Sub-Feller Semigroups Generated by Pseudodifferential Operators on Symmetric Spaces of Noncompact Type
Abstract
We consider global pseudodifferential operators on symmetric spaces of noncompact type, defined using spherical functions. The associated symbols have a natural probabilistic form that extend the notion of the characteristic exponent appearing in Gangolli’s Lévy–Khinchine formula to a function of two variables. The Hille–Yosida–Ray theorem is used to obtain conditions on such a symbol so that the corresponding pseudodifferential operator has an extension that generates a sub-Feller semigroup, generalising existing results for Euclidean space.
Keywords and Phrases. Riemannian symmetric space, Lie group, pseudodifferential operator, symbol, Feller process, Feller semigroup, generator, isotropic Sobolev space, spherical transform, fractional Laplacian.
MSC 2020. 43A85, 47D07, 47G20, 47G30, 60B15, 60G53
1 Introduction
Pseudodifferential operator theory is a powerful tool in the study of Feller–Markov processes on Euclidean space (see for example Knopova et al. (2015); Schilling (1998a, b, c); Schilling and Schnurr (2015), or Böttcher et al. (2013) §5 for a summary). Primarily developed by Niels Jacob and collaborators (see e.g. Jacob (1993, 1994); Hoh (1998)), this framework characterises sub-Feller semigroups and their generators as pseudodifferential operators (DOs) acting on , the Banach space of continuous, real-valued functions on that vanish at infinity. The associated symbols capture many properties of the sub-Feller processes and semigroups they represent, generalising the well-established relationship between Lévy processes and their characteristic exponents given by the Lévy–Khinchine formula. The key difference is that the Feller–Markov symbols typically depend on two variables instead of one — in the case of Feller processes, this is sometimes described as the Lévy characteristics having gained spatial dependence.
Manifold-valued Feller–Markov processes have also attracted interest in recent years (see Elworthy (1988); Hsu (2002) and Kunita (2019) §7 for excellent summaries), though the absence of a global harmonic analysis on general manifolds has so far prevented a DO approach. Lie groups and symmetric spaces come with their own harmonic analysis, however, in the form of the spherical transform (see Harish-Chandra (1958a, b) and Helgason (2001, 1984)). A natural question to ask then is to what extent a DO-based approach can be applied to the study of sub-Feller processes on Lie groups and symmetric spaces. Much work has already been done in this area, especially in the “constant coefficient” case of Lévy processes, in which the symbol depends only on its second argument. Here, the symbols are given by Gangolli’s Lévy–Khinchine formula (Gangolli (1964) Theorem 6.2), a direct analogue to the classical result. The first paper to use probabilistic DO methods on a Lie group was Applebaum and Cohen (2004), where DOs are used to study Lévy processes on the Heisenberg group. Pseudodifferential operator representations of semigroups and generators have also been found for Lévy processes on a general Lie group — see Applebaum (2011a, b) for the compact case and Applebaum (2013) Section 5 for the noncompact case. For Feller processes, DO representations have been found when the Lie group or symmetric space is compact — see Applebaum and Le Ngan (2020a, b).
This paper seeks to develop a more general theory of DOs for symmetric spaces of noncompact type, and apply it to seek conditions on a symbol so that the corresponding DO extends to the generator of some sub-Feller process. For the case, this question has been studied thoroughly in Jacob (1993, 1994), as well as Hoh (1998) Chapter 4.
The spherical transform enjoys many of the same properties as the Fourier transform on , and we find that several of the arguments in Jacob (1994) and Hoh (1998) generalise directly to the symmetric space setting. However, a direct transcription of Jacob and Hoh’s arguments is certainly not possible. One notable difference, for example, is the use of directional derivatives in the condition (2.2) of Jacob (1994), which on a manifold would depend on a choice of coordinate chart. Even on a symmetric space, there was no straightforward analogue for this condition, and a different approach is needed. We take a more operator-theoretic approach, and replace each of the partial derivatives () with the fractional Laplacian . This is an exciting object to work with, and we found it to be far more compatible with our harmonic analytical approach.
The structure of this paper will be the following. Section 2 presents a summary of necessary concepts and results from harmonic analysis on symmetric spaces, and introduces the system of symbols and DOs that will be used later on. We also introduce here the spherical anisotropic Sobolev spaces, a generalisation of the anisotropic Sobolev spaces first considered by Niels Jacob — see e.g. Jacob (1993, 1994).
In Section 3, we consider the Hille–Yosida–Ray theorem (see Theorem 3.1), and build on the work of Applebaum and Le Ngan (2020a, b), introducing a class of operators we will call Gangolli operators, which satisfy all but one of the conditions of Hille–Yosida–Ray. We prove that Gangolli operators are DOs in the sense of Section 2.5, and derive a formula for their symbols (Theorem 3.7).
Section 4 is concerned with seeking sufficient conditions for a Gangolli operator to extend to the generator a sub-Feller semigroup. Informed by the work of the previous section, this amounts to finding conditions that ensure
for some (see Theorem 3.1 (3)). This section perhaps most closely follows the approach of Jacob (1994) and Hoh (1998) Chapter 4, and where proofs are similar to these sources, we omit detail, and instead aim to emphasise what is different about the symmetric space setting. Full proofs may be found in my PhD thesis (Shewell Brockway, in preparation).
In Section 5 we present a large class of examples of symbols that satisfy the conditions found in Sections 3 and 4.
Notation. For a topological space , will denote the Borel -algebra associated with , and the space of bounded, Borel functions from , a Banach space with respect to the supremum norm. If is a locally compact Hausdorff space, then we write for the closed subspace of consisting of continuous functions vanishing at infinity, and for the dense subspace of compactly supported continuous functions. If is a smooth manifold and , then we write for the space of compactly supported -times continuously differentiable functions on .
2 Preliminaries
Let be a Riemannian symmetric space. By Theorem 3.3 of Helgason (2001), pp. 208, is diffeomorphic to a homogeneous space , where is a connected Lie group and is a compact subgroup of . Moreover, for some nontrivial involution on ,
where is the fixed point set of , and is the identity component of . Let and denote the Lie algebras of and , respectively. Note that is the eigenspace of the differential ; let denote the eigenspace. In fact, is a Cartan involution on , and the corresponding Cartan decomposition is
(2.1) |
Let denote the Killing form of , defined for each for all by . Assume that is of noncompact type, so that negative definite on and positive definite on . Since is nondegenerate, is semisimple.
