , is a fully supported Radon measure on , and is unitary from onto . Furthermore, is a regular Dirichlet form on , and there exists an -quasi-open subset of with such that is a quasi-homoemorphism that maps to the quasi-regular Dirichlet form on . In particular, there exists another -nest such that is an -nest.
Since is strongly subordinate to , it follows from Corollay 4.3 that the -nest is also an -nest. Clearly, is a topological homoemorphism for each . As is onto , is well-defined as the image Dirichlet form on . To verify that is a quasi-homeomorphism that maps to , it remains to show that forms an -nest. In fact, by the definition of the image Dirichlet form, we have
(5.1) | ||||
As a result, is -dense in . This confirms that is indeed an -nest.
For the second part of the statements, we observe that the existence of and the fact that can be chosen as a finely open set with respect to follow directly from [9, Theorem 6.6.5]. It is straightforward to verify that is strongly subordinate to and that is a Silverstein extension of . By applying Theorem 4.4 to these Dirichlet forms, we can further identify an MF 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 be a quasi-regular Dirichlet form on , and let be another (not necessarily quasi-regular) Dirichlet form on such that . Then there exists a locally compact separable metric space and a measurable map such that, by defining , is a unitary map from onto , and the following hold:
-
(1)
The image Dirichlet form is a regular Dirichlet form on , and serves as a quasi-homeomorphism that maps to the quasi-regular image Dirichlet form on .
-
(2)
There exists a Hunt process on and an MF with such that is properly associated with , and the subprocess is properly associated with .
6. Sandwiched Dirichlet form for domination
We consider two Dirichlet forms and on such that . The goal of this section is to explore the properties of a third Dirichlet form on that lies between and . Specifically, satisfies the sandwiching property:
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 . However, to facilitate exposition, we will also assume the quasi-regularity of and . 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 and are not quasi-regular. The following lemma illustrates this extension and demonstrates how quasi-regularity assumptions can be relaxed.
Lemma 6.1.
Let be a Dirichlet form on satisfying , where is assumed to be quasi-regular. Then, there exists a locally compact separable metric space and a measurable map such that, by defining , the operator is a unitary map from onto , and the following hold:
-
(1)
is a quasi-homoemorphism that maps to the quasi-regular image Dirichlet form on .
-
(2)
The image Dirichlet form is quasi-regular on .
-
(3)
The image Dirichlet form is regular on .
Particularly, the domination relationship
is preserved under the transformation .
Proof.
We begin by applying Corollary 5.2 to the Dirichlet forms and . This provides us with a locally compact separable metric space and a measurable map such that, by letting , is a unitary map from onto , acts as a quasi-homeomorphism that maps to the quasi-regular Dirichlet form , and is regular on . Since is onto , it follows that defines a Dirichlet form on . Additionally, the domination relationship
is preserved under the transformation . Next, we apply Corollary 5.2 to and . This yields another locally compact separable metric space and a measurable map such that, by letting , is a unitary map from onto , is a quasi-homeomorphism that maps to the quasi-regular image Dirichlet form , and the image Dirichlet form is regular on . Particularly, gives a Dirichlet form on .
Define , which is clearly Borel measurable from to . It is evident that , is a unitary map from onto , and
Since is a quasi-homeomorphism that maps to , it suffices to show that is a quasi-homoemorphism that maps to . To accomplish this, consider a -nest such that is a topological homeomorphism from to . Let be the -quasi-open subset of as specified in Theorem 5.1 for and . According to Remark 3.16, forms a -nest. Consider another -nest consisting of compact sets in , and define . Then is also a -nest, and is a topological homoemorphism. It can be verified that is a -nest by following the argument in (5.1). This completes the proof. ∎
From this point forward, we will further assume that both and are quasi-regular on . Let denote a Borel right process properly associated with . According to Corollary 4.6, there exists an MF with such that
Recall that is a bivariate smooth measure with respect to . Specifically, is smooth with respect to , and , where is the canonical measure of .
