Andrea N. Ceretani
Andrea N. Ceretani. Department of Mathematics of the Faculty of Exact and Natural Sciences, University of Buenos Aires, and Mathematics Research Institute “Luis A. Santaló” (IMAS), CONICET, Argentina.
aceretani@dm.uba.ar and Carlos N. Rautenberg
Carlos N. Rautenberg. Department of Mathematical Sciences and the Center for Mathematics and Artificial Intelligence (CMAI), George Mason University, Fairfax, VA 22030, USA.
crautenb@gmu.edu
Abstract.
We introduce a definition of the fractional Laplacian with spatially variable order and study the solvability of the associated Poisson problem on a bounded domain . The initial motivation arises from the extension results of Caffarelli and Silvestre, and Stinga and Torrea; however the analytical tools and approaches developed here are new. For instance, in some cases we allow the variable order to attain the values and leading to a framework on weighted Sobolev spaces with non-Muckenhoupt weights.
Initially, and under minimal assumptions, the operator is identified as the Lagrange multiplier corresponding to an optimization problem; and its domain is determined as a quotient space of weighted Sobolev spaces. The well-posedness of the associated Poisson problem is then obtained for data in the dual of this quotient space. Subsequently, two trace regularity results are established, allowing to partially characterize functions in the aforementioned quotient space whenever a Poincaré type inequality is available. Precise examples are provided where such inequality holds, and in this case the domain of the operator is identified with a subset of a weighted Sobolev space with spatially variant smoothness . The latter further allows to prove the well-posedness of the Poisson problem assuming functional regularity of the data.
Key words and phrases:
fractional order Sobolev space, spatially varying exponent, trace theorem, fractional
Laplacian with variable exponent, Hardy-type inequalities
2010 Mathematics Subject Classification:
35S15, 26A33, 65R20
CNR was partially supported by NSF grant 2012391. CNR was also been supported via the framework of MATHEON by the Einstein Foundation Berlin within the ECMath project SE15/SE19 and of the Excellence Cluster Math+ Berlin within project AA4-3 and the transition project OT6, and acknowledges the support of the DFG through the DFG-SPP 1962: Priority Programme “Non-smooth and Complementarity-based Distributed Parameter Systems: Simulation and Hierarchical Optimization” within Project 11. ANC was partially supported by Air Force Office of Scientific Research under Award NO: FA9550-19-1-0036, by CONICET grant PIP 0032CO, and by ANPCyT grant PICT 2016-1022.
The authors are very grateful to Prof. Harbir Antil from George Mason University for the valuable discussions on the paper.
1. Introduction
The goal of this work is twofold: (i) introduce the spectral fractional Laplacian associated with a homogeneous Dirichlet condition on a bounded domain , , in the case the fractional order is spatially variable and possibly attains the values and ; (ii) study the well-posedness of the equation
(1.1)
for some classes of data , and where is understood in an appropriate sense.
Motivated by the extension approach in by Caffarelli and Silvestre [5], or in bounded domains by Stinga and Torrea [17], we define to be the Lagrange multiplier associated to a suitable variational problem defined in an extended domain, for measurable functions with range contained in the interval . For a general class of functions , the domain of can be identified with a quotient space involving weighted Sobolev spaces,
(1.2)
where is the open semi-infinite cylinder (the extended domain) with base , and is a specific weight function. Roughly speaking, the spaces and are composed of functions that vanish on the lateral boundary of , and on the whole boundary (including the base ), respectively. Equation (1.1) is then solvable for every in the dual space of . For a smaller class of , the domain can be identified as a subset of a weighted Lebesgue space for some function , and the equation (1.1) is solvable when the right hand side is in . For an even smaller class of functions , this result is further improved since the domain of is identified with a subset of a weighted Sobolev space of functions with spatially variable smoothness, related to .
The main application that has motivated this work, in addition to the natural theoretical interest, is the recent paper [2]. There, initial results on an extension approach in Hilbert spaces on an open cylinder with base are given. However, the authors stopped short of defining due to the lack of a proper functional framework. The current paper aims to fill this gap. It is worth mentioning that none of the existing results in the literature are applicable to our case and new PDE and variational analysis tools are needed to study the current situation. For example, the extension approaches in [5, 17] assume to be a constant and avoid the extreme cases of and . In this setting, the nonlocal problem in , where is the -power of the realization of in with zero Dirichlet boundary conditions, can be equivalently formulated as a local one on a Sobolev space with a Muckenhoupt weight. On the other hand, our is a function which is allowed to touch the extreme cases 0 and 1 and therefore, the associated weights do not fulfill the Muckenhoupt property [2, Proposition 1]. In particular, fundamental results of type “” or Poincaré inequalities are not known in our case, leading to a more complex functional analytic framework.