Fix an -invariant inner product on , with respect to which (2.1) is an orthogonal direct sum. The Riemannian structure of is induced by the restriction of to .
There is a one to one correspondence between functions on and -right-invariant functions on , we denote both by . Similarly, we identify -invariant functions on with -bi-invariant functions on , and denote both by . Similar conventions will be used to denote standard subspaces of and ; for example will denote both the space of continuous, -invariant functions on , and the space of continuous, -bi-invariant functions on .
Equip with Haar measure, and for let denote the corresponding of real-valued functions. will denote the closed subspace of consisting of -bi-invariant elements. Since is unimodular, Haar measure is translation invariant, and can be projected onto the coset spaces and in a well-defined way. We continue to identify -bi-invariant functions on with functions on , as well as with -invariant functions on , in this setting.
2.1 Harmonic Analysis on Symmetric Spaces of Noncompact Type
For a thorough treatment of this topic, see Helgason (2001, 1984). Let denote the set of all left invariant differential operators on , and let denote the subspace of those operators that are also -right-invariant. A mapping is called a spherical if it is -bi-invariant, satisfies , and is a simultaneous eigenfunction of every element of .
Fix an Iwasawa decomposition , where is a nilpotent Lie subgroup of , and is Abelian. Let and denote respectively the Lie algebras of and . For each , let denote the unique element of such that . Harish-Chandra’s integral formula states that every spherical function on takes the form
(2.2) |
for some . Moreover, if an only if for some element of the Weyl group . A spherical function is positive definite if and only if .
The spherical transform of a function is the function given by
(2.3) |
Similarly, given a finite Borel measure on , the spherical transform of is the mapping given by
The spherical transform enjoys many useful properties, the most powerful being that it defines an isomorphism of the Banach convolution algebra with the space of Weyl group invariant elements of . The Borel measure is called Plancherel measure, and is given by
where denotes Harish-Chandra’s -function. According to the spherical inversion formula, for all and all ,
(2.4) |
There is also a version of Plancherel’s identity for the spherical transform, namely
(2.5) |
Let denote the subspace of consisting of -invariants. Then the image of under spherical transformation is a dense subspace of , and as such the spherical transform extends to an isometric isomorphism between the Hilbert spaces and . For more details, see for example Helgason (1984) Chapter IV § 7.3, pp. 454.
Similarly to classical Fourier theory, the most natural setting for the spherical transform is Schwarz space. A function is called rapidly decreasing if
(2.6) |
where denotes the geodesic distance on from to . Equivalently, if , where , then — see Gangolli and Varadarajan (1980) pp.167 for more details.
The (-bi-invariant) Schwarz space is the Fréchet space comprising of all rapidly decreasing, -bi-invariant functions , together with the family of seminorms given by the left-hand side of (2.6). By viewing the spaces and as finite dimensional vector spaces, we also consider the classical Schwartz spaces and , as well as -invariant subspaces and . The Euclidean Fourier transform
(2.7) |
defines a topological isomorphism between the spaces and in the usual way. Given and , the Abel transform is defined by
The Abel transform is fascinating in its own right, and we refer to Sawyer (2003) for more information. However, for our purposes we are mainly interested in its role in the following:
Theorem 2.1.
Writing for the spherical transform, the diagram
commutes, up to normalizing constants. Each arrow describes an isomorphism of Fréchet algebras.
2.2 Probability on Lie Groups and Symmetric spaces
We summarise a few key notions from probability theory on Lie groups and symmetric spaces. Sources for this material include Liao and Wang (2007) and Liao (2004, 2018).
Fix a probability space . Just as with functions on and , we may view stochastic processes on as projections of processes on whose laws are -right invariant. Let a stochastic process taking values on . The random variables
are called the increments of . Equipped with its natural filtration , is said to have independent increments if for all , is independent of , and stationary increments if
A process on is stochastically continuous if, for all and all with ,
A stochastically continuous process on with stationary and independent increments is called a Lévy process on . A process on is called a Lévy process if it is the projection of a Lévy process on , under the canonical surjection . Lévy processes on correspond precisely to the -invariant Feller processes on . The proof of this is similar to the well-known result for -valued Lévy processes.
The convolution product of two Borel measures on is defined for each by
(2.8) |
Note that since is semisimple, it is unimodular, and hence this operation is commutative. It is also clear from the definition that is -bi-invariant whenever and are.
Definition 2.2.
A family of finite Borel measures on will be called a convolution semigroup (of probability measures) if
-
1.
for all ,
-
2.
for all , and
-
3.
weakly as .
Note that must be an idempotent measure, in the sense that . By Theorem 1.2.10 on page 34 of Heyer (1977), must coincide with Haar measure on a compact subgroup of . We we will frequently take to be normalised Haar measure on , so that the image of after projecting onto is , the delta mass at .
One may also define convolution of measures on , and convolution semigroups on are defined analogously — see Liao (2018) Section 1.3 for more details. In fact, the projection map induces a bijection between the set of all convolution semigroups on and the set of all -bi-invariant convolution semigroups on — see Liao (2018) Propositions 1.9 and 1.12, pp. 11–13. We henceforth identify these two sets, but generally opt to perform calculations using objects defined on , for simplicity.
Let be a Lévy process on , and for each , let denote the law of . By Liao (2018) Theorem 1.7, pp. 8, is a convolution semigroup of probability measures on .
Definition 2.3.
We call the convolution semigroup associated with .
Let be a Lévy process on , and a Lévy process on for which . Let and denote the transition probabilities of and , respectively. Then for all , and ,
In particular, the prescription
defines a convolution semigroup on . By Liao (2018) Proposition 1.12, pp. 13, is -invariant, and there is a -bi-invariant convolution semigroup on for which
(2.9) |
It may be tempting to think that should be the convolution semigroup of . In fact, this is not the case: if it were, then we would have , which is not a -bi-invariant measure on . However, if we denote the convolution semigroup of by , and normalised Haar measure on by , then by Liao (2018) Theorem 3.14, pp. 88,
is a suitable choice for the -bi-invariant convolution semigroup on , for which (2.9) is satisfied. In particular, .
