845
Views
3
CrossRef citations to date
0
Altmetric
Articles

Fractional Klein-Gordon equation with singular mass. II: hypoelliptic case

, &
Pages 615-632 | Received 20 Apr 2021, Accepted 28 Jun 2021, Published online: 14 Jul 2021

Abstract

In this paper we consider a fractional wave equation for hypoelliptic operators with a singular mass term depending on the spacial variable and prove that it has a very weak solution. Such analysis can be conveniently realised in the setting of graded Lie groups. The uniqueness of the very weak solution, and the consistency with the classical solution are also proved, under suitable considerations. This extends and improves the results obtained in the first part [Altybay et al. Fractional Klein-Gordon equation with singular mass. Chaos Solitons Fractals. 2021;143:Article ID 110579] which was devoted to the classical Euclidean Klein-Gordon equation.

AMS Subject Classifications:

1. Introduction

The aim of this paper is to contribute to the study of the Klein-Gordon equation for positive (left) Rockland operator R (left-invariant hypoelliptic partial differential operator which is homogeneous of positive degree ν) on a general graded Lie group G, with a possibly singular mass term depending on the spacial variable; that is for T>0, and for s>0 we consider the Cauchy problem (1) utt(t,x)+Rsu(t,x)+m(x)u(t,x)=0,(t,x)[0,T]×G,u(0,x)=u0(x),ut(0,x)=u1(x),xG,(1) where m is a non-negative and possibly singular function/distribution.

The setting of Rockland operators on graded Lie groups allows one to consider both elliptic and subelliptic settings in (Equation1). The simplest example is that of the standard Klein-Gordon equation, when we take G=Rd to be the Euclidean space, and R=Δ to be the Laplacian on Rd. However, already on Rd, the setting of (Equation1) allows one to consider more general evolutions, for example, taking R=(1)mj=1d2mxj2m, for any integer m. Such operators are also Rockland operators on Rd, as we explain in the next section. However, the general setting of (Equation1) allows one to also consider hypoelliptic operators. The simplest example would be G being the Heisenberg group, and R the positive sub-Laplacian on it. More generally, if G is any stratified group (or a homogeneous Carnot group), and X1,,XN are the generators of its Lie algebra (satisfying the Hörmander condition), we can consider R=(1)mj=1NXj2m, for any integer m, where we understand Xj also as the derivative with respect to the vector field Xj.

The main feature of (Equation1) is that we will not assume any regularity on the mass coefficient m. Especially, we are interested in irregular m, for example being δ-distribution, or even δ2, if understood appropriately in the sense of multiplication of distributions. We note that in this situation the usual notion of weak solutions is not applicable to (Equation1) in view of the Schwartz impossibility result [Citation1] on the multiplication of distributions.

Thus, in this paper we work with the concept of very weak solutions. More specifically, we will show its applicability to the Cauchy problem (Equation1) for the Klein-Gordon equation for the Rockland operator R on the graded Lie group G with a singular mass depending on the spacial variable. This concept was introduced in [Citation2] to deal with the Schwartz impossibility result about multiplication of distributions [Citation1], in the context of wave type equations with singular coefficients. Later, this analysis was applied to other hyperbolic type equations with singular coefficients [Citation3–6]. The wave type equations with time-dependent coefficients on graded Lie groups were analysed in [Citation7] for Hölder coefficients, and in [Citation8] for distributional time-dependent coefficients, using the notion of very weak solutions. All these works deal with the time-dependent equations and in the recent papers [Citation9–12], the authors start to develop the notion of very weak solutions for equations with (irregular) space-depending coefficients.

The present paper is the extension and improvement of the results obtained in the first part [Citation9] which was devoted to the classical Klein-Gordon equation. In fact, the setting of [Citation9] was the Equation (Equation1) for G=Rd and R=Δ being the positive Laplacian on the Euclidean space. Consequently, the results here contain the results of [Citation9] as a special case, and we also use this chance to slightly correct the consistency statement given in that paper, see Remark 5.3, as well as a clarifying Remark 4.7.

2. Preliminaries

Let us briefly recall some basic concepts, terminology and notation on graded Lie groups that will be useful for the ideas we develop throughout this paper. For a more detailed exposition we refer to Folland and Stein [Chapter 1 in [Citation13]], or, to the more recent open access book, by Fischer and the second author [Chapter 3 in [Citation14]].

Let G be a nilpotent Lie group, and let g be its Lie algebra. Its lower series is the descending sequence {gi} of ideals defined inductively by g1=g, gi=[g,gi1], for i>1. If g admits a gradation of vector spaces as g=i=1gi, where all, but finitely many gi's are equal to {0}, and is such that [gi,gj]gi+j, for all i, j, then G is a graded Lie group. Graded Lie groups are naturally homogeneous Lie groups; that is g is equipped with a one-parameter family {Dr}r>0 of automorphisms of g of the form Dr=exp(Alogr), with A being a diagonalisable linear operator on g with positive eigenvalues. Such automorphisms shall be called dilations.

We have the following nested subclasses of Lie groups: nilpotenthomogeneousgradedstratified{Heisenberg,Engel,Cartan}. The cases of the Heisenberg, Engel and Cartan groups, are examples of graded Lie groups whose associated representation theory is well-understood in the sense that there exists a complete and explicit classification of the unitary, irreducible representations on them; see e.g. [Citation15,Citation16], as well as the analysis in [Citation17,Citation18]. For graded Lie algebras g of dimension n, the canonical family of dilations, is the one dictated by the gradation of g, and is given by (2) Xi(j)Dr=rviDrXi(j),(2) where Xi(j)gj, i=1,,n, and vi's are the same for all vectors Xi(j)gj. These vi's are called the dilations' weights.