The literature concerning possible definitions of with non-constant is restricted to the stochastic processes and stochastic calculus approaches and considers always the unbounded case ; see the monograph [3] and the references therein. By means of the Lévy-Khintchine representation formula, and the Fourier transform, the operator is determined to be of Lévy type. However, strong additional assumptions on are required to show that the operator is associated to a Feller or a Markov process. To name a few, these include assuming that is Lipschitz continuous and satisfies for some ;
see
[3, Example 3.5.9]. Neither of these restrictions are present in this work.
The paper is further motivated by several applications. The extension approach with spatially varying has shown
remarkable potential in image denoising: A rough choice of performs better than an optimal selected regularization parameter in total variation approaches; see [2]. This is indeed a game changer, especially the variable approach can enable one to replace the nonlinear (and degenerate) Euler-Lagrange equations in case of total variation by a linear one in the case of the variable fractional.
The variable approach can be also applied in geophysics: Models governed by a fractional Helmholtz equation (with constant fractional order) have shown good qualitative agreement with available magnetoteluric data, see [20]. Given the spatially long-range correlated heterogeneity of the medium, nonlocal models with spatially varying fractional order appear as an atractive tool to further obtain quantitative agreement.
Outline. The notation and main assumptions we make, specially those for the variable exponent , are specified in Section 2. In Section 3 we provide a succinct idea of the approach that we follow to study the fractional Laplacian with spatially variable order, , which is motivated by well-known results for the usual spectral fractional Laplacian.
Our main results begin from Section 4, where we introduce a definition of on the quotient space . Also in this section we prove the existence and uniqueness of a solution to the associated Poisson problem (1.1) for every in the dual space of . It is worth mentioning that the results in Section 4 require minimal conditions on the function , the weight , and the domain . The results given in Section 4, however, do not provide conditions for solvability of the Poisson problem when the right hand side of the elliptic equation is a (regular) real valued function defined only on .
In a second approach, we are able to better identify the domain of as a quotient space also, now on a Sobolev space that consist of functions in that formally vanish on the lateral boundary of . Differently from the construction given in Section 4, this second approach requires some extra conditions on both, and . These conditions are intimately related with the existence of -trace results for functions in , as well as with the existence of a Poincaré inequality in ; thus, we postpone the second construction until Section 7.
In Section 5, we first study the -traces of functions in , for . In particular, we are able to characterize -dependent integrability and differential regularity of restrictions of functions in to . Subsequently, we are able to prove the existence of a Poincaré inequality for in Section 6, for a special class of non-constant functions.
Our results finish in Section 7, where the details on the second definition of are given. Here, we identify the domain of with a subset of a weighted Lebesgue space for some weight , provided vanishes only on a set of zero measure and a Poincaré inequality holds for functions in . Further, we improve this result for the case when is the -dimensional unit square and satisfies some extra conditions. In this latter case we identify the domain of with a subset of a Sobolev space of functions with variable smoothness on . The paper closes with Section 8 that includes, in addition to conclusions, a number of open questions and future research directions.
More general elliptic operators of the form
with spatially variable fractional order , can be defined by extending the ideas in this paper in a natural way.
2. Notation and main assumptions
We assume that , , is a non-empty bounded open set with a Lipschitz boundary (except in Section 4, where no condition is imposed on the boundary). We denote by the open semi-infinite cylinder with base , by the lateral boundary of , and by the cylinder with the base , that is,
A generic point in is denoted by , where and .
A function is said to be a weight if is positive and finite almost everywhere. For an open set , and a weight , we denote by the space of measurable functions such
The space endowed with the norm is a Banach space. Further, given we say that a weight satisfies the condition, and write , if is locally integrable, that is,
For a weight , we define the weighted Sobolev space as the subset of of functions with weak gradients such that . Endowed with the norm
is a Banach space; see [12]. Notice that is a larger class of weights than the Muckenhoupt weights . The latter is also used to define weighted Sobolev spaces; see [19]. Throughout the paper we assume and denote the (Hölder) conjugate exponent of by .
The measurable function , which will characterize the spatially variable order of the fractional Laplacian, is assumed to satisfy:
(H1)
for almost all .
We use the notation to emphasize the dependence of the function on the spatial variable , and use to denote a constant in the interval .