In this way, Lévy processes on may be understood through the study of -bi-invariant convolution semigroups on . The corresponding Lévy processes on are called -bi-invariant Lévy processes. For such a process , with -bi-invariant convolution semigroup , the Hunt semigroup of ) is given by
(2.10) |
Note that since , we have . In fact, forms a strongly continuous operator semigroup on , and the restriction of each to yields a strongly continuous semigroup on . is a left invariant Feller semigroup in each of these cases (see Ngan (2019) pp. 82–83).
Restricting to the -bi-invariant functions in this way will be advantageous, as we have the spherical transform at our disposal. As an early application of this, we prove the following useful eigenvalue relation for the Hunt semigroup of a -bi-invariant convolution semigroup.
Proposition 2.4.
Let be a -bi-invariant convolution semigroup on , and let denote the the restriction to of the Hunt semigroup associated with . Then for all , and ,
Proof.
Let , and . Observe that since each is invariant under all translations by ,
for each . Integrating over and applying a Fubini argument,
We can now apply the beautiful integral formula for spherical functions,
(2.11) |
(c.f Helgason (1984) pp. 400–402), to conclude
as desired. ∎
The infinitesimal generator of a Lévy process on is given by the celebrated Hunt formula (Hunt (1956) Theorem 5.1). We describe a version of this next, specialising to the -bi-invariant case most relevant to our work on symmetric spaces. We first introduce a local coordinate system on , defined in terms of the orthogonal decomposition (2.1).
Definition 2.5.
Let be a basis of , ordered so that is a basis of . A collection of smooth functions of compact support is called a system of exponential coordinate functions if there is a neighbourhood of for which
(2.12) |
The may be chosen so as to be -right-invariant for , and such that
For more details, see Liao (2018) pp.36–37, 83.
The choice of basis of enables us to view as a matrix, for each . A vector is said to be -invariant if
Similarly, a real-valued matrix is -invariant if
A Borel measure on is called a Lévy measure if , , and .
We state a useful corollary of the famous Hunt formula. For more details, including a proof, see Section 3.2 of Liao (2018), pp. 78.
Theorem 2.6.
Let be the infinitesimal generator associated with a -bi-invariant Lévy process on . Then , and there is an -invariant vector , an -invariant, non-negative definite, symmetric matrix , and a -bi-invariant Lévy measure such that
for all and . Moreover, the triple is completely determined by , and independent of the choice of exponential coordinate functions .
Conversely, given a triple of this kind, there is a unique -bi-invariant convolution semigroup of probability measures on with infinitesimal generator given by .
Since is semisimple, has no non-zero -invariant elements. This means that for the class of manifold we are considering, -bi-invariant Lévy generators will take the form
(2.13) |
Given such a Lévy generator, we write for the diffusion part of . By the discussion surrounding (3.3) in Liao (2018), pp. 75, , and so for each there is such that
(2.14) |
Moreover, is a -invariant quadratic polynomial function on .
Theorem 2.7 (Gangolli’s Lévy–Khinchine formula).
Let be a -bi-invariant convolution semigroup of probability measures on with infinitesimal generator , and let denote the diffusion part of . Then , where
(2.15) |
and is given by (2.14).
This result was first proven in Gangolli (1964), see also Liao and Wang (2007). For a proof of the specific statement above, see page 139 of Liao (2018).
The function given by (2.15) will be called the Gangolli exponent of the process .
Remark 2.8.
2.3 Positive and Negative Definite Functions
By viewing as a finite-dimensional real vector space, we may consider positive and negative definite functions on , defined in the usual way.
Proposition 2.9.
-
1.
For all , is positive definite.
-
2.
Let be a finite -bi-invariant Borel measure. Then is positive definite.
Proof.
Let , , , and , and note that
Therefore, by the Harish-Chandra integral formula (2.2),
(2.16) |
Part 1 follows.
For part 2, observe that since (2.16) holds for all , we can replace each by its complex conjugate. Therefore, for all , , , and . Taking complex conjugates,
for all , , , and , and hence
∎
By choosing a basis of , we may identify it with , and apply classical results about positive (resp. negative) definite functions on Euclidean space to functions on , to obtain results about positive (resp. negative) definite functions in this new setting.
One useful application of this is the Schoenberg correspondence, which states that a map is negative definite if and only if and is positive definite for all . This is immediate by the Schoenberg correspondence on — see Berg and Forst. (1975) page 41 for a proof.
Proposition 2.10.
Let be the Gangolli exponent of a Lévy process on . Then is negative definite.
Proof.
Let is a Lévy process on , and let be the law of , for all . Then forms a convolution semigroup on . By Proposition 1.12 of Liao (2018) (pp. 13), arises as the projection onto of a -bi-invariant convolution semigroup on . By Proposition 2.9, the spherical transform of each is positive definite, and by the Schoenberg correspondence, for each , there is a negative definite function on such that and . In fact, since is a convolution semigroup, it must be the case that
By uniqueness of Gangolli exponents, , a negative definite function. ∎
We finish this subsection with a collection of results about negative definite functions, which will be useful in later sections.
Proposition 2.11.
Let be a continuous negative definite function. Then
-
1.
For all ,
-
2.
(Generalised Peetre inequality) For all and ,
-
3.
There is a constant such that
(2.17)
Proof.
These results follow from their analogues on — see Hoh (1998) page 16. ∎
2.4 Spherical Anisotropic Sobolev Spaces
Suppose is a real-valued continuous negative definite function, and let . We define the (spherical) anisotropic Sobolev space associated with and to be
where denotes the space of -bi-invariant tempered distributions. One can check that each is a Hilbert space with respect to the inner product
These spaces are a generalisation of the anisotropic Sobolev spaces first introduced by Niels Jacob, see Jacob (1993), and developed further by Hoh, see Hoh (1998). For the special case , we will write . Note also that , by the Plancherel theorem. In this case, we will omit subscripts and just write for the inner product.
Note that is a non-negative function, since it is negative definite and real-valued. We impose an additional assumption, namely that there exist constants such that
(2.18) |
Analogous assumptions are made in Jacob (1994) (1.5) and Hoh (1998) (4.2), and the role of (2.18) will be very similar.
Theorem 2.12.
Let be a real-valued, continuous negative definite symbol, satisfying (2.18). Then
-
1.
and are dense in each , and we have continuous embeddings
-
2.
We have continuous embeddings
whenever with . In particular, for all .
-
3.
Under the standard identification of with its dual, the dual space of each is isomorphic to , with
(2.19) for all .
- 4.
