2. Preliminary results
Here we provide some auxiliary results on the limiting Feller process for the scaled Markov chains governing the sizes and times of jumps. These results are obtained by combining the method of proving well-posedness of processes generated by operators of order at most one (from [
20]) with general convergence results from [
8].
Our main assumptions are as follows:
Condition (A) on the distribution of times: the family
is given by (
4) with
with some constants
, and
;
Condition (B) on the distribution of velocities: is a family of probability laws on such that the family of measures is tight (in particular, uniformly bounded);
Condition (C) on the first order regularity of spatial distributions: the derivative of with respect to x, exists as a family of signed vector-valued uniformly bounded measures such that the family is tight;
Condition (D) on the second order regularity: the second order derivatives of with respect to x exist as uniformly bounded and tight families of signed measures, and .
It is known (see e.g. Theorem 19.28 of [
24] or Theorem 8.1.1 of [
20]) that if the chains with transitions
(with a family of transitions given by (
5)) converges to a Feller process
, as
, then the generator
of the corresponding limiting semigroup can be obtained as the limit
We shall need the following simple result (a proof can be found e.g. in [
22] or [
21]):
Lemma 2.1.
Let be a probability density on such that for with some and such that (the latter condition comes from the requirement that ). Then for any Lipschitz continuous vanishing at zero, it follows that
where L is the Lipschitz constant of f.
Applying this result, yields
The next simple proposition is our first (preliminary) result.
Proposition 2.1. Assume conditions (A) - (C) hold. Then operator (14) generates a Feller process in and a corresponding Feller semigroup in , which has as an invariant core.
The conditions are just slightly different from that of Theorem 5.1.1 from [
20] (or Theorem 5.14.1 from [
22]). The proof is exactly the same. We omit it, but will show the arguments below in a more involved case of processes with a boundary.
Thus the limit (
12) exists for functions
F from the core of the limiting process. Hence the standard results (Theorem 19.28 of [
24]) implies the following direct consequence of Proposition 2.1:
Proposition 2.2. Assume conditions (A) - (C) hold. Then the chains with transitions arising from (5) converge in distribution to the Feller process , as .
Let us define the right continuous inverse process to
by the formula
As usual, by we denote the left continuous modification of .
Once Propositions 2.1 and 2.2 are obtained, the fundamental Theorem 3.6 from [
8] can be applied to conclude the following:
Corollary 1. The leading (or overshooting) and lagging CTRWs, and , from (6) converge in distribution to the process and, respectively, to the process , which is the right continuous version of the process .
By (
8), this implies the following:
Corollary 2. The intermediate CTRWs converge to , which is the spatial coordinate of the point of intersection of the line, joining and , with the boundary hyperplane .
Our aim is to introduce and to analyse the fractional pseudo-differential equations that govern the evolution of the processes , , .
3. Main results
3.1. Material derivatives
Integrating by parts, (
14) can be rewritten as
Recalling the expression for the standard (right) fractional derivative of an order
:
one can say that the integral
represents (up to a positive multiplier that we shall neglect) the fractional material derivative, where the material derivative (in the direction
v) is defined as
Thus we can write down the generator
L in the form
where the r.h.s. is the averaged (over
v) fractional material derivative.
Continuing the analogy, let us note that if
vanishes at
, then the right Riemann-Liouiville derivative can be defined as the restriction of
to the space of functions vanishing for
, that is, as the operator
Similarly, operators (
16) and (
17), reduced to the space of smooth functions vanishing for
, take the forms
and
Thus represents a multidimensional analog of the standard Riemann-Liouville fractional derivative (of variable order in our case), so that the inverse of this operator (if well defined) will represent a multidimensional analog of the standard Riemann-Liouville fractional integral.
When looking for a probabilistic interpretation of fractional derivatives in [
23], it was noted that the Riemann-Liouville derivative is obtained from the free (without boundary) fractional operator by restricting it to the space of functions vanishing identically beyond the boundary, which in terms of the underlying stochastic process means its killing on the attempt to cross the boundary. In its turn, the Caputo-Dzherbashian derivative is obtained from the free fractional operator by restricting it to the space of functions that are constant beyond the boundary, which in terms of the underlying stochastic process means its stopping on the attempt to cross the boundary. This interpretation of fractional derivatives led to the natural extension of the fractional derivatives not only to a two-sided case, but to a variety of multidimensional settings. However, while killing on the boundary has always clear meaning, stopping for a multidimensional jump-type process really depends on the way one projects the result of the final jump (that crosses the boundary) to the boundary, which leads to several different version of the Caputo-Dzherbashian fractional operators. Three of them are analysed in this paper.