Throughout the paper we consider the function defined by
and such that for a given , and , the function satisfies that
(H2)
, and if constant, then
for all . Here is the standard Euler-Gamma function.
Assumptions (H1) and (H2) imply that . However, it is known that (in general) is not expected to be of Muckenhoupt type, see [2, Proposition 1].
Given , we denote by the truncated cylinder of height , that is,
and define the sets and accordingly. The restriction of the weight to is also denoted by .
Example 2.1.
A possible choice for the function is given by
where ,
is a closed subset with zero-measure of and and .
This type of functions are useful in image processing where the set is the approximated set of edges/discontinuities of a certain image that one tries
to recover; see [2].
The two examples for that are of relevance to us are defined by
(2.1)
where . It follows that (H2) is satisfied given that .
3. The extended domain approach
This section is devoted to briefly review the well-known extension domain approach to define the spectral fractional Laplacian, see for instance [5, 17, 8].
Throughout this section, we assume that is constant.
We denote by the sequence of eigenvalues of the Laplace operator supplemented with a Dirichlet boundary condition, and consider an orthonormal basis of of associated eigenfunctions. The spectral fractional Laplacian is defined by
(3.1)
on the space
For extensions of (3.1) to non-homogeneous boundary conditions, we refer to [1].
It is worth mentioning that if
and for .
Here, is the closure in of the space of infinitely continuous differentiable functions with compact support in , and is the Lions-Magenes space [18].
Moreover, is the fractional Sobolev space of order ,
endowed with the norm
The extension approach introduced by Caffarelli and Silvestre [6], see [17, 7] for the case of bounded domains, establishes that if (dual space of ) then the unique solution to the elliptic equation
is given by , where satisfies
(3.2)
see [7, Lemma 2.2]. Here, denotes the dual pairing between and . Moreover, is the -trace operator for functions in the space
More precisely,
is the unique bounded linear operator that satisfies for every that vanishes on ; which is also onto over , that is
admits a unique solution for any , the harmonic extension operator
where is the solution to problem (3.4), is well-defined, linear, and bounded. Then one finds that the spectral fractional Laplacian given by (3.1) satisfies
(3.5)
for all and all , which provides an equivalent definition for . This second approach is our starting point to study the fractional Laplacian with spatially variable order: We identify a space of traces on which we can define the fractional Laplacian by a formula analogous to (3.5).
4. Abstract definition and solution to
We consider in this section an abstract derivation of the spatially variable fractional Laplacian . The advantage of this initial approach is that it requires minimal assumptions, namely (H1) and (H2), which are primarily sufficient conditions to have ; this leads to an appropriate definition of the associated weighted Sobolev spaces. Also, it is worth noticing that the arguments in this section do not require any assumption on the regularity of the boundary .
This path starts with the proper derivation of the trace space for the weighted Sobolev spaces in study. For this matter,
we consider the space
and endow it with the semi-norm
Note that is a norm on the subset of functions in that vanish at or . Subsequently, we define and as the completion in of the infinitely differentiable functions in with compact support in and , respectively, that is:
where
(4.1)
The only portion of the boundary where functions in do not necessarily vanish is the cap.
A few words are in order concerning and . Note that and are both pre-Hilbert spaces when endowed with the inner product
It follows then that their completion, and , are Hilbert spaces; in particular for there exist Cauchy sequences and in such that
If there is no risk of confusion, and in order to simplify notation, occasionally we simply write
and analogously we treat .
Given that , then we observe that is a closed subspace of . Thus, we can define an abstract space of traces on of functions in as the quotient space
We then define
i.e., the abstract trace on of a function is identified with the equivalence class that contains . The space is then endowed with the usual norm
Note that
(4.2)
is a linear and bounded operator, and that is a Hilbert space, given that and are also Hilbert spaces. We denote its inner product as . Further notice that, by definition, . Unless it is not clear from the context, we denote the class simply by . The following result establishes the existence of the harmonic extension operator.
Theorem 4.1.
Let and . The minimization problem:
()
for
admits a unique solution that, as , converges strongly to the unique solution to
()
for
Proof.
The existence of a solution to () follows from arguments of the direct methods for calculus of variations: The functional is non-negative, coercive, and weakly lower semicontinuous; for the latter part note that is also weakly lower semicontinuous. Uniqueness follows from the strict convexity of .
Since , there exists such that . Thus, given that is a minimizer of ,
(4.3)
for every . Then, by basic theory for penalty functions (see [13, Lemma 1 in Chapter 10]) we have that
(4.4)
Thus, by (4.3) we have that the sequence is bounded in , so it admits a weakly convergent subsequence, say
(4.5)
Further, by (4.4) we observe that . Next we show that with being the minimizer to ().