-
5.
Let . Then for all , there is such that
(2.20) for all .
-
6.
There exist continuous embeddings
for all , where .
Lemma 2.13.
Let . Then .
Proof.
By standard arguments, one may check that , for all . Writing , we have , and hence whenever . By Proposition 7.2 on page 450 of Helgason (1984), there are such that
(2.23) |
Let be such that for all . Then
(2.24) |
whenever . ∎
Proof of Theorem 2.12.
Much of this theorem may be proved by adapting proofs from the case. For example, to prove Theorem 2.12 (1), let denote the space of all measurable functions on for which , a Hilbert space with respect to the inner product
By viewing as a real vector space and using inequality (2.23) to relate to Lebesgue measure, the proof of Theorem 3.10.3 on page 208 of Jacob (2001a) may be easily adapted to show that
is continuous. Noting Theorem 2.1, Theorem 2.12 (1) follows.
Proofs of Theorem 2.12 (2)–(5) are almost identical to their -based counterparts, see Jacob (1994) §1, or Hoh (1998) pp. 46–48.
By Theorem 2.12 (4), Theorem 2.12 (6) will follow if we can prove the existence of a continuous embeddings
(2.25) |
for all . Let and . By Lemma 2.13, , and by the spherical inversion formula (2.4),
for all . By the Cauchy–Schwarz inequality,
for all , where . It follows that
The embedding (2.25) may then be obtained using a density argument. ∎
2.5 Pseudodifferential Operators and Their Symbols
A measurable mapping will be called a negative definite symbol if it is locally bounded, and if for each , is negative definite and continuous. If in addition is continuous in its first argument, we will call a continuous negative definite symbol.
Let denote the set of all measurable functions on .
Theorem 2.14.
Let be a negative definite symbol, and for each and , define
(2.26) |
Then
-
1.
Equation (2.26) defines a linear operator .
-
2.
If is a continuous negative definite symbol, then .
-
3.
If is -bi-invariant in its first argument, then is -bi-invariant for all .
Proof.
Definition 2.15.
Operators of the form (2.26), where is a negative definite symbol, will be called (spherical) pseudodifferential operators on .
An important subclass of these operators first appeared for irreducible symmetric spaces in Applebaum (2013), with the symbol arising as the Gangolli exponent of a -bi-invariant Lévy process. Note that just as in the classical Euclidean case, the symbols arising from Lévy processes are spatially independent, in the sense that they are constant in their first argument. We explore some specific examples of this below. In Section 3, we introduce a large class of examples pseudodifferential operators with spatial dependence.
Example 2.16.
-
1.
Diffusion operators with constant coefficients. Since is semisimple, the generator of a -bi-invariant diffusion-type Lévy process on takes the form , where is an -invariant, non-negative definite symmetric matrix (c.f. (2.13)). As already noted, ; let denote the -eigenvalue of . Note that is the Gangolli exponent of .
We claim that is a continuous negative definite symbol, and the associated pseudodifferential operator is . To see this, let denote the convolution semigroup generated by , and let be the associated Hunt semigroup, as defined in (2.10). Then, given and ,
(2.27) By the spherical inversion formula (2.4), for all ,
Recalling that whenever , a Fubini argument may be applied to conclude that . By Proposition 2.4 and Theorem 2.7,
and so
By (2.27), for all and ,
(2.28) Now, if and , then
Moreover, , since , and is a -invariant polynomial function. By the dominated convergence theorem, we may bring the limit through the integral sign in (2.28) to conclude that
(2.29) for all and .
-
2.
Brownian motion. As a special case of the above, is a pseudodifferential operator with symbol .
-
3.
Killed diffusions. With minimal effort, the results of Example 2.16 (1) may be extended to include killing. To see this, note first that such operators are always of the form , where is a diffusion operator of the form considered above, and . The associated -eigenvalues must satisfy
and hence using (2.29) as well as the spherical inversion theorem,
for all and .
-
4.
Lévy generators. More generally, if is the infinitesimal generator of a -bi-invariant Lévy process on , and if is the corresponding Gangolli exponent, then is a continuous negative definite symbol, and is the corresponding pseudodifferential operator. This is proven in Applebaum (2013) Theorem 5.1 in the case where is irreducible, and later in this paper as a special case of Theorem 3.7.
3 Gangolli Operators and the Hille–Yosida–Ray Theorem
We will soon define the class of pseudodifferential operators that will be of primary interest. In this section, we motivate this definition with a short discussion of the Hille–Yosida–Ray theorem, and prove that our class of operators are pseudodifferential operators in the sense of Definition 2.15. We finish the section with some examples.
Let be a locally compact, Hausdorff space, let be a closed subspace of , and let denote the space of all real-valued functions on . A -semigroup defined on is called sub-Feller if for all , and all ,
A linear operator is said to satisfy the positive maximum principle, if for all and such that , we have .
The following theorem is an extended version of the Hille–Yosida–Ray theorem, which fully characterises the operators that extend to generators of sub-Feller semigroups on . Similar versions in which may found in Hoh (1998), pp. 53, and Jacob (2001a), pp. 333. For a proof, see Ethier and Kurtz (1986), pp. 165.
Theorem 3.1 (Hille–Yosida–Ray).
A linear operator on is closable and its closure generates a strongly continuous, sub-Feller semigroup on if and only if the following is satisfied:
-
1.
is dense in ,
-
2.
satisfies the positive maximum principle, and
-
3.
There exists such that is dense in .
In their papers Applebaum and Le Ngan (2020a, b), Applebaum and Ngan found necessary and sufficient conditions for an operator defined on to satisfy Theorem 3.1 (2), for the cases , and . We will focus primarily on the case , since this is the realm in which the spherical transform is available.
A mapping will be called a -bi-invariant Lévy kernel if it is -bi-invariant in its first argument, and if for all , is a -bi-invariant Lévy measure. Fix a system of exponential coordinate functions, as defined in Definition 2.5, and adopt all of the notation conventions from this definition.
Definition 3.2.
An operator will be called a Gangolli operator if there exist mappings (), as well as a -bi-invariant Lévy kernel , such that for all and ,
(3.1) | ||||
and if for all ,
-
1.
.
-
2.
is an -invariant, non-negative definite, symmetric matrix.
Remarks 3.3.
- 1.
- 2.