In the case of graded Lie groups, or more generally in the case of nilpotent Lie groups, the exponential map (on the group) is a diffeomorphism from g onto G, under the group law that has been defined accordingly to the structure of g due to the Baker-Campbell-Hausdorff formula; see, e.g. [Citation19]. More generally, this identification allows for the transmission of ideas from the infinitesimal level of the Lie algebra g to the level of the group G. Additionally, when g is homogeneous, then, the dilations can be transported to the group side, while the Lebesgue measure dx on g is the bi-invariant Haar measure on G, and the number Q that satisfies d(Dr(x))=rQdx, that is the sum of the eigenvalues of the matrix A, shall be called the homogeneous dimension of G.

On the other hand, any element πGˆ of the unitary dual of G, with π acting on some separable Hilbert space Hπ, gives rise to the representation dπ on the space of smooth vectors Hπ on the infinitesimal level; that is we can define dπ(X)v:=limt01t(π(exp(tX))vv),Xg, vHπ. The above definition, due to the Poincaré-Birkhoff-Witt Theorem (see, e.g. [Citation20], see also a discussion in [Citation14]), that identifies that space of left-invariant operators in g with the universal enveloping Lie algebra U(g), can be extended to any TU(g), i.e. we can write dπ(T); or, with an abuse of notation, π(T).

A remarkable class among left-invariant operators, that generalises the notion of the sub-Laplacian on the bigger class of graded groups, is that of Rockland operators, which are usually denoted by R. The latter is a class of operators that are hypoelliptic on G [Citation21], and homogeneous of a certain positive degree. So, by Rockland operators we understand the homogeneous left-invariant hypoelliptic differential operators on G. For additional characterisations of the Rockland operators, we refer to [Citation22–24], as well as to a presentation in [Citation14].

We recall that R and π(R), are densely defined on their domains D(G)L2(G), and HπHπ, respectively (cf. [Proposition 4.1.15 in [Citation14]]. The latter implies that the positivity of R, as required for our purposes, amounts to the condition (Rf,f)L2(G)0,fD(G). We remark that, for a positive Rockland operator R, the spectrum of the operator π(R), with πGˆ{1}, is discrete [Citation25], which allows us to choose an orthonormal basis for Hπ such that the self-adjoint operator π(R) can be identified with the infinite dimensional matrix with diagonal elements πk,k2πk2, with πkR+.

Let us now recall that the group Fourier transform of a function fL1(G) at πGˆ is the bounded operator fˆ(π) (often denoted by π(f)) on Hπ given by (fˆ(π)v1,v2)Hπ:=Gf(x)(π(x)v1,v2)Hπdx,v1,v2Hπ. If fL2(G)L1(G), then fˆ(π) is a Hilbert-Schmidt operator, and we have the following isometry, known as the Plancherel formula (3) G|f(x)|2dx=Gˆπ(f)HS2dμ(π),(3) where μ stands for the Plancherel measure on G. For a detailed exposition of the Plancherel Theorem and the relevant theory, we refer to [Citation19,Citation26], or to [Section 1.8, Appendix B.2 in [Citation14]].

Finally, since the action of a Rockland operator R is involved in our analysis, let us make a brief overview of some related properties.

Definition 2.1

Homogeneous Sobolev spaces

For s>0, p>1, and R a positive homogeneous Rockland operator of degree ν, we define the R-Sobolev spaces as the space of tempered distributions S(G) obtained by the completion of S(G)Dom(Rsν) for the norm fL˙sp(G):=RpsνfLp(G),fS(G)Dom(Rpsν), where Rp is the maximal restriction of R to Lp(G).Footnote1

Let us mention that, the above R-Sobolev spaces do not depend on the specific choice of R, in the sense that, different choices of the latter produce equivalent norms, see [Proposition 4.4.20 in [Citation14]].

In the scale of these Sobolev spaces, we recall the next proposition as in [Proposition 4.4.13 in [Citation14]].

Proposition 2.2

Sobolev embeddings

For 1<q~0<q0< and for a,bR such that ba=Q1q~01q0, we have the continuous inclusions L˙bq~0(G)L˙aq0(G), that is, for every fa˙Lbq~0(G), we have fa˙Laq0(G), and there exists some positive constant C=C(q~0,q0,a,b) (independent of f) such that (4) fL˙aq0(G)CfL˙bq~0(G).(4)

In the sequel we will make use of the following notation:

Notation 2.3

  • When we write ab, we will mean that there exists some constant c>0 (independent of any involved parameter) such that acb;

  • if α=(α1,,αn)Nn is some multi-index, then we denote by [α]=i=1nviαi, its homogeneous length, where the vi's stand for the dilations' weights as in (Equation2), and by |α|=i=1nαi, the length of it;

  • for suitable fS(G) we have introduced the following norm fHs(G):=fL˙s2(G)+fL2(G);

  • when regulisations of functions/distributions on G are considered, they must be regarded as arising via convolution with Friedrichs-mollifiers; that is, ψ is a Friedrichs-mollifier, if it is a compactly supported smooth function with Gψdx=1. Then the regularising net is defined as (5) ψϵ(x)=ϵQψ(Dϵ1(x)),ϵ(0,1],(5) where Q is the homogeneous dimension of G.

3. Estimates for the classical solution

Here and thereafter, we consider a fixed power s>0 of a fixed, positive (in the operator sense) Rockland operator R that is assumed to be of homogeneous degree ν. Moreover, the coefficient m in (Equation1) will be regarded to be non-negative on G.