By weak lower semicontinuity of and (4.4), we observe:
that is . The fact that is a minimizer to () follows by selecting an arbitrary such that , then the previous to last inequality above yield
i.e., is a minimizer. Further, by strict convexity, minimizers to () are unique, so that the entire sequence satisfies
(4.6)
and also . Using (4.4), this limit is equivalent to
Theorem 4.1 ensures the existence of the abstract weighted harmonic extension operator
where is the solution to ().
In addition, the map is linear and bounded: Linearity follows directly from the examination of the first order conditions. For boundedness, consider (4.3) with instead of , where solves () and , to obtain . Then, by taking the limit as we observe
Then, by considering the infimum over all , we obtain
The well-posedness of the map allows us to establish a definition for the fractional Laplacian with spatially variable order.
Definition 4.2.
Let be the dual space of . The operator
is determined as follows: for , then is defined by
(4.8)
Remark 4.3.
The relation of the above definition with the classical spectral fractional Laplacian (3.5) is straightforward in light of the abuse of notation disclosed at the beginning of the chapter; in which case we can write
Furthermore, by a formal integration-by-parts formula and using the fact that is weighted harmonic,
we obtain that is equal to the generalized Neumann trace of when restricted
to . Then, as for the cassical case with constant order , can be understood as a Dirichlet-to-Neumann map.
Remark 4.4.
In view of Theorem 4.1, the expression in (4.8) is equivalent to
where is the unique solution to ().
The operator is well-defined as we see next, and it can be seen as the Lagrange multiplier associated to the harmonic extension problem.
Proposition 4.5.
For each , there exists a unique such that
Proof.
Initially, note that is the solution to (). For convenience, we write the constraint in () as , where is defined by . Since the operator is linear and bounded, also it is , and hence . Thus, is linear, bounded, and surjective. Therefore, there exists a unique Lagrange multiplier such that
which proves the statement.
∎
In view of Remark 4.3, we can also interpret as the Neumann trace of the extension
onto .
Remark 4.6.
It follows that is a bounded linear operator given that is linear and bounded.
We are now able to determine existence of solutions to the Poisson problem with spatially variant Laplacian.
Theorem 4.7.
Let . The equation
(4.9)
admits a unique solution in that is given by , where solves
(4.10)
for
Proof.
Since is linear and bounded, we have that
is a linear functional over . Then, there exists a solution to the problem (4.10) and the solution is unique due to strict convexity of .
Note that via necessary and sufficient conditions of optimality for (4.10), the unique solution satisfies:
(4.11)
and then is identical to its harmonic extension, i.e., . To see the latter, we consider in (4.11) and observe that by density
(4.12)
where we have used the fact that the functions in vanish on .
Moreover, we also (trivially) have so that satisfies first order optimality conditions for () for .
Hence, by convexity (uniqueness) . Also, by definition of the operator and
(4.11), we have
To prove uniqueness, consider a solution to (4.9) with and notice that
Then, satisfies first order optimality conditions for
whose unique minimizer is the zero function. Then, by convexity, , so that and hence .
∎
Remark 4.8(Truncated cylinder ).
It is worth mentioning that exactly the same construction with replaced by the truncated cylinder , , leads to a definition of by means of an extension problem on , as well as to the existence and uniqueness of solution to the associated Poisson problem. We care about because it makes the problem tractable from an implementation point of view [2, 15].
A few words are in order concerning Theorem 4.7; although it provides a solvability result for the elliptic problem, it does not establish existence of solutions based on maps defined on . That is, we would like to address the question: Under what conditions on , does the equation admit a solution? This question is answered in Section 7 and it is intimately related to the following trace results.
5. Trace theorems
In this section we identify a trace operator that properly relates values of maps on a Sobolev space in to their values at . For this matter, in addition to (H1) and (H2), we assume that the measurable function satisfies:
(H3)
The set of points on which is zero has measure zero, i.e., where
We define to be the closure in of the infinitely differentiable functions in
with compact support in , that is,
where is given in (4.1). Then, formally speaking, is the set of functions in that vanish on . We now prove the regularity of restrictions of functions in on the boundary.
Theorem 5.1(Trace theorem).
Provided that (H1), (H2), and (H3) hold true, there exists a unique bounded linear operator
that satisfies for all , where the weight is defined by
The same statement is true if we replace by the space , for every .
Proof.