For a Gangolli operator given by (3.1), and for each , we will denote by the operator obtained by freezing the coefficients of at . Explicitly, for all and ,
For each , is the generator of a killed -bi-invariant Lévy process on . We continue to adopt the notation for the diffusion part, and for the -eigenvalue of .
Consider the following continuity conditions on the coefficients of :
-
(c1)
are continuous, for .
-
(c2)
For each , the mappings and are continuous from to .
Lemma 3.4.
Let be a Gangolli operator, and define by
(3.2) |
Suppose (c1) and (c2) hold. Then is a continuous negative definite symbol.
Proof.
That is continuous in its first argument is immediate from (c1) and (c2). Fix and consider . By Theorem 2.6, there is a convolution semigroup generated by , and by Theorem 2.7, the corresponding Gangolli exponent is a continuous negative definite mapping on , given by
Therefore is continuous, and negative definite since for fixed , is a non-negative constant. ∎
Definition 3.5.
The symbols described by Lemma 3.4 will be referred to as Gangolli symbols, due to their connection with Gangolli’s Lévy–Khinchine formula.
Remarks 3.6.
-
1.
Gangolli exponents are precisely those Gangolli symbols constant in their first argument.
-
2.
The set of all Gangolli symbols forms a convex cone.
Theorem 3.7.
Let and be as in Lemma 3.4. Then is a pseudodifferential operator with symbol .
Proof.
Let denote the non-local (i.e. jump) part of , so that
(3.3) |
for all and . By design,
(3.4) |
For the diffusion part of , note that for each , is an operator of the form considered in Example 2.16 (3), and in particular satisfies
(3.5) |
for all .
Consider now the jump part . By Lemma 2.3 on page 39 of Liao (2018), for each fixed , and for all , the integrand on the right-hand side of (3.3) is absolutely integrable with respect to . Therefore, (3.3) may be used to extend the domain of so as to include . We do so now, and (without any loss of precision) denote the extension by .
Let us proceed similarly to Applebaum and Le Ngan (2020b) Section 5, and define for each a linear functional by
Then , and hence
for all and . Moreover, the integrand on the right-hand side is absolutely -integrable, for all and . Since , and for all (Theorem 5.3 (b) of Liao (2018)),
Thus, for all and , is absolutely -integrable, and
(3.6) |
A standard argument involving the functional equation (2.11) for spherical functions may now be applied in precisely the same way as in Applebaum and Le Ngan (2020b) (5.3)–(5.7), to infer that
(3.7) |
for all and .
Finally, let , and observe that by the spherical inversion formula
(3.8) | ||||
Claim.
For all and ,
Proof of Claim.
This is a fairly standard differentiation-through-integration-sign argument. First note that by translation invariance of , it suffices to prove the claim for . Now,
The claim will follow if we can apply the dominated convergence theorem to bring the above limit through the integral sign. By the mean value theorem, for each and ,
for some , and hence for all . By Helgason Theorem 1.1 (iii), , for some some constant . Thus, for , and ,
and clearly , since . Hence we may apply dominated convergence as desired, and the claim follows.
Example 3.8.
-
1.
Let be non-negative, and let be a Gangolli exponent. Then given by
is a Gangolli symbol. Indeed, by Theorem 2.7, there exists a sub-diffusion operator and a -bi-invariant Lévy measure such that for all ,
and hence for all and ,
If , where and is an -invariant, non-negative definite symmetric matrix, then the characteristics are are
Since is non-negative, continuous and -bi-invariant, the conditions of Definition 3.2 are easily verified for these characteristics, as are (c1) and (c2).
-
2.
Hyperbolic plane. As described in Helgason (1984) (pp. 29–31), the Poincaré disc model of the hyperbolic plane is isomorphic to . Moreover, is a symmetric space of noncompact type, with spherical functions are given by the Legendre functions
(see Helgason (2001) Proposition 2.9, pp. 406). Since is irreducible and , by Theorem 3.3 of Applebaum and Le Ngan (2020b), diffusion operators on must be multiples of the Laplace–Beltrami operator, and the symbols of Feller processes take the simplified form
for all and . The constant coefficient (i.e. Lévy) case of this formula was discovered by Getoor — see Getoor (1961) Theorem 7.4.
4 Construction of Sub-Feller Semigroups
In this section we tackle the third condition of Hille–Yosida–Ray (Theorem 3.1), when . To this end, we seek conditions on a symbol so that, for some ,
(4.1) |
Our approach is based primarily on Jacob (1994) and Hoh (1998) Section 4. Now that we are on the level of operators, there are more arguments that closely resemble these sources. In these cases, proofs are not expanded in great detail, and may be omitted entirely to save space. Instead, we aim to emphasise what does not carry over from the Euclidean space setting.
For a mapping and for each , , define
(4.2) |
Observe that if for all , then , and we may consider the spherical transform , given by
To motivate the introduction of , consider the case , . In this case, the so-called frequency shift property for the Fourier transform says that
(4.3) | ||||
where ∧ denotes the Fourier transform taken in the first argument of . Hoh (1998) and Jacob (2001b) make use of bounds on , and will assume an analogous role in work to come.
As in previous work, let be a fixed real-valued, continuous negative definite function satisfying (2.18) for some fixed . The next lemma is an analogue of Lemma 2.1 of Jacob (1994). See also Hoh (1998) Lemma 4.2, pp. 48. The primary difference in this work is the presence of integer powers of , which replace the multinomial powers of of the setting.
One advantage of this approach is that () has a global definition that does not depend on our choice of local coordinates. Another advantage is that we know its symbol — see equations (4.8) and (4.9) below.
Lemma 4.1.
Let , and suppose for all . Suppose that for each , there is a non-negative function such that
(4.4) |
for all , . Then there is a constant such that
(4.5) |
for all , where denotes the usual norm on the Banach space .
Remarks 4.2.
-
1.
As in (2.21), .
- 2.
-
3.
Under the conditions of the lemma, and using the Fubini theorem, we have the following: for all and ,
(4.6) Fubini’s theorem does indeed apply here — a suitable bound for the integrand on the first line of (4.6) may be found by noting that, by (4.4),
(4.7) for all and . By Theorem 2.1, , and the usual bound (2.23) on the density of Plancherel measure may be applied, similarly to (2.24), to conclude that the right-hand side of (4.7) is -integrable.
Proof of Lemma 4.1.