Theorem 6.2.
Let and be given as above. A quasi-regular Dirichlet form on is sandwiched between and , i.e.,
if and only if there exists an MF with the properties
such that .
Proof.
To demonstrate the sufficiency, we note that Corollary 3.17 indicate that . For convenience, let and . Consider . We will show that . Since , -q.e., and , it suffices to prove that
(6.1) |
In fact, it follows from and (3.10) that
Note that . Thus, we can easily derive (6.1). By utilizing , -q.e., and , it is straightforward to verify that forms an algebraic ideal in . The condition also implies that for all non-negative .
Conversely, let be a quasi-regular Dirichlet form such that . By applying Theorem 4.4 to and , we obtain an MF with such that . Let and . It follows from that . (Note that is -dense in .) Thus, , -q.e. It remains to show
(6.2) |
In fact, from Corollary 3.3, we know that is an -quasi-open set. Consider the part Dirichlet form of on :
Applying Theorems 2.4 and 4.4 to and , we find that is strongly subordinate to . Moreover, it follows from Proposition 3.11 and Theorem 4.2 that there exists a bivariate smooth measure on with respect to such that . Particularly, (6.2) is established. This completes the proof. ∎
Remark 6.3.
According to Proposition 3.11, the sandwiched Dirichlet form can also be expressed as
(6.3) | ||||
for some -quasi-open set such that , -q.e., and a bivariate smooth measure on with respect to such that .
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 is local. Then, the quasi-regular Dirichlet form is sandwiched between and if and only if there exists an -quasi-open set with , -q.e., and a smooth measure on with respect to with such that
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 is strongly subordinated to , which implies that . Then, the quasi-regular Dirichlet form is sandwiched between and if and only if there exists a positive measure on with such that
Proof.
It is sufficient to observe that a positive measure satisfying is a bivariate smooth measure, as is . ∎
6.2. Sandwiched by Silverstein extension
When is a Silverstein extension of , we have
The result in Theorem 6.2 indicates that the sandwiched Dirichlet form can be expressed as for some such that
Particularly, only charges (off diagonal) and (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 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
and . Let
where . Additionally, define
It is evident that is a regular, symmetric Dirichlet form on properly associated with the absorbing Brownian motion on , and is a regular, symmetric Dirichlet form on . The canonical measure corresponding to is
Since the -norm is equivalent to the -norm on , it follows that every sing-point set contained in is not -polar, all -quasi-continuous functions are continuous, and all -quasi-open sets are open. Moreover, is quasi-regular on , and is a Silverstein extension of .
Let be a quasi-regular (not necessarily symmetric) Dirichlet form on that is sandwiched between and . In its representation (6.3), the (-quasi-)open set has four possibilities:
The first case, where , is straightforward to address: We must have in (6.3). In other words, the sandwiched Dirichlet form is identical to . The second and third cases are analogous, so it suffices to discuss the second case, where . The bivariate smooth measure on satisfies
From these conditions, we conclude that only charges . Particularly,
for some constant .
It remains to consider the final case where . Similarly, only charges , with and . Specifically, we have
for , where and . Note that is not necessarily symmetric. It is symmetric if and only if (), in which case admits the Beurling-Deny decomposition
where .
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 and are symmetric, and is the active reflected Dirichlet space of 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 as and refer to as . It is important to note that is a Silverstein extension of ; see [9, Theorem 6.6.3].
Let be the MF corresponding to the domination . Then, is the part Dirichlet form of on . The set is precisely the “boundary” (of ), as considered in [18, Definition 4.2]. The following result demonstrates that the Dirichlet form sandwiched between and its active reflected Dirichlet space can be expressed as a perturbation of the part Dirichlet form of on some -quasi-open set (, -q.e.), thereby recovering the main result, Theorem 4.6, of [18].
Corollary 6.7.