For the sake of brevity, we define so that
Let and be such that and . Initially, we write
(5.1)
where is the partial derivative of with respect to the coordinate.
Let . Multiplying (5.1) by and then integrating from to with respect to , we find:
where:
Now, we notice that and that
since and . Thus,
Multiplying the last expression by , we obtain:
(5.2)
Next, we shall estimate and . A direct use of the Hölder’s inequality yields:
We now estimate in several steps. With the change of variables in the inner integral of , we obtain:
(5.3)
By adding and substracting in the exponent of , we rewrite the r.h.s. of (5.3) as
where . Then, by the Hölder’s inequality, we find:
Applying the Hölder’s inequality on the integral with respect to , we obtain:
Finally, we have:
Using the above estimations for and in (5.2), and observing that , we obtain:
from which we have:
(5.4)
since .
Raising inequality (5.4) to the power and then integrating over , we find:
Therefore, and
where . Notice that is an arbitrary, but fixed, number in , so that in this case we can fix to depend only on . The operator is the unique bounded linear extension of the mapping to .
Let us finally see that the same trace result holds true when we replace by , where . If , it follows from (5.4) that
from which, exactly as before, we find
(5.5)
If, on the contrary, , then we select in (5.4) and obtain (5.5) in the same way. The trace operator is now obtained as before.
∎
Remark 5.2.
If is constant, then both and are also constants. Hence, and , so it follows from Theorem 5.1 that
This is in accordance to the classical case, see [14, Theorem 3.2].
If, additionally, , then we observe that and the trace operator given in [7] (see also Section 3) coincide for functions in . From this, we find that is just given by the restriction to of the map in [7]. However, a deeper result is true; see Theorem 7.3.
In Theorem 5.1, we have characterized the integrability of functions in the trace space of . We aim now to identify the “smoothness” of functions in this trace space. This is a more complicated task since we aim at determining a space with a spatially variable smoothness associated to the function .
For simplicity, from now on we assume that is the -dimensional unit square . The forthcoming analysis requires one final assumption on the functions and :
(H4)
For almost every and , it holds true that
where .
Assumption (H4) enables us to use a Hardy-type inequality (see Lemma 5.8 below) for two specially chosen weights, which is a key ingredient to prove the subsequent improvement of the trace result in Theorem 5.9.
Example 5.3.
Let , , and constant; see (2.1) in Example 2.1. Suppose that satisfies:
(5.6)
for some , , and . Notice that the only point where is allowed to be zero is .
For this particular setting, although , i.e., is not a Muckenhoupt weight (see [2]), we find that (H4) holds true as we see next.
To simplify the notation below, we write . Since for all , we have:
Therefore, by Tonelli’s Theorem, we have that belongs to , which in turn implies that
for almost all , by Fubini’s Theorem.
Minor changes in the above arguments yield the same conclusion for functions with a finite number of zeros and a local behavior as (5.6) around each of them.
Next, in Definition 5.4 we present a Sobolev space of functions where smoothness is spatially dependent and related to .
First, we introduce the required notation.
For , let be given by
where
and the notation for means that the th-coordinate of is replaced by , that is:
Definition 5.4.
The space is defined by
(5.7)
with the norm
(5.8)
where
and
In order to address that controls locally the differential regularity of elements in , consider the following. For , let be the fractional Sobolev space of order , that is,
equipped with the norm
If , we have . Then, note the following lemma that can be found in [14] (see also [11]).
Lemma 5.5.
Let . There exists a positive constant such that
for every that satisfies for all , where
We now can show the relation between and the classical Sobolev spaces.
Theorem 5.6.
If constant, then
Proof.
Let and consider . Since , we have:
Then, a direct calculation yields:
for all since is constant by assumption (H2).
Therefore,
where . In addition, we notice that since is constant. Now the conclusion follows from Lemma 5.5 with .
∎
Remark 5.7.
In light of the previous result, it seems that a more appropriate notation for would be . We avoid this for the sake of brevity.
The following lemma is a key tool for the improvement of the result in Theorem 5.1. The proof can be found in [16, Sect. 2.6].
Lemma 5.8(Weighted Hardy-type inequality).
Let be a weight function defined in the interval . If
then
(5.9)
for all absolutely continuous functions in that satisfy , where
and .
Now we are in shape to prove the improvement of Theorem 5.1.
Theorem 5.9(Improved trace theorem).
Provided that (H1) to (H4) hold true, there exists a unique bounded linear operator
that satisfies for all .
The same statement is true if we replace by the space , for every .
Proof.