Let and be fixed. The fractional Laplacian satisfies a well-known eigenrelation
(4.8) |
which may be proven using subordination methods and properties of the Laplace-Beltrami operator on a symmetric space, using similar techniques to Section 5.7 of Applebaum (2014), pp. 154–7. One can also show using standard methods that
(4.9) |
for all and . Then, using the definition of the spherical transform,
for all and all . Applying this to , we have for all ,
and summing over ,
(4.10) |
for all . Let be the smallest positive number such that
Then, rearranging (4.10),
(4.11) |
for all .
Remark 4.3.
The constant
(4.12) |
appearing in the proof of Lemma 4.1 will remain relevant throughout this chapter.
Let now be a continuous negative definite symbol, -bi-invariant in its first argument, and -invariant in its second (for example, could be taken to be a Gangolli symbol, as in (3.2)). Similarly to Jacob (1994) §4 and Hoh (1998) (4.26), we write
(4.13) |
where and , for some fixed . Observe that is necessarily a negative definite symbol. Though may not be, we may still define the operator in a meaningful way, by
By decomposing in this way, we view it as a perturbation of a negative definite function by . The assumptions we place on will control the size of this perturbation, as well as ensuring certain regularity properties of acting on the anisotropic Sobolev spaces introduced in Section 2.4.
Assumptions 4.4.
In the notation above, we impose the following:
-
1.
There exist constants such that for all with ,
(4.14) -
2.
Let , , and suppose that for all . Suppose further that for , there exists such that
(4.15) for all , , where (c.f. (4.2)).
Remarks 4.5.
Theorem 4.6.
Subject to Assumptions 4.4, for all , extends to a continuous operator from to , and extends to a continuous operator from to .
Proof.
The proof of the first part is omitted, since it is an easy adaptation of the proof of Theorem 4.8 on page 55 of Hoh (1998) — first proved as Corollary 3.1 in Jacob (1994).
The second part is also proved similarly to Theorem 4.8 of Hoh (1998), the main difference being that takes the place of the transformed symbol, as discussed previously (see (4.3)). By (4.16) and the Plancherel theorem,
Then, using (4.15), Lemma 4.1 and Young’s convolution inequality111See Simon (2015) Theorem 6.6.3, page 550. Here, we are again identifying with a Euclidean space.,
for all . Hence, for all ,
and extends to a bounded linear operator . ∎
Under an additional assumption, we are able to obtain a more powerful result.
Theorem 4.7.
Suppose Assumptions 4.4 hold, and suppose further that satisfies . Then extends to a continuous linear operator from .
We first need a technical lemma.
Lemma 4.8.
Proof.
This is a special case of a bound obtained in Hoh (1998) — see page 50, lines 5–11. ∎
Proof of Theorem 4.7.
By Theorem 4.6, it suffices to prove that extends to a continuous operator from . Given ,
(4.19) | ||||
Also, by Theorem 4.6 and Theorem 2.12 (2),
(4.20) |
where . We will estimate
Our method is similar to that in Theorem 4.3 of Hoh (1998), and so some details are omitted. The map replaces the transformed symbol once again.
To prove (4.1), we seek solutions to the equation
(4.22) |
for a given function and . Consider the bilinear form defined by
Theorem 4.9.
Suppose Assumptions 4.4 hold with . Then extends continuously to .
Proof.
This proof is very similar to those of Jacob (1994) Lemma 3.2, pp. 160, and Hoh (1998) Theorem 4.9, pp. 56, and so we give only a sketch.
Let . Using Assumption 4.4 (1) and the fact that is continuous, there is such that . Plancherel’s identity may then be used to show that
Furthermore, methods similar to the proof of Theorem 4.7 are used to show that
(4.23) |
where
(4.24) |
By Theorem 2.12 (2), there is such that , and thus
for all , which proves the theorem. ∎
The following assumption will ensure that for sufficiently large, is coercive on . We will then use the Lax–Milgram theorem to obtain a weak solution to (4.22).
Assumption 4.10.
Remark 4.11.
The next theorem is an analogue of Theorem 3.1 of Jacob (1994).
Theorem 4.12.
Proof.
Theorem 4.13.
Let . Then (4.22) has a weak solution in the following sense: for all there is a unique such that for all ,
Proof.
Apply the Lax–Milgram theorem (Theorem 1 of Evans (1998), pp. 297) to , using the linear functional . ∎
Having found a weak solution to (4.22), the next task is to prove that this solution is in fact a strong solution that belongs to . This will be achieved using the Sobolev embedding of Theorem 2.12 (6).
Just as in Jacob (1994) Theorem 3.1 and Hoh (1998) Theorem 4.11, we have a useful lower bound for the pseudodifferential operator acting on , when .
Theorem 4.14.
Proof.
The proof is formally no different to the sources mentioned: let , and use (4.25) and Theorem 2.12 (5) to prove that
(4.26) |
for some . Recall the estimate (4.21) of from the proof of Theorem 4.7. In light of Assumption 4.10 and the particular form chosen for , one can use (4.21) to show that
where is a constant. Using Theorem 2.12 (5) once again, let such that
Then, by the above,
(4.27) |
Combining (4.26) and (4.27), we get
∎
The proof of the next theorem makes use of a particular family of bounded linear operators on , which will play the role of a Friedrich mollifier, but in the noncompact symmetric space setting.
First note that by identifying with via our chosen basis, it makes sense to consider Friedrich mollifiers on . For and , let
where is a constant chosen so that . This mollifier is used frequently in Evans (1998) (see Appendix C.4, pp. 629), and Jacob (1994) and Hoh (1998) use it to pass from a weak solution result to a strong solution result.
Observe that for all . Using Theorem 2.1, let be such that
where denotes the Euclidean Fourier transform (see equation (2.7)). For , let be the convolution operator defined on by
The most important properties of needed for the proof of Theorem 4.17 are stated below, and proven in the Section 6.
Proposition 4.15.
-
1.
for all and .
-
2.
For all , is a self-adjoint contraction of .
-
3.
for all , and , and if , then
-
4.
For all and , as .
The following commutator estimate will also be useful in the proof of Theorem 4.17.
Lemma 4.16.
Proof.