3.2. Stopped and killed limiting generators
Using Lemma 2.1 allows us to conclude that the limits of
as
, exist for
and equal, respectively,
Remark 1. Let us comment for clarity that the processes described by these three generators differ only by the last jump, that is, by the jump that is meant to cross the boundary. Saying more precisely, the last jump means really the last period of motion with a constant velocity. In the last jump (when ) just does not occur at all (hence the shift is multiplied by the indicator ). In the last jump occurs in full (thus confirming the term overshooting). In the last period of motion with constant velocity is interrupted exactly on crossing the boundary , which makes this process the most natural one from author’s point of view.
One can rewrite these expressions in the following equivalent forms:
Integrating by parts yields
which is exactly the averaged fractional material derivative (
18).
We shall denote by the general operators with * denoting either lag, or lead, or int. Our main technical results, given below, concern the existence of well defined Feller processes in generated by .
To work with these operators let us denote by (or sometimes shorter by ) the subspace of of functions vanishing at the boundary , and by the subspace of with all partial derivatives belonging to .
The elementary properties of are collected in the following proposition, its proof being obtained by a direct inspection that we omit.
Proposition 3.1. Assume conditions (A)-(B). Then all three are bounded operators from to . Moreover, the image of and belong to .
We are also interested in the versions of these processes killed on the boundary
. The semigroups arising from killed processes act on the space
. It is seen that in this space all three operators
coincide. Let us denote them by
:
As expected, this is nothing else, but the operator (
19), which represents (up to a sign) a multi-dimensional analog of the Riemann-Liouville fractional operator.
3.3. Formulation of the technical results: stopped and killed limiting processes
Theorem 3.1. Under conditions(A)-(C) the operator generates a Feller semigroup in the space with being its invariant core. Moreover, this semigroup is also strongly continuous in . Finally, the potential operator is well defined as a bounded operator both in and .
Theorem 3.2.
Under conditions(A)-(D) the operators and generate Feller semigroups and , respectively, in the space such that all points of the boundary are rest points for the corresponding Feller processes. For an invariant core can be taken as the subspace of consisting of functions with the derivative with respect to w vanishing at the boundary . For an invariant core can be taken as the subspace of consisting of functions with the averaged material derivative
vanishing at the boundary . Moreover, these semigroups are also strongly continuous in these cores considered as Banach subspaces of .
Remark 2. In the case of symmetric distribution of velocities, e.g. if , the spaces and coincide.
Theorem 3.3. Under conditions(A)-(D) the operator generates a Feller semigroup in the space with an invariant core . This semigroup is strongly continuous in this core considered as a Banach subspace of .
The proof of all these results follow the same line of arguments. We shall give details for Theorem 3.1 in
Section 4 and briefly comment on modifications arising in other cases.
The following result is a straightforward but important corollary of the theorems given above.
Proposition 3.2. The semigroups , , represent different extensions of the semigroup from the space to the space . The domain of the operator lies in the intersection of the domains of the operators , , .
Finally, when working with
, we shall need to use functions from the domain that are not differentiable up to the boundary. Let us introduce the following functional space
, which is the subspace of
consisting of functions
such that the spatial gradient
exists and belongs to
and, with respect to the second variable,
F is locally Hölder in the sense that the function
is well defined and belongs to
. It is seen that for any
formulas (
20), (
21), and (
22) yield well defined functions from
. Consequently, using the fact that the generator of any Feller semigroup is a closed operator, and approximating the functions from
by the functions from the corresponding invariant cores of
(given by the above Theorems) we obtain the following fact.
Proposition 3.3. The space belongs to the domain of the generators of all semigroups , , , and the space belongs to the domain of the generator of the semigroup .
3.4. Main results on the limiting fractional equations
Let us start with the analogs of the Riemann-Liouville fractional operators.
Theorem 3.4.
(i) For any there exists a unique classical solution (classical in the sense that G lies in the domain of ) to the equation
(ii) The solution G has the following path integral (probabilistic) interpretation:
where is given by (15).
(iii) If , thenas well.
Proof. Statements (i) and (iii) are direct consequences of Theorem 3.1. Representation (
28) is the standard probabilistic representation for the potential operator that is routinely derived from the Dynkin martingale (see detail of a similar derivation in the proof of the next Theorem). □
Theorem 3.5.
(i) For any there exists a unique classical (in the sense that it belongs to the domain of ) solution of the multi-dimensional fractional Cauchy problem (with material fractional derivatives)
where * in denotes either lag, or int, or lead;
(ii) This solution has the following probabilistic representation:
and where is given by (15).
(iii) If, then, in case of either lag or int, this solutionFbelongs to the space.