For the sake of simplicity, we give the proof only for ; with the natural changes,
the proof adapts straightforward to the case .
Let . Initially, we write:
(5.10)
where:
where as in Definition 5.4. Next, we shall estimate and separately. For this, we introduce the auxiliary function given by
We have:
where and denote the partial derivative of with respect to the first and second coordinates, respectively.
Interchanging the order of integration in the first term of the right hand side of the above inequality, and introducing the
change of variable in the second one, we find:
(5.11)
where:
and
The function is absolutely continuous in and satisfies for almost all . Additionally, by definition we observe that
The operator is the unique bounded linear extension of the map to .
The proof when is replaced by where is identical.
∎
Remark 5.10(Surjectivity of trace operator).
Although the previous result represents an improvement on the -trace characterization for functions in , nothing can be said about the surjectivity of the trace operator for non-constant.
Remark 5.11.
If is constant, then it follows by Theorem 5.6 that . Hence, the trace result in Theorem 5.1 is again in accordance to the classical case, see [14, Theorem 2.8] (see also Remark 5.2). Moreover, if and , we observe that
(5.17)
since , so in this case we further partially recover the trace result in [7, Lemma 2.2] given that .
6. Cases where the Poincaré inequality holds
We address now in this section cases and conditions on not constant that are sufficient for the Poincaré inequality to hold true. Two results are given, one in the entire cylinder and one in the truncated cylinder; see Theorem 6.1 and Theorem 6.3 respectively. From now on until the end of the section, we assume that constant, see (2.1) in Example 2.1; and is given by
(6.1)
where for and is a finite collection of non-empty open subsets of that satisfies .
In other words, we assume that is a step function in with range contained in the interval . Our first example is given by the next theorem which basically states that the Poincaré inequality holds provided that all pieces of the partition of touch the boundary .
then there exists a positive constant that satisfies
(6.3)
for all .
Proof.
The proof is quite direct, thanks to the existence of a Poincaré inequality for functions in that vanish on a subset of non-zero measure of : Let and . For every , the function belongs to and vanishes on a portion with non-zero measure of , by (6.2). Then, by the Poincaré inequality, we have
(6.4)
where is a positive constant that depends only on and , and is the gradient of with respect to the first coordinates.
Multiplying (6.4) by , then integrating for , and finally adding up for , we obtain
Next we prove that the truncated domain allows a much more amenable result than the one in the complete cylinder . In particular, we prove that (6.1) is a sufficient condition for the Poincaré inequality to hold; the result is given in next Theorem 6.3. The proof requires the following auxiliary lemma, see [10, Theorem 5.2] for its proof.
Lemma 6.2(Classical Hardy inequality).
Let and let be a differentiable function almost everywhere in that satisfies . If
then
where .
We are now in a position to present the final result in this section.
Theorem 6.3.
Assume that constant and is given by (6.1). For every there exists a positive constant that satisfies
(6.5)
for all .
Proof.
Let and . Initially, we write:
(6.6)
where
We denote by a positive constant that may depend only on and the partition , whose numerical value may be different from one line to another.
Let . We define
and observe that
(6.7)
where
For each fixed , the function belongs to . Thus, by the Poincaré-Wirtinger’s inequality, we obtain:
From this, similarly as in the proof of Theorem 6.1, we find:
(6.8)
Let be the extension by zero of to . Notice that is differentiable almost everywhere in since for all , and, trivially, satisfies . Also, observe that
since and is bounded in .
Then, by the classical Hardy inequality in Lemma 6.2 with , we have:
(6.9)
We now observe that:
where is the partial derivative of with respect to the coordinate. Then, by the Hölder’s inequality on the inner integral, we have:
We are now in a position to give a new definition for the operator , and to solve the associated Poisson problem for right hand sides defined on . The arguments below are very similar to those developed in Section 4 but now we assume some extra conditions on the function and the domain , which enable a better characterization of the domain of . We present the ideas for the semi-infinite cylinder , but the same arguments are valid for a truncated one .
From now on, we assume that the functions and satisfy hypotheses (H1), (H2), and (H3). Further, we assume that the Poincaré inequality holds true, that is there exists such that
For example, this is satisfied under the assumptions of Theorem 6.1 (see Theorem 6.3 for the case of a truncated cylinder). In particular, this implies that
algebraically and topologically. We endow with the norm .
Under the hypotheses assumed, we have established in Theorem 5.1 an -trace operator
(7.1)
and proved it is bounded, linear, and such that for all . Note that this operator is not, however, surjective. Subsequently, consider
which is a closed subspace of . Hence, a space of abstract traces on of functions in can be defined as the quotient space
Remark 7.1.