The quasi-regular (not necessarily symmetric) Dirichlet form is sandwiched between and if and only if there exists an -quasi-open set with , -q.e., and a smooth measure with respect to with such that
where is the part Dirichlet form of on . Particularly, is symmetric.
Proof.
It suffices to prove the necessity. For convenience, we take every function to be its -quasi-continuous -version. According to, e.g., [9, Proposition 6.4.1], admits the Beurling-Deny decomposition as follows: For ,
where denotes the energy measure defined in [9, (4.3.8)]. For , we have
where
Thus, is the jumping measure of and serves as its killing measure. Utilizing [9, (4.3.34)], we further conclude that
(6.4) |
where denotes the energy measure of for the Dirichlet form . Let denote the local Dirichlet space of , as defined in [9, (4.3.31)]. The measure is well-defined for each using [9, Theorem 4.3.10]. According to the definition of the reflected Dirichlet space (see [9, Definition 6.4.4]), we have
where for ,
and for . We will prove that
(6.5) | ||||
is strongly local in the sense of [9, Proposition 6.4.1 (i)]. Once this is established, the uniqueness of Beurling-Deny decomposition for implies that
Note that is the canonical measure of . Consequently, the bivariate smooth measure in the expression (6.3) for reduces to a smooth measure such that , since .
It remains to demonstrate that (6.5) is strongly local. Let denote the -quasi-support of , and assume that is constant in an -quasi-open set with . Consider a sequence of increasing -quasi-open sets such that , -q.e., along with two sequences of functions and in such that and , -q.e. on for each . We aim to prove that
(6.6) |
Corollary 3.3 indicates that is -quasi-open. Given that , it follows from (6.4) and [9, Proposition 4.3.1 (ii)] that
Thus, we have
Since is constant on (), we can similarly obtain that , which further implies . Therefore, (6.6) is established. This completes the proof. ∎
7. Application to the Laplacian with Robin boundary conditions
Let (with ) be a non-empty open set with the boundary . Consider the Sobolev space
with the norm
where represents the distributional derivative. For , define
Moreover, we let
where denotes the space of all continuous real-valued functions on , and
where denotes the space of all infinitely differential functions on with compact support.
It is well known that is a regular, symmetric Dirichlet form on , while is a regular, symmetric Dirichlet form on ; see, e.g., [3, §2]. Denote by and the Hunt processes on and corresponding to these forms, respectively. For convenience, let Cap be the -capacity of , and if the form symbol preceding the quasi-notion is omitted, it defaults to the quasi-notion corresponding to . For instance, denotes the quasi-continuous version of with respect to , and the statement q.e. on asserts that is identically equal to on outside some polar set with respect to .
7.1. Local Robin boundary
In this subsection, we aim to investigate the (not necessarily symmetric) Dirichlet form on () that is sandwiched between and , i.e.,
(7.1) |
To avoid trivial cases, we always assume that , meaning that is not polar (see, e.g., [3, Proposition 2.5]).
Definition 7.1.
A positive Borel measure on is called quasi-admissible if it satisfies the following conditions:
-
(1)
is quasi-closed.
-
(2)
, regarded as a measure on with , is smooth with respect to the part Dirichlet form of on .
Remark 7.2.
It is important to emphasize that the defining set of plays a crucial role in determining admissibility. For two admissible measures, even if their trivial extensions on 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 on , we define a quadratic form on as follows:
(7.2) | ||||
These forms encapsulate all Dirichlet forms that are sandwiched between and , as demonstrated in the following result. It should be emphasized that the quasi-regularity of is not essential to this characterization, as we explained in Lemma 6.1.
Theorem 7.3.
A quasi-regular (not necessarily symmetric) Dirichlet form on is sandwiched between and if and only if there exists a quasi-admissible measure such that .
Proof.
If 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 is automatically a Dirichlet form. Particularly, we have the following.
Corollary 7.4.
Let be a coercive closed form on that satisfies (7.1). Then is a local and symmetric Dirichlet form.
Proof.