Let and , and observe that by Proposition 4.15 (2),
for all , so . For , let , as previously (c.f. (4.2)). Then by (4.6) and Proposition 4.15 (1), for all ,
Applying (4.6) once more,
(4.28) |
for all . From here, a straightforward adaptation to the proof of Hoh (1998) Theorem 4.4, pp. 51–52, with (4.28) replacing Hoh (1998) (4.23), completes the proof of the lemma. ∎
We are now ready to state and prove that, subject to our conditions, a strong solution to (4.22) exists, and belongs to an anisotropic Sobolev space of suitably high order.
Theorem 4.17.
Proof.
Let . By Theorem 2.12 we also have , and so by Theorem 4.13 there is a unique such that
(4.30) |
The proof follows that of Jacob (1994) Theorem 4.3, pp. 163 and Hoh (1998) Theorem 4.12, pp. 59, using induction to show that that for , and in particular, that . The family of operators take over role of the Friedrich mollifiers of Jacob (1994) and Hoh (1998). By Proposition 4.15 these operators satisfy the properties needed for the proof to carry over with little alteration. Lemma 4.16 and Theorem 4.14 replace Hoh (1998) Theorem 4.4 and 4.11, respectively. ∎
Theorem 4.18.
Proof.
Fix with . Let denote the linear operator on with domain , defined by for all . By a similar argument to that on page 60 of Hoh (1998), one can show using that is a operator core for , with
for all . Here, Theorem 2.12 (6) replaces Hoh (1998) Proposition 4.1, and Theorem 4.7 replaces Hoh (1998) Theorems 4.8 and 4.11.
Let be as in Theorem 4.17. We show that for all . Given , choose a sequence in such that as . Then for all , and thus for all . ∎
Corollary 4.19.
5 A Class of Examples
We now present a class of Gangolli symbols that satisfy the conditions of Corollary 4.19. Let such that . We consider symbols of the form
(5.1) |
where is a positive constant, is a Gangolli exponent satisfying (2.18), is non-negative, and is a Gangolli exponent satisfying, for some ,
(5.2) |
By Example 3.8, the mappings and are both Gangolli symbols, and hence so is .
For each and , let
(5.3) |
Observe that is of the form (4.13): since has compact support, , and if , then .
Proof.
For Assumption 4.4 (2), note that in the case we are considering,
and so, for ,
for all and . By (5.2),
For each , a noncommutative version of the multinomial theorem tells us that
(5.4) |
for some coefficients , where and . Expanding the right-hand side of (5.4) using the fact that each is a derivation will give a large sum of terms of the form
where the are constants, and are products of powers of , each with degree at most . Let be the set of all the ’s and the set of all the ’s, so that
(5.5) |
The following bound will be useful.
Lemma 5.2.
For all , there is a constant such that
(5.6) |
for all and .
Proof.
Proof.
It is clear by construction that for all .
To verify the rest of Assumption 4.4 (2), it will be useful to assume that is even. Note that this is an acceptable assumption, since if is odd, we may replace it with — the conditions of Corollary 4.19 will still be satisfied. Let . We seek for which
(5.7) |
Let . Assume first that is even, so that . By (5.5) and Lemma 5.2,
since . Now, for all , and therefore,
where
Let
(5.8) |
Then , since each is a continuous function of compact support. Moreover,
(5.9) |
In particular, we have verified (5.7) when is even.
Assume now that is odd, so that . Since is even, note also that . Applying to both sides of (5.5),
(5.10) |
The families and now each consist of differential operators of degree at most .
Now, is the infinitesimal generator of the process obtained by subordinating Brownian motion on by the standard -stable subordinator on . By standard subordination theory (see Applebaum (2014) §5.7, pp. 154) may be expressed as a Bochner integral
(5.11) |
where denotes the heat semigroup generated by .
Given , and ,
(5.12) | ||||
Let denote the heat kernel associated with . For the term of (5.12), note that , and so
By Lemma 5.2 and the fact that ,
(5.13) |
where is as in (5.6), and . Thus
where
(5.14) |
Since and for all , it follows that for all ,
(5.15) |
We claim that . Clearly , since it is a continuous function of compact support. Each of the operators is a positivity preserving contraction of , and so
By Fubini’s theorem, , with
It follows by (5.14) that , and that
(5.16) |
For the term of (5.12), observe that by Lemma 6.1.12 of Davies (2007), pp. 169, as well as the Fubini theorem,
Hence, using the product formula for ,
(5.17) | ||||
Let and be as in (5.13). Then for all ,
In exactly the same way, for ,
and also
where the constants are chosen so that for all and ,
and . Such constants exist by Lemma 5.2. Now,
for all , and hence by (5.17),
Since , it follows that for all and ,
(5.18) |
where
(5.19) |
Observe that for all . Indeed, , and , hence , and are all continuous functions of compact support. Thus , and, since is an -contraction,
Noting that , it follows by Fubini’s theorem that , with
(5.20) |
Substituting (5.18) and (5.15) into (5.12), we obtain the pointwise estimate
(5.21) |
for all , and , where the () are given by (5.14) and (5.19). Hence by (5.10), for all ,
where
(5.22) |
and is still assumed to be odd. As already noted, for all , and hence . Moreover, by (5.16) and (5.20),
(5.23) |
for some positive constant . In particular, we have verified (5.7) when is odd. ∎
6 Proof of Proposition 4.15
-
1.
Let and . Using a change of variable ,
-
2.
The map is symmetric under , and hence is real-valued. Therefore, given and ,
To see that is a contraction, note that for all , and so by Plancherel’s identity
for all and all .
-
3.
Let and . By Theorem 2.1, , and hence there is such that
Then, using Proposition 2.11 (3),
for all . Let . By Plancherel’s identity,
By Proposition 4.15 (1), , for all , and hence
That is, .
Next, suppose . Then, since ,
as desired.
- 4.
Acknowledgement.
Many thanks to David Applebaum for his advice and support with writing this paper. Thanks also to the University of Sheffield’s School of Mathematics and Statistics, and to the EPSRC for providing PhD funding while this research was carried out.
References
- Anker (1990) J.P. Anker. Fourier multipliers on Riemannian symmetric spaces of the noncompact type. Annals of Mathematics, 132:597–628, 1990.
- Applebaum (2011a) D. Applebaum. Infinitely divisible central probability measures on compact Lie groups - regularity, semigroups and transition kernels. Ann. Probab., 39(6):2474 – 2496, 2011a.
- Applebaum (2011b) D. Applebaum. Pseudo differential operators and Markov semigroups on compact Lie groups. J. Math. Anal. Appl., 384(2):331–348, 2011b.