Due to the absence of density results of the type “” for non-Muckenhoupt weights, we are not in a position to assure that the spaces and are actually the same.
Immediately from here, via the isomorphism theorems, we can argue that there is an isomorphism
(7.2)
Moreover, one can simply consider to be given by . However, in order to identify with a subset of functions defined on , we need further information related with the structure of the function space .
as , and observe that is surjective by definition. In this setting we identify the abstract -trace of with the equivalence class that contains . The space is then endowed with the usual quotient norm
As before, we have . Note that is a Hilbert space, given that and are also Hilbert spaces.
Identically as in Theorem 4.1, we argue the existence of the weighted harmonic extension operator
where is the solution to
for
The well-posedness of the map allows us to establish a definition for the fractional Laplacian with spatially variable order.
Definition 7.2.
Let be the dual space of . The operator
is determined as follows: for , then is defined by
Since Proposition 4.5 holds true with the usual changes, the operator is then well-defined and Theorem 4.7 is also proven mutatis mutandis: For a , the equation
(7.4)
admits a unique solution that is given by , where solves
(7.5)
for
Although this approach seems equivalent to the one in Section 4, in this setting we have a more detailed representation of the elements . In fact, within this approach, there exists an injection
which is linear and bounded. Linearity follows directly, and boundedness follows given that for arbitrary ,
where we have used the linearity of and that , and then
In order to see that is an injection, suppose that , then so that , and the class is the zero element of . This identification allows us to consider to be the identity, and identify the continuous embedding
For a schematic relationship between the trace operators , , the isomorphism and the embedding , see Figure 1.
An amenable consequence of this identification is given in Theorem 7.4, however in first place we address the reduction to case where , a constant, where we obtain that is recovered as the domain of .
Figure 1. Diagram relating the operators , the isomorphism , and the operator .
Theorem 7.3.
Let be constant and suppose that functions in satisfy a Poincaré inequality, then
and therefore,
Proof.
Given that is constant, we have that is constant, and hence . Additionally, by Remark 5.2, we have that . Then, there is only left to prove that for each there exists a sequence in convergent in the sense of to a , and such that . We divide the proof into four steps for the sake of clarity.
Step 1: Let be arbitrary. Since for or , and for , it follows that is dense in . Then, there exists a sequence in such that
as . We denote and to their spectral decomposition where as . Further, define by
where each satisfies the Bessel equation:
Since the Poincaré inequality is valid for functions in , by the construction of the proof in [7, Proposition 2.1], we have that , and
and thus
(7.6)
as . Note that since has compact support, the support of is uniformly away from .
Step 2: For and , we consider a smooth non-increasing function such that:
and notice that the function belongs to . By direct calculation we have that
(7.7)
as .
Step 3: For and , we consider the shifted cylinder
and the weighted space , where
Further, let defined by reflection as
and note that , i.e.,
for all squares .
Let be the usual mollifier operator, i.e.,
where belongs to , , and . Since and since is bounded and with Lipschitz boundary, it follows that for ,
for any ; see [9]. Given that the support of is uniformly away from , and that if , it follows that
(7.8)
as . Note in addition, that for sufficiently large , we have
Step 4: In view of (7.6), (7.7), and (7.8), by appropriately selecting a sequence , we observe
as , so that . In particular,
and ; the result is then proven.
∎
Next we can establish the well-posedness of the elliptic equation of interest.
Theorem 7.4.
Assume that (H1), (H2), and (H3) holds true, and functions in satisfy a Poincaré inequality. For every , the equation
(7.9)
admits a unique solution .
Proof.
The conclusion follows from the existence and uniqueness of solution to the same problem with right hand side in since we identify , so that by means of .
∎
The result above can be refined in terms of regularity if in addition we observe (H4), and consider . In this case, the injection is given as
leading to our last theorem, whose proof is obtained as for Theorem 7.4.
Theorem 7.5.
In addition to the hypotheses of Theorem 7.4, consider and assume that (H4) holds true. Then, for every
The problem in the truncated cylinder is treated identically, and Theorem 7.4 and Theorem 7.5 still hold true under the obvious changes.
8. Conclusions and open questions
This paper continues the program initiated in [2] and provides a rigorous definition
of the variable order fractional Laplacian. The proposed theoretical framework enables solutions to Poisson equation
on bounded Lipschitz domains . The techniques introduced in the paper are new and
none of the existing works applies to our setting. However, the existing setting, where is a constant, can be
recovered from our proofs as a special case.