The next two propositions prove the existence and uniqueness of the classical solution to the Cauchy problem (Equation1), in the cases where the coefficient m is such that mL(G) or mL2Qνs(G), where, in the second case, we must additionally require Q>νs.

Proposition 3.1

Let mL(G), m0, and suppose that (u0,u1)Hsν2(G)×L2(G). Then, there exists a unique solution uC([0,T];Hsν2(G))C1([0,T];L2(G)) to the Cauchy problem (Equation1), that satisfies the estimate (6) u(t,)Hsν2(G)+tu(t,)L2(G)(1+mL(G)){u1L2(G)+u0Hsν2(G)},(6) uniformly in t[0,T].

Proof.

Multiplying the Equation (Equation1) by ut and integrating over G, we get (7) (utt(t,),ut(t,)L2(G)+Rsu(t,),ut(t,)L2(G)+m()u(t,),ut(t,)L2(G))=0,(7) for all t[0,T]. It is easy to check that (utt(t,),ut(t,)L2(G))=12tut(t,),ut(t,)L2(G),(Rsu(t,),ut(t,)L2(G))=12tRs2u(t,),Rs2u(t,)L2(G), and (m()u(t,),ut(t,)L2(G))=12tm()u(t,),m(),u(t,)L2(G). Denoting by E(t):=ut(t,)L2(G)2+Rs2u(t,)L2(G)2+m()u(t,)L2(G)2, the energy functional estimate of the system (Equation1), the Equation (Equation7) implies that tE(t)=0, and consequently also that E(t)=E(0), for all t[0,T]. By taking into consideration the estimate (8) m()u0L2(G)2mL(G)u0L2(G)2,(8) by the above, it follows that each positive term that E(t) consists of, is bounded itself. That is, we have that (9) m()u(t,)L2(G)2u1L2(G)2+Rs2u0L2(G)2+mL(G)u0L2(G)2,(9) while also that (10) ut(t,)L2(G)2,Rs2u(t,)L2(G)2u1L2(G)2+u0Hsν2(G)2+mL(G)u0L2(G)2(1+mL(G)){u1L2(G)2+u0Hsν2(G)2},(10) uniformly in t[0,T], where we use Rs2u0L2(G),u0L2(G)u0Hsν2(G). Observe that, to prove (Equation6), it remains to show the desired estimate for the norm u(t,)L2(G). To this end, we first apply the group Fourier transform to (Equation1) with respect to xG and for all πGˆ, and we get (11) uˆtt(t,π)+π(R)s uˆ(t,π)=fˆ(t,π),uˆ(0,π)=uˆ0(π),uˆt(0,π)=uˆ1(π),(11) where fˆ(t,π) denotes the group Fourier transform of the function f(t,x):=m(x)u(t,x). Taking into account the matrix representation of π(R), we rewrite the matrix equation (Equation11) componentwise as the infinite system of equations of the form (12) uˆtt(t,π)k,l+πk2suˆ(t,π)k,l=fˆ(t,π)k,l,(12) with initial conditions uˆ(0,π)k,l=uˆ0(π)k,l and uˆt(0,π)k,l=uˆ1(π)k,l, for all πGˆ and for any k,lN, where now fˆ(t,π)k,l can be regarded as the source term of the second order differential equation as in (Equation12).

Now, let us decouple the matrix equation in (Equation12) by fixing πGˆ, and treat each of the equations represented in (Equation12) individually. If we denote by v(t):=uˆ(t,π)k,l,β2s:=πk2s,f(t):=fˆ(t,π)k,l, and v0:=uˆ0(π)k,l,v1:=uˆ1(π)k,l, then (Equation12) becomes (13) v(t)+β2sv(t)=f(t),v(0)=v0,v(0)=v1,(13) with β>0. By solving first the homogeneous version of (Equation13), and then by applying Duhamel's principle (see e.g. [Citation27]), we get the following representation of the solution of (Equation13) (14) v(t)=cos(tβs)v0+sin(tβs)βsv1+0tsin((ts)βs)βsf(s)ds.(14) Assuming without loss of generality that T1, and using the estimates |cos(tβs)|1, t[0,T], and |sin(tβs)|1, for large values of the quantities tβs, while for small values of them, the estimates |sin(tβs)|tβsTβs, inequality (Equation14) yields |v(t)||v0|+T|v1|+tsL2[0,T]f(t)L2[0,T]|v0|+|v1|+f(t)L2[0,T], where we have applied the Cauchy-Schwarz inequality. Now the last estimate, if substituting back our initial functions in t, gives |uˆ(t,π)k,l|2|uˆ0(π)k,l|2+|uˆ1(π)k,l|2+fˆ(t,π)k,lL2[0,T]2, where the latter holds uniformly in πGˆ and for each k,lN. Recall that for any Hilbert-Schmidt operator A, one has AHS2=k,l|Aφk,φl|2, for any orthonormal basis {φ1,φ2,}, summing the above over k, l we get uˆ(t,π)k,lHS2uˆ0(π)k,lHS2+uˆ1(π)k,lHS2+k,l0T|fˆ(t,π)k,l|2dt. Next we integrate the last inequality with respect to the Plancherel measure μ on Gˆ, so that using the Plancherel identity (Equation3), we obtain (15) u(t,)L2(G)2u0L2(G)2+u1L2(G)2+Gk,l0T|fˆ(t,π)k,l|2dtdμ(π),(15) and if we use Lebesgue's dominated convergence theorem, Fubini's theorem and the Plancherel formula we have (16) Gk,l0T|fˆ(t,π)k,l|2dtdμ=0TGk,l|fˆ(t,π)k,l|2dμdt=0Tf(t,)L2(G)2dt.(16) Now, by (Equation9), and the formula of f we have (17) f(t,)L2(G)2=m()u(t,)L2(G)2mL(G)m()u(t,)L2(G)2(1+mL(G))2{u1L2(G)2+u0Hsν2(G)2}.(17) Combining the inequalities (Equation15), (Equation16) and (Equation17) we get (18) u(t,)L2(G)2(1+mL(G))2{u1L2(G)2+u0Hsν2(G)2},(18) uniformly in t[0,T]. The claim (Equation6) now follows by (Equation10) and (Equation18). Finally, the uniqueness of u is an immediate consequence of (Equation6), and the proof is complete.