Proof. (i) We claim that there exists a function from the domain of the generator such that and .
For the case of either lag or int such a function can be easily chosen from the space (implying, by Proposition 3.1, that ). In fact, one can take , and must be chosen from the requirement that its material derivative vanishes on the boundary .
The case of
is a bit more subtle, as
cannot be chosen from
. By Proposition 3.3, we can search for an appropriate
in the space
. And this is possible, because, as follows from (
24), if
, then
Consequently, for a given smooth , one can choose such that the last two terms in the last expression cancel.
With
chosen in the way, required above, we see that the function
belongs to
and satisfies the equation
Since
, we can conclude by Theorem 3.4, that there exists a unique classical solution
of problem (
31). Therefore, by Proposition 3.2, the function
belongs to the domain of
and represents the unique solution of the original problem.
(ii) Representation (
30) is obtained by the straightforward application of the Dynkin martingale. Namely, since
is a Feller process, it follows that the process
is a martingale for any
. By (
29),
. Then (
30) follows from Doob’s optional sampling theorem and an evident observation (see the end of the proof of Theorem 3.1, if necessary) that the stopping time
has a finite expectation.
(iii) This follows from Theorem 3.4 (iii) and the observation that whenever . □
Remark 3. Of course, once Theorem 3.5 or 3.4 are proved, one can use formulas (30) or (28) to define generalised solutions for the corresponding problems for an arbitrary continuous function ϕ.
3.5. Modification: motions with a fixed random acceleration or parameter depending velocity
For a particle in random media a reasonable model is represented by a process that moves with a constant acceleration between random stops, see e.g. [
25]. This suggests to look at a modification of Lévy walks, that can be called Lévy runs, where, after each switching, the particle starts moving with some constant acceleration (rather than velocity, as in Lévy walks) drawn randomly from some distribution. Fractional equations arising in the natural scaling limit of such processes are straightforward modifications of the above case with constant velocity.
Namely, the corresponding operator (
14) of the limiting Markov process without a boundary changes to the operator
where
is the distribution of accelerations chosen at the position
x. The Riemann-Liouville-type operator (
27) of the killed process changes to the operator
Similar modifications can be written for the three versions of Caputo-Dzherbashian fractional derivatives. All results above have straightforward extension to this new model with constant accelerations between switching times.
This model is related to the model with parameter dependent velocity suggested in [
26]. To combine these models we can suggest to substitute
in (
2) by a more general smooth function
such that
for all
v. The theory above can be carried out for this situation with more or less obvious corrections. Namely, possible growth of
in
v should be compensated by the assumption of the existence of appropriate moments of
.
4. Proofs of Theorems 3.1–3.3
4.1. Approximations
To build the processes generated by (including ) we shall use appropriate approximations. For an let be a smooth function such that for and for . Let denote the operator obtained by changing to in the formula for . One sees that all are bounded operators in the space such that the images of and belong to . Consequently, all generate Feller semigroups in and hence the corresponding Feller processes in . For the cases of and all points of the boundary are rest points for these processes.
We are going to construct the processes generated by as the limits of the corresponding processes generated by . To perform these limits we are going to show that the semigroups are uniformly (in ) bounded as semigroups in certain subspaces of .
4.2. Proof of Theorem 3.1
Recall that we consider the operator
as a bounded operator in
. Denoting
we obtain
Differentiating with respect to
w (taking into account that
and that
F vanishes on the boundary
) yields
The last two terms cancel yielding
Differentiating with respect to
x yields
Since is bounded by times the -norm of F, it follows that all terms in this expression apart from the first one (that generates a contraction semigroup) are uniformly bounded in . Moreover, since vanishes at the boundary , it follows that also vanishes at this boundary. Therefore, due to the perturbation theory, the operators generate strongly continuous semigroups in , which are uniformly bounded in .
Consequently, we may conclude that the derivatives of
are uniformly bounded functions (at least for
t from any compact interval, which is sufficient for our purposes) for any initial
. Therefore, writing
we conclude that, for
,
as
. Hence the functions
converge to a function
.
Convergence for extends to the convergence for by the standard density argument. Therefore the family of contraction operators converges to a family , as . Clearly the limiting family is also a strongly continuous semigroup of contractions in .
Writing
and noting that by (
37) the first term is of order
, as
, allows one to conclude that
belongs to the domain of the generator of the semigroup
in
and that it is given there by (
27).
To show that is an invariant core, we can apply to the procedure applied above to . Namely, differentiating we find that, on the partial derivatives of F, the operator acts as the diagonal operator (with on the diagonal) plus a uniformly bounded operator. Thus, again referring to the standard perturbation theory, we conclude that the operators act as a uniformly bounded strongly continuous semigroup in .