The following are open questions and topics for future research:
•
The study of as regularizer in optimization problem, i.e.,
and the optimal selection of in a bilevel framework.
•
The extension to more general settings of the Poincaré inequality type result presented in
Section 6.
•
The surjectivity of the new trace operator is still open (cf. Remark 5.10).
•
We have introduced Sobolev spaces with -dependent weights for the extension
problem and -dependent differentiability for the space on . New approaches need
to be established to prove additional regularity of solutions to in these Sobolev spaces.
•
Extensions to parabolic, semilinear and obstacle type problems are of interest.
•
The authors in [2] proposed a numerical method for the truncated problem.
But the numerical analysis of this problem is completely open. Also, convergence of the
truncated solution to the full solution is of interest as well.
•
Optimal control problems with variable order PDEs as constraints.
References
[1]
Harbir Antil, Johannes Pfefferer, and Sergejs Rogovs.
Fractional operators with inhomogeneous boundary conditions:
Analysis, control, and discretization.
Commun. Math. Sci., 16(5):1395–1426, 2018.
[2]
Harbir Antil and Carlos N. Rautenberg.
Sobolev spaces with non-Muckenhoupt weights, fractional elliptic
operators, and applications.
SIAM Journal on Mathematical Analysis, 51(3):2479–2503, 2019.
[3]
David Applebaum.
Lévy processes and stochastic calculus.
2009.
[4]
Haim Brezis.
Functional Analysis, Sobolev Spaces and Partial Differential
Equations.
Springer New York, 2011.
[5]
Luis Caffarelli and Luis Silvestre.
An extension problem related to the fractional Laplacian.
Communications in Partial Differential Equations,
32(7-9):1245–1260, 2007.
[6]
Luis A. Caffarelli, Sandro Salsa, and Luis Silvestre.
Regularity estimates for the solution and the free boundary of the
obstacle problem for the fractional Laplacian.
Inventiones Mathematicae, 171(2):425–461, 2007.
[7]
Antonio Capella, Juan Dávila, Louis Dupaigne, and Yannick Sire.
Regularity of radial extremal solutions for some non-local semilinear
equations.
Communications in Partial Differential Equations,
36(8):1353–1384, 2011.
[8]
Eleonora Di Nezza, Giampiero Palatucci, and Enrico Valdinoci.
Hitchhiker’s guide to the fractional Sobolev spaces.
Bull. Sci. Math., 136(5):521–573, 2012.
[9]
Vladimir Gol’dshtein and Alexander Ukhlov.
Weighted sobolev spaces and embedding theorems.
Transactions of the American Mathematical Society,
361(7):3829–3850, 2009.
[11]
Alois Kufner, Oldřich John, and Svatopluk Fučík.
Function spaces.
Academia, 1977.
[12]
Alois Kufner and Bohumír Opic.
How to define reasonably weighted sobolev spaces.
Commentationes Mathematicae Universitatis Carolinae,
25(3):537–554, 1984.
[13]
David G. Luenberger.
Optimization by Vector Space Methods.
John Wiley & Sons, Inc., New York, NY, USA, 1st edition, 1997.
[14]
Ales̆ Nekvinda.
Characterization of traces of the weighted Sobolev space
on .
Czechoslovak Mathematical Journal, 43(4):695–711, 1993.
[15]
Ricardo H. Nochetto, Enrique Otarola, and Abner J. Salgado.
A PDE approach to fractional diffusion in general domains: A priori
error analysis.
Found. Comput. Math., 15(3):733–791, 2015.
[16]
Bohumír Opic and Alois Kufner.
Hardy-type inequalities, volume 219.
Longman Scientific and Technical, 1990.
[17]
Pablo Raúl Stinga and José Luis Torrea.
Extension problem and Harnack’s inequality for some fractional
operators.
Communications in Partial Differential Equations,
35(11):2092–2122, 2010.
[18]
Luc Tartar.
An introduction to Sobolev spaces and interpolation spaces,
volume 3 of Lecture Notes of the Unione Matematica Italiana.
Springer, Berlin; UMI, Bologna, 2007.
[19]
Bengt O Turesson.
Nonlinear potential theory and weighted Sobolev spaces, volume
1736.
Springer Science & Business Media, 2000.
[20]
Chester Weiss, Bart van Bloemen Waanders, and Harbir Antil.
Fractional operators applied to geophysical electromagnetics.
Geophysical Journal International, 220(2):1242–1259, 2020.