Proposition 3.2

Assume that Q>νs, and let mL2Qνs(G)LQνs(G), m0. If we suppose that (u0,u1)Hsν2(G)×L2(G) then there exists a unique solution uC([0,T];Hsν2(G))C1([0,T];L2(G)) to the Cauchy problem (Equation1) satisfying the estimate (19) u(t,)Hsν2(G)+tu(t,)L2(G)1+mL2Qνs(G)1+mLQνs(G)12u1L2(G)+u0Hsν2(G),(19) uniformly in t[0,T].

Proof.

Proceeding as in the proof of Proposition 3.1, we have (20) E(t)=E(0), t[0,T],(20) where the energy estimate E is given by E(t)=ut(t,)L2(G)2+Rs2u(t,)L2(G)2+m()u(t,)L2(G)2. Now, applying Hölder's inequality, we get (21) mu0L2(G)2mLq(G)u0L2q(G)2,(21) where 1<q,q<, and (q,q) conjugate exponents, to be chosen later. Observe that if we apply (Equation4) for u0Hsν2(G), b=sν2, a = 0, and q0=2QQνs, then q~0=2, and we have (22) u0Lq0(G)Rs2u0L2(G)<.(22) Choosing 2q=q0 in (Equation21) so that q=QQνs, we get q=Qνs, so that (23) mu0L2(G)2mLQνs(G)Rs2u0L2(G)2<,(23) and by (Equation20) we can estimate (24) m()u(t,)L2(G)2u1L2(G)2+u0Hsν2(G)2+mu0L2(G)2u1L2(G)2+u0Hsν2(G)2+mLQνs(G)u0Hsν2(G)21+mLQνs(G)u1L2(G)2+u0Hsν2(G)2,(24) uniformly in t[0,T]. Additionally, (Equation20), under the estimate (Equation24), implies (25) ut(t,)L2(G)2,Rs2u(t,)L2(G)21+mLQνs(G)u1L2(G)2+u0Hsν2(G)2.(25) To show our claim (Equation19), it suffices to show the desired estimate for the solution norm u(t,)L2(G). First we observe that by the Sobolev embeddings (Equation4) and Hölder's inequality, using (Equation23) with m instead of m, and m2LQνs(G)=mL2Qνs(G)2, one obtains mu(t,)L2(G)2mL2Qνs2Rs2u(t,)L2(G)2, where the last combined with (Equation25) yields (26) mu(t,)L2(G)2m2Qνs21+mLQνs(G)u1L2(G)2+u0Hsν2(G)2.(26) Finally, using arguments similar to those we developed in Proposition 3.1, together with the estimate (Equation26) we get u(t,)L2(G)2u0L2(G)2+u1L2(G)2+m()u(t,)L2(G)2u1L2(G)2+u0Hsν2(G)21+m2Qνs21+mLQνs(G), uniformly in t[0,T]. The uniqueness of u is immediate by the estimate (Equation19), and this finishes the proof of Proposition 3.2.

4. Existence and uniqueness of the very weak solution

Here, we consider the case where the mass-term in (Equation1) satisfies some moderateness properties. The latter is satisfied in cases where, for instance, m has strong singularities, namely when m=δ or δ2. This follows by Proposition 4.8 for δ, while we can understand δ2 as an approximating family or in the Colombeau sense.

Definition 4.1

Moderateness

  1. Let X be a normed space of functions on G. A net of functions (fϵ)ϵX is said to be X-moderate if there exists NN such that fϵXϵN, uniformly in ϵ(0,1].

  2. A net of functions (uϵ)ϵ in C([0,T];Hsν2(G))C1([0,T];L2(G)) is said to be C([0,T];Hsν2(G))C1([0,T];L2(G))-moderate, or for brevity, C1-moderate, if there exists NN such that supt[0,T]{u(t,)Hsν2(G)+tu(t,)L2(G)}ϵN, uniformly in ϵ(0,1].

Definition 4.2

Negligibility

Let Y be a normed space of functions on G. Let (fϵ)ϵ, (f~ϵ)ϵ be two nets. Then, the net (fϵf~ϵ)ϵ is called Y-negligible, if the following condition is satisfied (27) fϵf~ϵYϵk,(27) for all kN, ϵ(0,1]. In the case where f=f(t,x) is a function also depending on t[0,T], then the negligibility condition (Equation27) can be regarded as fϵ(t,)f~ϵ(t,)Yϵk, kN, uniformly in t[0,T]. The constant in the inequality (Equation27) can depend on k but not on ϵ.

In Definitions 4.3 and 4.6, we introduce the notion of the unique very weak solution to the Cauchy problem (Equation1). Our definitions are similar to the one introduced in [Citation2], but here we measure moderateness and negligibility in terms of Lp(G) or C1-seminorms rather than in terms of Gevrey-seminorms.