It is sufficient to note that the transformation 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 -semigroup 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 on , let
where . In other words, is the part of on which is locally finite. The measure is called admissible if implies for any Borel set . The main result, Theorem 4.1, of [4] states that if is admissible, then
(7.3) | ||||
is a Dirichlet form that satisfies (7.1).
In fact, is clearly an open subset of , and is a Radon measure on . If is admissible, then is a Radon smooth measure with respect to the part Dirichlet form of on . According to [9, Theorems 3.3.9 and 5.1.6], is a regular Dirichlet form on . Particularly, is closed and, consequently, quasi-closed. Therefore, we conclude that , when restricted to , is quasi-admissible.
Remark 7.7.
Consider the case , where each single-point subset of is a polar set. If is a smooth measure with , it is always possible to construct another smooth measure , which is equivalent to , such that is nowhere Radon on in the sense that for any non-empty open subset of (see [20, IV, Theorem 4.7]). Note that with is quasi-admissible, and , defined as (7.2), is a quasi-regular Dirichlet form on , which differs from . However, the admissibility argument in this example leads to because the part of , on which 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 that does not charge any polar sets.
Let be a positive measure on , which can be extended to in the way that . Assume that charges no polar sets, i.e., implies for any subset . Under these conditions, the integral () is well defined for any quasi-continuous function . Consider a sequence , which is dense in with respect to the norm . We define the set
Note that , q.e., and is independent of the choice of the sequence , as can be demonstrated by the following fact.
Lemma 7.8.
Let be a positive measure on such that implies for any subset . Then is a quasi-closed set such that , q.e. on for any . Furthermore, the restriction of to is quasi-admissible.
Proof.
For any , there exists a subsequence such that , q.e. Thus, , q.e. on . For the second part, it suffices to show that , as a measure on , is smooth with respect to the part Dirichlet form of on . 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 be a positive Borel measure on , and let be a positive, symmetric measure on , where denotes the diagonal of . The pair is called quasi-admissible if the following conditions hold:
-
(1)
is quasi-closed.
-
(2)
is smooth with respect to the part Dirichlet form of on , where .
Remark 7.10.
For any quasi-admissible pair , we define the symmetric quadratic form on as follows:
(7.4) | ||||
The following fact is elementary.
Lemma 7.11.
Let be a quasi-admissible pair. Then is a quasi-regular and symmetric Dirichlet form on .
Proof.
Since , it follows that is dense in . Clearly, is a symmetric, positive form that satisfies the Markovian property. To show that it is closed, consider an -Cauchy sequence . This sequence is also Cauchy in both and , and moreover, is Cauchy in . Hence, there exist , , and such that converges to and with respect to the and -norms, respectively, and with respect to the -norm. By taking a subsequence if necessary, we can assume that , q.e., (which indicates that q.e. on ,) , -a.e., and , -a.e. Since charges no polar sets, we obtain that , -a.e., and for -a.e. . Particularly, and . This establishes the closedness of .
It remains to demonstrate that is quasi-regular on . Let , which is quasi-admissible in the sense of Definition 7.1 with , and define as in (7.2). Notably, we have
This implies that any -nest is also an -nest. Similarly, consider the part Dirichlet form of on . From
it follows that any -nest is also an -nest. According to [9, Theorem 5.1.4], any -nest is an -nest. Thus, the quasi-notions for and are equivalent. Therefore, the quasi-regularity of follows directly from the quasi-regularity of and . 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 on and a symmetric, positive Borel measure on . Let be the part of on which and are locally finite, i.e.,
(7.5) |
where . The pair is called admissible if for any polar set . For such a pair, the key findings from [25, Theorems 3.2 and 3.3] indicate that
is closable, and its closure is given by
In fact, is clearly an open subset of , and is a Radon measure on . Thus, is a Radon smooth measure on with respect to the part Dirichlet form of on . It is noteworthy that , where represents the family of all continuous functions on that vanish at infinity (see [14, page 132]), and is dense in with respect to the uniform norm by the Stone-Weierstrass theorem. Particularly, is a regular, symmetric Dirichlet form on , which admits the Beurling-Deny decomposition for as follows:
where .