Finally, the potential operator
is known to be expressed via the semigroup by the following formula:
Since the coordinate
increases faster than certain Poisson process
with the generator
with some
, one can very roughly (but sufficiently for us) estimate the probability that
by the probability
Thus the potential operator is a bounded operator in , as was claimed. Quite similarly, one shows that this operator is bounded in the space .
4.3. Proof of Theorem 3.2
Differentiating
and
with respect to
x shows again (as for the case of
) that the action of these operators on the spatial derivatives is the same as that of
and
, respectively, up to some uniformly (in
) bounded operators. Moreover,
and
vanish at the boundary for any
. New features arise when differentiating with
w. After some cancellations, similar to the case of
we find that, for
,
and
It follows that if , then , and thus the subspace is invariant under the action of the semigroup . Similarly, if the averaged material derivative vanishes at the boundary , then , and thus the subspace is invariant under the action of the semigroup .
Arguing now as for case of the killed process we find that, for any the functions converge, as , in the space to some functions . Extending this convergence by the density argument we conclude that the contraction operators converge strongly in the space to some contraction operators that form a strongly continuous semigroup in the space such that the space belongs to the core of its generator.
Similarly we find that, for any the functions converge in the space to some functions . Extending this convergence by the density argument we conclude that the contraction operators converge strongly in the space to some contraction operators that form a strongly continuous semigroup in the space such that the space belongs to the core of its generator.
However, we cannot complete the proof as for the killed process, because it is not obvious that the derivatives of or with respect to w remain bounded under the action of the corresponding semigroups. Therefore, in this case, we have to use the second order regularity condition (D), to work with the second order derivatives and then show, in the same way as for the first order derivatives, that the semigroups and are strongly continuous in the space . For instance, in the case of , we first check that the subspace of consisting of functions with the first and the second derivatives in w vanishing at the boundary is invariant under , and then show the convergence, as , of the functions for F in this subspace, the convergence being in the space . Then we extend this convergence by the density argument to all F from , and thus complete the proof.
4.4. Proof of Theorem 3.3
Differentiating with respect to
x yields
It follows that the space is invariant under the action of the semigroup , as in the case of the semigroup . However, unlike the latter, the generator does not vanish on the boundary . The rest of the proof is the same as for Theorem 3.2.
5. Extension: including waiting times
In the literature on Lévy walks one often assumes additionally that a particle waits some random time after a move, before starting a new one.
Allowing for additional waiting time means that the transitions (
2) are modified and turn to the transitions
with some family of probabilities
with the tails given by some
with
. To be more concrete, we assume, analogously to (
4) that
has density
such that
with some constants
and
.
Then the scaled version (
5) extends to the transitions
The corresponding prelimiting operator (
12) converges on the set of smooth functions to the operator
To get (
44) one just writes down
and apply Lemma 2.1 to both terms.
Thus, the sequential shift of the second (time) coordinate in (
43) turns to the sum of independent shifts, when passing to the limit.
Straightforward extension of Propositions 2.1, 2.2 yields the following:
Proposition 5.1. Assume that conditions (A) - (C) and (42) hold and is continuously differentiable. Then operator (44) generates a Feller process in and a corresponding Feller semigroup in , which has as an invariant core. The chains with transitions arising from (43) converge in distribution to the Feller process , as .
Let us now write down the corresponding extensions of stopped processes. Since we first wait and then jump, we will be stopped if either the waiting time is crossing the boundary
or, otherwise, if we cross the boundary when moving. Thus the lagging stopped version of (
43) will be
Similarly other transitions are defined by adding additional waiting times to the transitions of .
To find the limiting generator, we are looking for the limit of
. By (
42), as
,
To deal with this expression we again use (
45) and Lemma 2.1 yielding
Similar calculations work for other
leading to the following formulas:
The results for and the corresponding processes extend to the version with additional waiting times. However, to avoid technical complications, we make additional simplifying assumption:
Condition (E): for the results below concerning we assume that the distribution of velocities is symmetric: for all x; for the results concerning we assume that either for all x, or for all x; nothing additional for and .
Theorem 5.1. Under conditions of Proposition 5.1 supplemented by Condition (E), the results of Theorems 3.1, 3.2, 3.3, as well as Theorems 3.4, 3.5, extend literally to the operator .
Proof. The extension of all proofs is straightforward. Let us note only that condition (E) for is needed, while otherwise the boundary conditions of spaces and do not coincide and therefore neither can be chosen as an invariant subspace for such that the application of to this subspace belongs to . The condition (E) for is needed for choosing in the extension of the proof of Theorem 3.5. □