Definition 4.3

Very weak solution

Let (u0,u1)Hsν2(G)×L2(G). Then, if there exists a non-negative L(G)-moderate (or a L2Qνs(G)LQνs(G)-moderate, if we require to have Q>νs) approximating net (mϵ)ϵ, mϵ0, to m, so that the family (uϵ)ϵC([0,T];Hsν2(G))C1([0,T];L2(G)) which solves the ϵ-parametrised problem (28) t2uϵ(t,x)+Rsuϵ(t,x)+mϵ(x)uϵ(t,x)=0,(t,x)[0,T]×G,uϵ(0,x)=u0,ϵ(x),tuϵ(0,x)=u1,ϵ(x),xG,(28) for all ϵ(0,1], is C1-moderate, then net (uϵ)ϵ is said to be a very weak solution to the Cauchy problem (Equation1).

Remark 4.4

In Definition 4.3 above we ask for mϵ to approximate m, to allow for more flexibility. This should be understood as follows. If mD(G) is a distribution, we can understand it as a regularisation, namely, the assumption in Definition 4.3 is that there is a Friedrichs mollifier ψ0 such that mϵ=mψϵ. However, the word approximation allows for more flexibility, for example, we can in principle generate an approximating family with a net m~ϵ such that the one we will discuss in (Equation31). Moreover, this context allows us to start with m being more singular than a distribution: for example, if m=δ2 we can think of an approximating family mϵ=ψϵ2. See also Remark 4.7 for a continuation of this discussion.

We now formulate the very weak existence result, corresponding to two possibilities of regularising with families (mϵ)ϵ with different properties, corresponding to the existence results in Propositions 3.1 and 3.2.

Theorem 4.5

Let (u0,u1)Hsν2(G)×L2(G). Then the Cauchy problem (Equation1) has a very weak solution.

Proof.

Let u0,u1 be as in the hypothesis. If (mϵ)ϵ is L(G)-moderate (or L2Qνs(G)LQνs(G)-moderate), then, since mϵ0, by using (Equation6) (or (Equation19), respectively) we get uϵ(t,)Hsν2(G)+tuϵ(t,)L2(G)ϵN,NN, for all t[0,T] and for any ϵ(0,1]. This means that the family of solutions (uϵ)ϵ is C1-moderate, and completes the proof of Theorem 4.5.

The uniqueness of the very weak solution to the Cauchy problem (Equation1) can be understood as if a negligible change of the net (mϵ)ϵ does not affect the asymptotic behaviour of the family of the very weak solutions. In other words, negligible changes of the approximation mϵ of m lead to negligible changes in the solution family uϵ, with an appropriate choices of norms to understand the negligibility. Formally, we have the following definition.

Definition 4.6

Let X and Y be normed spaces of functions on G. We say that the Cauchy problem (Equation1) has an (X,Y)-unique very weak solution, if for all X-moderate nets mϵ0,m~ϵ0, such that (mϵm~ϵ)ϵ is Y-negligible, it follows that uϵ(t,)u~ϵ(t,)L2(G)CNϵN, NN, uniformly in t[0,T], and for all ϵ(0,1], where (uϵ)ϵ and (u~ϵ)ϵ are the families of solutions corresponding to the ϵ-parametrised problems (29) t2uϵ(t,x)+Rsuϵ(t,x)+mϵ(x)uϵ(t,x)=0,(t,x)[0,T]×G,uϵ(0,x)=u0,ϵ(x),tuϵ(0,x)=u1,ϵ(x),xG,(29) and (30) t2u~ϵ(t,x)+Rsu~ϵ(t,x)+m~ϵ(x)u~ϵ(t,x)=0,(t,x)[0,T]×G,u~ϵ(0,x)=u~0,ϵ(x),tu~ϵ(0,x)=u~1,ϵ(x),xG,(30) respectively.

Remark 4.7

We note that in Definition 3 in the previous paper [Citation9], the word ‘regularisation’ needs to be understood, in general, as an approximation not necessarily depending on the classical convolution and specific mollifiers. In this case, our definition of the uniqueness of the very weak solutions here includes also the version in Definition 3 in [Citation9], but Definition 4.6 makes it more rigorous. To clarify this further, we can take, for example, mϵ to be a regularisation of m by a convolution (if m is a distribution), and take (31) m~ϵ=mϵ+e1/ϵ.(31) Then the net (mϵm~ϵ)ϵ is L-negligible, and so it is suitable to be used in Definition 4.6. If m=δ2, we can take mϵ=ψϵ2 for a Friedrichs mollifier ψ, and still, for example, m~ϵ as in (Equation31). We also note that Definition 4.6 can be also interpreted as stability. In fact, in Definition 4.6 we do not assume mϵ to approximate m since we can prove the required property without this assumption (as in Theorems 4.9 and 4.10). This allows for our results to be applicable to cases like m=δ2, since with this approach we do not need to explain in which sense mϵ=ψϵ2 approximates m=δ2.

We now give some clarification of the moderateness assumption of the regularisations (or approximations). Let us underline that, the global structure of E-distributions, implies that the assumption on the Lp-moderateness, for p[1,], is natural for nets that arise as regularisations of a compactly supported distribution in E via convolutions with a mollifier as in (Equation5).

Proposition 4.8

Let vE(G), and let vϵ=vψϵ be obtained as the convolution of v with a mollifier ψϵ as in (Equation5). Then the regularising net (vϵ)ϵ is Lp(G)-moderate for any p[1,].

Proof.