- Applebaum (2013) D. Applebaum. Aspects of recurrence and transience for Lévy processes in transformation groups and noncompact Riemannian symmetric pairs. J. Aust. Math. Soc., 94(3):304–320, 2013.
- Applebaum (2014) D. Applebaum. Probability on Compact Lie Groups. Springer, 2014.
- Applebaum and Cohen (2004) D. Applebaum and S. Cohen. Lévy processes, pseudo-differential operators and Dirichlet forms in the Heisenberg group. Ann. Fac. Sci. Toulouse Math., Ser. 6, 13(2):149–177, 2004.
- Applebaum and Le Ngan (2020a) D. Applebaum and T. Le Ngan. The positive maximum principle on Lie groups. J. London Math. Soc., 101:136–155, 2020a.
- Applebaum and Le Ngan (2020b) D. Applebaum and T. Le Ngan. The positive maximum principle on symmetric spaces. Positivity, 24:1519–1533, 2020b.
- Berg and Forst. (1975) C. Berg and G Forst. Potential Theory on Locally Compact Abelian Groups. Springer, 1975.
- Böttcher et al. (2013) B. Böttcher, R. Schilling, and J. Wang. Lévy Matters III, Lévy Type Processes: Construction, Approximation and Sample Path Properties. Springer, 2013.
- Davies (2007) E.B. Davies. Linear Operators and their Spectra. Cambridge Univ. Press, 2007.
- Elworthy (1988) K.D. Elworthy. Geometric Aspects of Diffusions on Manifolds. Springer, 1988.
- Ethier and Kurtz (1986) S.N. Ethier and T.G Kurtz. Markov Processes: Characterization and Convergence. Wiley, 1986.
- Evans (1998) L. Evans. Partial Differential Equations. American Math. Soc., 1998.
- Gangolli (1964) R. Gangolli. Isotropic infinitely divisible measures on symmetric spaces. Acta Math., 111:213–246, 1964.
- Gangolli and Varadarajan (1980) R. Gangolli and V.S. Varadarajan. Harmonic Analysis of Spherical Functions on Real Reductive Groups. Springer, 1980.
- Getoor (1961) R. K. Getoor. Infinitely divisible probabilities on the hyperbolic plane. Pacific J. Math., 11(4):1287–1308, 1961.
- Harish-Chandra (1958a) Harish-Chandra. Spherical functions on a semi-simple Lie group I. Amer. J. Math., 80(2):241–310, 1958a.
- Harish-Chandra (1958b) Harish-Chandra. Spherical functions on a semi-simple Lie group II. Amer. J. Math., 80(3):553–613, 1958b.
- (20) S. Helgason. Supplementary Notes to: S. Helgason: Groups and Geometric Analysis, Math. Surveys and Monographs, Vol.83, Amer. Math. Soc. 2000. http://www-math.mit.edu/~helgason/group-geoanal-vol83.pdf.
- Helgason (1984) S. Helgason. Groups and Geometric Analysis. Academic Press, Inc., 1984.
- Helgason (2001) S. Helgason. Differential Geometry, Lie Groups and Symmetric Spaces. AMS; New Ed edition, 2001.
- Heyer (1977) H. Heyer. Probability Measures on Locally Compact Groups. Springer-Verlag, 1977.
- Hoh (1998) W. Hoh. Pseudo Differential Operators Generating Markov Processes. Habilitationsschrift, Bielefeld, 1998.
- Hsu (2002) E.P. Hsu. Stochastic Analysis on Manifolds. American Math. Soc., 2002.
- Hunt (1956) G.A. Hunt. Semi-groups of measures on Lie groups. Trans. Amer. Math. Soc., 81:264–293, 1956.
- Jacob (1993) N. Jacob. Further pseudodifferential operators generating Feller semigroups and Dirichlet forms. Revista Matemática Iberoamericana, 9(2):373–407, 1993.
- Jacob (1994) N. Jacob. A class of Feller semigroups generated by pseudo differential operators. Math. Zeitschrift, 215(1):1432–1823, 1994.
- Jacob (2001a) N. Jacob. Pseudo Differential Operators and Markov Processes, volume I. Imperial College Press, 2001a.
- Jacob (2001b) N. Jacob. Pseudo Differential Operators and Markov Processes, volume II. Imperial College Press, 2001b.
- Knopova et al. (2015) V. Knopova, R. Schilling, and J. Wang. Lower bounds of the Hausdorff dimension for the images of Feller processes. Stat. and Probab. Letters, 97:222–228, 2015.
- Kunita (2019) H. Kunita. Stochastic Flows and Jump Diffusions. Springer, 2019.
- Liao (2004) M. Liao. Lévy Processes on Lie groups. Cambridge Univ. Press, 2004.
- Liao (2018) M. Liao. Invariant Markov Processes Under Lie Group Actions. Springer, 2018.
- Liao and Wang (2007) M. Liao and L. Wang. Lévy–Khinchin formula and existence of densities for convolution semigroups on symmetric spaces. Potential Analysis, 27(2):133–150, 2007.
- Ngan (2019) T.L. Ngan. The Positive Maximum Principle on Lie Groups and Symmetric Spaces. PhD thesis, The University of Sheffield, 2019. URL http://etheses.whiterose.ac.uk/id/eprint/22779.
- Sawyer (2003) P. Sawyer. The Abel transform on symmetric spaces of noncompact type. Amer. Math. Soc. Transl., 210(2):331–355, 2003.
- Schilling (1998a) R. Schilling. Conservativeness and extensions of Feller semigroups. Potential Analysis, 9:91–104, 1998a.
- Schilling (1998b) R. Schilling. Conservativeness of semigroups generated by pseudo differential operators. Potential Analysis, 9:91–104, 1998b.
- Schilling (1998c) R. Schilling. Growth and Hölder conditions for sample paths of Feller processes. Probab. Theory Relat. Fields, 112:565–611, 1998c.
- Schilling and Schnurr (2015) R. Schilling and A. Schnurr. The symbol associated with the solution of a stochastic differential equation. Electron. J. Probab., 15:1369–1393, 2015.
- Shewell Brockway (in preparation) R Shewell Brockway. PhD thesis, University of Sheffield, in preparation.
- Simon (2015) B. Simon. Real Analysis, A Comprehensive Course in Analysis, Part 1. American Math. Soc., 2015.