It is straightforward to verify that and comprise a quasi-admissible pair with , and , as defined in (7.4), is identical to .
Clearly, is dominated by , while is not dominated by due to Corollary 7.4. The following result characterizes all symmetric Dirichlet forms that are sandwiched between and .
Theorem 7.13.
Let be a quasi-admissible pair. A quasi-regular and symmetric Dirichlet form on is sandwiched between and if and only if there exists another quasi-admissible pair with the properties
(7.6) | ||||
such that .
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 -quasi-open set with along with a (symmetric) bivariate smooth measure on with respect to the part Dirichlet form of on such that and
Recall that the bivariate smooth measure satisfies that is smooth with respect to and . Consequently, , which implies .
Let us verify that is quasi-admissible. Note that the quasi-notion of corresponds to that of the part Dirichlet form of on . As stated in Corollary 3.3, is quasi-open, and hence, is quasi-closed. Since is smooth with respect to the part Dirichlet form of on , it follows from the remarks preceding Corollary 3.2 that the restriction of to is smooth with respect to the part Dirichlet form of on . It is notable that is also smooth with respect to the part Dirichlet form of on , as it is smooth with respect to . Therefore, we deduce that is smooth with respect to the part Dirichlet form of on , where . This verifies the quasi-admissibility of .
It remains to demonstrate that . Note that
For , we have
since . This establishes that . The opposite inclusion can be verified similarly. This completes the proof. ∎
One may be interested in local Dirichlet forms sandwiched between and . According to Theorem 7.13, these forms are characterized as follows.
Corollary 7.14.
Let be a quasi-admissible pair with . A quasi-regular and local Dirichlet form on is sandwiched between and if and only if there exists a quasi-admissible measure with the properties
such that , as defined in (7.2).
Proof.
It suffices to note that the Dirichlet form in Theorem 7.13 is local if and only if . ∎
7.3. Resolution of an open problem
We close this section with a special consideration of the admissible pair in Example 7.12. Recall that is defined as (7.5), , and the regular Dirichlet form corresponds to the quasi-admissible pair of measures
with the defining set .
Let , which is an admissible measure with as discussed in Example 7.6. Let be defined as (7.3) (with ). Since and , where , it follows from Corollary 7.14 that is sandwiched between and . The following result establishes that it is the largest among the local Dirichlet forms sandwiched between and .
Corollary 7.15.
If is a quasi-regular and local Dirichlet form on that is sandwiched between and , then is dominated by .
Proof.
Particularly, if is an admissible measure such that is sandwiched between and , then is dominated by . 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 be a Hausdorff topological space with the Borel -algebra assumed to be generated by the continuous functions on , and let be a -finite measure on with support . Denote by be the universal completion of .
A.1. Quasi-regularity
Let be a Dirichlet form on . An increasing sequence of closed subsets of is called an -nest if is -dense in , where . A set is called -polar if there exists an -nest such that . A statement is said to hold in the sense of -quasi-everywhere (abbreviated as -q.e.) if it holds outside an -polar set. A function on is called -quasi-continuous if there exists an -nest such that is finite and continuous on for each , denoted by . If , -a.e., and is -quasi-continuous, then is referred to as an -quasi-continuous -version of . A set is termed -quasi-open if there exists an -nest such that is an open subset of with respect to the relative topology for each . The complement of an -quasi-open set is called -quasi-closed. It is important to emphasize that these concepts are related solely to the symmetric part of .
Definition A.1.
A Dirichlet form on is called quasi-regular if the following conditions hold:
-
(1)
There exists an -nest consisting of compact sets.
-
(2)
There exists an -dense subset of whose elements have -quasi-continuous -versions.
-
(3)
There exists having -quasi-continuous -versions and an -polar set such that separates the points of .