Recall, that for vE(G) we can find mN and compactly supported continuous functions fβC(G) such that v=|β|mβfβ, where |β| denoted the length of the multi-index β. Considering the convolution of v with a mollifier ψϵ as in (Equation5) yields vϵ=vψϵ=|β|mβfβψϵ=|β|mβfβψϵ, where each term in the above sum can be rewritten as βfβψϵ=βfβ(yx1),ψϵ(x)=(1)|β|fβ(yx1),βψϵ(x)=(1)|β|ϵQfβ,βψ(Dϵ1(x)=(1)|β|ϵQ[β]fβ,(βψ)(Dϵ1(x), where [β] stands for the homogeneous length of β.

Finally, since fβ,ψ are compactly supported, we get fβ,(βψ)(Dϵ1)Lp(G), for all p, and this finishes the proof of Proposition 4.8.

We note that thanks to Proposition 4.8 the assumption of Theorem 4.5 can be relaxed to (u0,u1){Hsν2(G)E(G)}×{L2(G)E(G)}. The following theorems show the uniqueness of the very weak solution to the Cauchy problem (Equation1) under different assumptions on the nets (mϵ)ϵ.

Theorem 4.9

Suppose that (u0,u1){Hsν2(G)E(G)}×{L2(G)E(G)}. Then the very weak solution to the Cauchy problem (Equation1) is (L(G),L(G))-unique.

Proof.

Let (uϵ)ϵ and (u~ϵ)ϵ be the families of solutions corresponding to the Cauchy problems (Equation29) and (Equation30), respectively. If we denote by Uϵ(t,):=uϵ(t,)u~ϵ(t,), then Uϵ satisfies (32) t2Uϵ(t,x)+RsUϵ(t,x)+mϵ(x)Uϵ(t,x)=fϵ(t,x),(t,x)[0,T]×G,Uϵ(0,x)=0,tUϵ(0,x)=0,xG,(32) where fϵ(t,x):=(m~ϵ(x)mϵ(x))u~ϵ(t,x).

The solution of the Cauchy problem (Equation32) can be expressed in terms of the solution to the corresponding homogeneous Cauchy problem using Duhamel's principle. Indeed, if for a fixed σ, Vϵ(t,x;σ) is the solution of the homogeneous problem (33) t2Vϵ(t,x;σ)+RsVϵ(t,x;σ)+mϵVϵ(t,x;σ)=0,in (σ,T]×G,Vϵ(t,x;σ)=0,tVϵ(t,x;σ)=fϵ(σ,x),on {t=σ}×G,(33) then Uϵ is given by Uϵ(t,x)=0tVϵ(t,x;σ)dσ.

Since by Minkowski's integral inequality we know 0tVϵ(t,;σ)dσL2(G)0tVϵ(t,;σ)L2(G)dσ, using the energy estimate (Equation6) to control L2(G)-norm of the solution Vϵ to the homogeneous problem (Equation33), and subsequently of Uϵ, we get Uϵ(t,)L2(G)0TVϵ(t,;σ)L2(G)dσ(1+mϵL(G))0Tfϵ(σ,)L2(G)dσ(1+mϵL(G))m~ϵmϵL(G)0Tu~ϵ(σ,)L2(G)dσ, where we use the estimate fϵ(σ,)L2(G)=(m~ϵmϵ)()u~ϵ(σ,)L2(G)m~ϵmϵL(G)u~ϵ(σ,)L2(G). Now, using the fact that (mϵ)ϵ is L(G)-moderate, while also that the net (u~ϵ)ϵ, as being a very weak solution to the Cauchy problem (Equation29), is C1-moderate and that (mϵm~ϵ)ϵ is L-negligible, we get that Uϵ(t,)L2(G)ϵN1+N0TϵN2dσ=TϵN1N2+N, for some N1,N2N, and for all NN, ϵ(0,1]. That is, we have Uϵ(t,)L2(G)ϵk, for all kN, and the last shows that the net (uϵ)ϵ is the unique very weak solution to the Cauchy problem (Equation1).

Alternative to Theorem 4.9 conditions on the nets (mϵ)ϵ,(m~ϵ)ϵ that guarantee the very weakly well-posedness of (Equation1) are given in the following theorem.

Theorem 4.10

Let Q>νs, and suppose that (u0,u1){Hsν2(G)E(G)}×{L2(G)E(G)}. Then the very weak solution to the Cauchy problem (Equation1) is (L(G),L2Qνs(G))-unique. Moreover, the very weak solution to the Cauchy problem (Equation1) is also (L2Qνs(G)LQνs(G),L2Qνs(G))-unique and (L2Qνs(G)LQνs(G),L(G))-unique.

Proof.

We will only prove the (L(G),L2Qνs(G))-uniqueness as the other two uniqueness statements are similar. Proceeding as we did in the proof of Theorem 4.9, we arrive at Uϵ(t,)L2(G)(1+mϵL(G))0Tfϵ(σ,)L2(G)dσ=(1+mϵL(G))0T(m~ϵmϵ)()u~ϵ(σ,)L2(G)dσ. for all t[0,T]. Additionally, by applying Hölder's inequality, together with the Sobolev embeddings (Equation4), we have (m~ϵmϵ)()u~ϵ(t,)L2(G)m~ϵmϵL2Qνs(G)Rs2u~ϵ(t,)L2(G), where since (u~ϵ), as being the very weak solution corresponding to the L(G)-moderate net (m~ϵ)ϵ, is C1-moderate, we have Rs2u~ϵ(t,)L2(G)ϵN1,forsome N1N. Summarising the above, and since m~ϵmϵL2Qνs(G)ϵN, NN, we obtain Uϵ(t,)L2(G)ϵk, kN, uniformly in t, and this finishes the proof of Theorem 4.10.

5. Consistency of the very weak solution with the classical one

The next theorems stress the conditions, on the coefficient m and on the initial data u0,u1, under which, the classical solution to the Cauchy problem (Equation1) can be recaptured by its very weak solution. In the statements below, we understand the classical solutions as those given by Proposition 3.1 or Proposition 3.2, depending on the assumptions. By the ‘regularisations’ mϵ=mψϵ below we understand the convolution with non-negative Friedrichs mollifiers ψ0.

Theorem 5.1

Let Q>νs. Consider the Cauchy problem (Equation1), and let (u0,u1)Hsν2(G)×L2(G). Assume also that mL2Qνs(G)LQνs(G), m0, and that (mϵ)ϵ, is a regularisation of the coefficient m. Then the regularised net (uϵ)ϵ converges, as ϵ0, in L2(G) to the classical solution u given by Proposition 3.2.

Proof.

Let u be the classical solution of (Equation1) given by Proposition 3.2, and let (uϵ) be the very weak solution of the regularised analogue of it as in (Equation28). Then, we get t2(uuϵ)(t,x)+Rs(uuϵ)(t,x)+mϵ(x)(uuϵ)(t,x)=ηϵ(t,x),(uuϵ)(0,x)=0,t(uuϵ)(0,x)=0, where (t,x)[0,T]×G, and ηϵ(t,x):=(mϵ(x)m(x))u(t,x). If we denote by Uϵ the difference Uϵ(t,x):=(uuϵ)(t,x), the above can be rewritten equivalently as (34) t2Uϵ(t,x)+RsUϵ(t,x)+mϵ(x)Uϵ(t,x)=ηϵ(t,x),Uϵ(0,x)=0,tUϵ(0,x)=0.(34) Therefore, if we denote by Wϵ(t,x;σ) the solution to the corresponding homogeneous problem with the initial data at {t=σ}×G Wϵ(t,x;σ)=0,andtWϵ(t,x;σ)=ηϵ(σ,x), then by Proposition 3.2 we get Wϵ(t,;σ)L2(G)(1+mϵL2Qνs(G))1+mϵLQνs(G)1/2ηϵ(σ,)L2(G)(1+mϵL2Qνs(G))1+mϵLQνs(G)1/2mϵmL2Qνs(G)×Rs2u(σ,)L2(G), uniformly in t[σ,T] and σ[0,T], where we apply Hölder's inequality and the Sobolev embeddings (Equation4). Since mL2Qνs(G), we have mϵmL2Qνs(G)0, so that taking the limit of the above as ϵ0, we get (35) Wϵ(t,;σ)L2(G)0,(35) uniformly in t[σ,T] and σ[0,T]. Now, Duhamel's principle allows us to express the solution to the inhomogeneous problem with respect to the homogeneous one as (36) Uϵ(t,x)=0tWϵ(t,x;σ)dσ,(36) so that, by (Equation35), (Equation36), and Minkowski's integral inequality 0tWϵ(t,;σ)dσL2(G)0tWϵ(t,;σ)L2(G)dσ, we obtain Uϵ(t,)L2(G)Tsupσ[0,T]Wϵ(t,;σ)L2(G)0,asϵ0. This means that uϵu with respect to L2(G)-norm, and this finishes the proof of Theorem 5.1.

In the following theorem we denote by C0(G) the space of continuous functions on G vanishing at infinity, that is, such that for every ϵ>0 there exists a compact set K outside of which we have |f|<δ. We also denote by B(G) the space of simple and compactly supported functions on G. Both C0(G) and B(G), if endowed with the norm L(G), are Banach spaces.

Theorem 5.2

Consider the Cauchy problem (Equation1), and let (u0,u1)Hsν2(G)×L2(G). Assume also that mC0(G)B(G), m0, and that (mϵ)ϵ, mϵ0, is a regularisation of the coefficient m. Then the regularised net (uϵ)ϵ converges, as ϵ0, in L2(G) to the classical solution u given by Proposition 3.1.

Before giving the proof of Theorem 5.2, let us make the following observation: If mC0(G)B(G), then mϵL(G)C<, uniformly in ϵ(0,1].

Proof

Proof of Theorem 5.2.

First observe that for m, (mϵ)ϵ as in the hypothesis, we have mϵL(G) for each ϵ(0,1]. Now, as in (Equation34), if we denote by Wϵ the solution to the problem t2Wϵ(t,x;σ)+RsWϵ(t,x;σ)+mϵ(x)Wϵ(t,x;σ)=0,Wϵ(t,x;σ)=0,tWϵ(t,x;σ)=ηϵ(σ,x)on {t=σ}×G, where ηϵ(t,x):=(mϵ(x)m(x))u(t,x), then by Proposition 3.1 we obtain Wϵ(t,;σ)L2(G)(1+mϵL(G))ηϵ(σ,)L2(G)(1+mϵL(G))mϵmL(G)u(σ,)L2(G), uniformly in t[σ,T] and σ[0,T]. Now, by Lemmas 3.1.58 and 3.1.59 in [Citation14] we have mϵmL(G)0,as ϵ0, so that by the above we get (37) Wϵ(t,;σ)L2(G)0,as ϵ0,(37) uniformly in t[σ,T] and σ[0,T]. Finally, by Duhamel's principle if Uϵ is the solution to the non-homogeneous problem (Equation34), then by (Equation37) we get Uϵ(t,)L2(G)0, and this completes the proof of Theorem 5.2.