It is called regular if is a locally compact separable metric space, is a fully supported Radon measure on , and is -dense in as well as uniformly dense in . It is noteworthy that a regular Dirichlet form is always quasi-regular.
It is well known that there exists an -tight right process
on properly associated with a quasi-regular Dirichlet form in the sense that is an -quasi-continuous -version of for any and , where is the transition semigroup of and is the -semigroup corresponding to (see [20, IV, Theorem 3.5]). Denote by the resolvent of , defined as
for all and . The notation related to general Markov processes can be found in, e.g., [7] and [30]. The symbol represents the cemetery, and denotes the lifetime of . Note that the -tightness of means that there exists a sequence of increasing compact sets such that
(A.1) |
where and denotes the first hitting time of any nearly Borel measurable set with respect to .
Without loss of generality, we may further assume that is a Lusin space (see [20, IV, Remark 3.2 (iii)]), and that is Borel measurable in the sense that for all and (see, e.g., [12, Corollary 3.23]). In this context, is referred to as a Borel right process. As established in [36], exists in for , -a.s. (This fact will be utilized only in § 3.3.)
A set is called -polar (for ) if there exists a nearly Borel measurable set such that . A statement holds in the sense of q.e. if it holds outside an -polar set. A nearly Borel measurable set is said to be (-)invariant if for every , . The restriction of a Borel right process 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 is called an -inessential set (for ) if and is -invariant. Obviously, an -inessential set is -polar. Note that any -polar set is contained in an -inessential Borel set (see, e.g., [9, Theorem A.2.15]). A numerical function defined q.e. on is called finely continuous q.e. if there exists an -inessential set such that is nearly Borel measurable and finely continuous with respect to the restricted right process . A set is called q.e. finely open if there exists an -inessential set such that is a nearly Borel measurable and finely open set for .
The following lemma summarizes the relationships between several -quasi-notions and concepts related to .
Lemma A.2.
Let be a Borel right process properly associated with the quasi-regular Dirichlet form on . Then the following statements hold:
-
(1)
A set is -polar, if and only if it is -polar.
-
(2)
An increasing sequence of closed subsets of is an -nest if and only if
(A.2) for q.e. (equivalently, for -a.e. ).
-
(3)
If is -quasi-continuous, then is finely continuous q.e. Conversely, if is finely continuous q.e., then is -quasi-continuous.
-
(4)
A set is -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 in this proof to an -polar set such that (A.2) holds for .)
The necessity of the fourth statement can be concluded by repeating the proof of [9, Theorem 3.3.3]. In particular, if is -quasi-open, there exists an -inessential Borel set such that is finely open with respect to . Conversely, let be q.e. finely open, and let be an -inessential set for such that is nearly Borel measurable and finely open with respect to . Take that is strictly positive on , and define
Note that is -excessive with respect to (see [9, Lemma A.2.4 (ii)]), and . It follows from, e.g., [19, Theorem 2.6] that . Thus, the third statement indicates that both and are -quasi-continuous. It is evident that , where and is the set of all regular points for . Note that is -polar, and hence -polar. Since is -quasi-open, we can conclude that both and are -quasi-open. This completes the proof. ∎
Remark A.3.
The third statement indicates the following fact: if is finely continuous q.e., then there exists an -inessential Borel set such that is not only nearly Borel measurable with respect to but also Borel measurable on .
A.2. Quasi-homeomorphism
Let be a second topological space with the Borel measurable -algebra , and let be a measurable map. Define , the image measure of under . Then
is an isometry. If is onto, i.e., , then
is a Dirichlet form on , referred to as the image Dirichlet form of under . We denote as .
Definition A.4.
Let be another Dirichlet form on with . The Dirichlet form is called quasi-homeomorphic to if there exists an -nest , an -nest and a map such that
-
(1)
is a topological homeomorphism from to for each .
-
(2)
.
-
(3)
, the image Dirichlet form of under .
Such a map is called a quasi-homeomorphism from to .
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 . 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.