Remark 5.3

We note that in Theorem 4 in the paper [Citation9], one wrote the assumption that mL(Rd) in the consistency result. This may be not sufficient in general. Indeed, to be more accurate, it is better to ask m to be in the subspace C0(Rd)B(Rd) of L(Rd). In this way we obtain a correction to the statement of Theorem 4 in [Citation9] as a special case of Theorem 5.2 with G=Rd and R being the positive Laplacian Δ.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The authors are supported by the Fonds Wetenschappelijk Onderzoek (FWO) Odysseus 1 grant G.0H94.18N: Analysis and Partial Differential Equations and by the Methusalem programme of the Ghent University Special Research Fund (BOF) (Grant number 01M01021). Michael Ruzhansky is also supported by Engineering and Physical Sciences Research Council (EPSRC) grant EP/R003025/2.

Notes

1 When p = 2, we will write R2=R for the self-adjoint extension of R on L2(G).

References

  • Schwartz L. Sur l'impossibilite de la multiplication des distributions. C R Acad Sci Paris. 1954;239:847–848.
  • Garetto C, Ruzhansky M. Hyperbolic second order equations with non-regular time dependent coefficients. Arch Ration Mech Anal. 2015;217:113–154.
  • Altybay A, Ruzhansky M, Tokmagambetov N. Wave equation with distributional propagation speed and mass term: numerical simulations. Appl Math E-Notes. 2019;19:552–562.
  • Munoz JC, Ruzhansky M, Tokmagambetov N. Wave propagation with irregular dissipation and applications to acoustic problems and shallow water. J Math Pures Appl. 2019;123:127–147.
  • Ruzhansky M, Tokmagambetov N. Very weak solutions of wave equation for Landau Hamiltonian with irregular electromagnetic field. Lett Math Phys. 2017;107:591–618.
  • Ruzhansky M, Tokmagambetov N. Wave equation for operators with discrete spectrum and irregular propagation speed. Arch Ration Mech Anal. 2017;226:1161–1207.
  • Ruzhansky M, Taranto C. Time-dependent wave equations on graded groups. Acta Appl Math. 2021;171:Article ID 21. 24pp.
  • Ruzhansky M, Yessirkegenov N. Very weak solutions to hypoelliptic wave equations. J Differ Equ. 2020;268:2063–2088.
  • Altybay A, Ruzhansky M, Sebih ME, et al. Fractional Klein-Gordon equation with singular mass. Chaos Solitons Fractals. 2021;143:Article ID 110579.
  • Altybay A, Ruzhansky M, Sebih ME, et al. Fractional Schrödinger equations with singular potentials of higher-order. Rep Math Phys. 2021;87:129–144.
  • Altybay A, Ruzhansky M, Sebih ME, et al. The heat equation with strongly singular potentials. Appl Math Comput. 2021;399:Article ID 126006.
  • Garetto C. On the wave equation with multiplicities and space-dependent irregular coefficients. Trans Amer Math Soc. 2020. To appear. arXiv:2004.09657.
  • Folland GB, Stein EM. Hardy spaces on homogeneous groups. Princeton (NJ)/Tokyo: Princeton University Press/University of Tokyo Press; 1982. (Mathematical notes; vol. 28).
  • Fischer V, Ruzhansky M. Quantizations on Nilpotent Lie groups. Birkhäuser/Springer; 2016. (Progress in mathematics; vol. 314).
  • Dixmier J. Sur les représentations unitaires des groupes de Lie nilpotents. Canad J Math. 1957;10:321–348.
  • Taylor ME. Noncommutative microlocal analysis. I. Vol. 52. Mem. Math. Soc.; 1984.
  • Chatzakou M. On (λ,μ)-classes on the Engel group. In: Georgiev V, Ozawa T, Ruzhansky M, Wirth J, editors. Advances in harmonic analysis and partial differential equations. Birkhauser/Springer; 2020. p. 37–49. (Trends math.).
  • Chatzakou M. Quantizations on the Engel and the Cartan groups. J Lie Theory. 2021;31:517–542.
  • Corwin LJ, Greenleaf FP. Representations of nilpotent Lie groups and their applications. Part I: basic theory and examples. Cambridge: Cambridge University Press; 1990. (Cambridge studies in advanced mathematics; vol. 18).
  • Birkhoff G. Representability of Lie algebras and Lie groups by matrices. Ann Math. 1937;38:526–532.
  • Helffer B, Nourrigat J. Caracterisation des opérateurs hypoelliptiques homogénes invariants á gauche sur un groupe de Lie nilpotent gradué. Comm Partial Differ Equ. 1979;4:899–958.
  • Beals R. Opérateurs invariants hypoelliptiques sur un groupe de Lie nilpotent. In: Séminarie Goulaouic-Schwartz 1976/1977: Équations aux dérivées partielles at analyse fonctionnelle, Exp. No. 19; 1977. p. 8.
  • Rockland C. Hypoellipticity on the Heisenberg group-representation-theoretic criteria. Trans Amer Math Soc. 1978;240:1–52.
  • ter Elst AFM, Robinson M. Spectral estimates for positive Rockland operators. In: Algebraic groups and Lie groups. Cambridge: Cambridge Univ. Press; 1997. p. 195–213. (Austral. math. soc. lect. ser.; vol. 9).
  • Hulanicki A, Jenkins JW, Ludwig J. Minimum eigenvalues for positive, Rockland operators. Proc Amer Math Soc. 1985;94:718–720.
  • Dixmier J. C-algebras. Translated from French by Francis Jellet. Vol. 15. North-Holland Mathematical Library; 1977.
  • Evans LC. Partial Differential Equations. American Mathematical Society; 1998.