Home Carleman estimates for singular parabolic equations with interior degeneracy and non-smooth coefficients
Article Publicly Available

Carleman estimates for singular parabolic equations with interior degeneracy and non-smooth coefficients

  • Genni Fragnelli and Dimitri Mugnai ORCID logo EMAIL logo
Published/Copyright: February 16, 2016

Abstract

We establish Carleman estimates for singular/degenerate parabolic Dirichlet problems with degeneracy and singularity occurring in the interior of the spatial domain. Our results are completely new, since this situation is not covered by previous contributions for degeneracy and singularity on the boundary. In addition, we consider non-smooth coefficients, thus preventing the use of standard calculations in this framework.

1 Introduction

Controllability issues for parabolic problems have been a mainstream topic in recent years, and several developments have been pursued: starting from the heat equation in bounded and unbounded domain, related contributions have been found for more general situations. A common strategy in showing controllability results is to prove that certain global Carleman estimates hold true for the operator which is the adjoint of the given one.

In this paper we focus on a class of singular parabolic operators with interior degeneracy of the form

(1.1)ut-(a(x)ux)x-λb(x)u,

associated to Dirichlet boundary conditions and with (t,x)QT:=(0,T)×(0,1), T>0 being a fixed number. Here a and b degenerate at the same interior point x0(0,1), and λ satisfies suitable assumptions (see condition (2.4) below). The fact that both a and b degenerate at x0 is just for the sake of simplicity and shortness. All the stated results are still valid if they degenerate at different points. The prototypes we have in mind are a(x)=|x-x0|K1 and b(x)=|x-x0|K2 for some K1,K2>0. The main goal is to establish global Carleman estimates for operators of the form given in (1.1).

Such estimates for uniformly parabolic operators without degeneracies or singularities have been largely developed (see, e.g., [32]). Recently, these estimates have been also studied for operators which are not uniformly parabolic. Indeed, as pointed out by several authors, many problems coming from Physics (see [36]), Biology (see [21]) and Mathematical Finance (see [35]) are described by degenerate parabolic equations. In particular, new Carleman estimates (and consequently null controllability properties) were established in [1], and also in [15, 40], for the operator

ut-(aux)x+c(t,x)u,(t,x)QT,

where a(0)=a(1)=0, aC1(0,1) and cL(QT) (see also [13, 12, 25] for problems in non-divergence form).

An interesting situation is the case of parabolic operators with singular inverse-square potentials. First results in this direction were obtained in [46] for the non-degenerate singular potentials with heat-like operator

(1.2)ut-Δu-λ1|x|2u,(t,x)(0,T)×Ω,

with associated Dirichlet boundary conditions in a bounded domain ΩN containing the singularity x=0 in the interior (see also [47] for the wave and Schrödinger equations and [16] for boundary singularity). Similar operators of the form

ut-Δu-λ1|x|K2u,(t,x)(0,T)×Ω,

arise, for instance, in quantum mechanics (see, e.g., [4, 19]), or in combustion problems (see, e.g., [6, 10, 20, 33]), and is known to generate interesting phenomena. For example, in [4] and in [5] it was proved that, for all values of λ, global positive solutions exist if K2<2, whereas instantaneous and complete blow-up occurs if K2>2. In the critical case, i.e., K2=2, the value of the parameter λ determines the behavior of the equation. If λ1/4 (which is the optimal constant of the Hardy inequality, see [9]) global positive solutions exist, while, if λ>1/4, instantaneous and complete blow-up occurs (for other comments on this argument we refer to [45]). We recall that in [46], Carleman estimates were established for (1.2) under the condition λ1/4. On the contrary, if λ>1/4, in [22] it was proved that null controllability fails.

We remark that the non-degenerate problems studied in [4, 16, 22, 45, 47, 46] cover the multidimensional case, while here we treat the case N=1, like Vancostenoble [45], who studied the operator that couples a degenerate diffusion coefficient with a singular potential. In particular, for K1[0,2) and K22-K1, she established Carleman estimates for the operator

ut-(xK1ux)x-λ1xK2u,(t,x)QT,

unifying the results of [14] and [46] in the purely degenerate operator and in the purely singular one, respectively. This result was then extended in [24] and in [23] to the operators

(1.3)ut-(a(x)ux)x-λ1xK2u,(t,x)QT,

for axK1,K1[0,2) and K22-K1. Here, as before, the function a degenerates at the boundary of the space domain, and Dirichlet boundary conditions are in force.

We remark the fact that all the papers cited so far, with the exception of [22], consider a singular/degenerate operator with degeneracy or singularity appearing at the boundary of the domain. For example, in (1.3) as a one can also consider the double power function

a(x)=xk(1-x)κ,x[0,1],

where k and κ are positive constants. To the best of our knowledge, [8, 29, 30] are the first papers dealing with Carleman estimates (and, consequently, null controllability) for operators (in divergence and in non-divergence form with Dirichlet or Neumann boundary conditions) with mere degeneracy at the interior of the space domain (for related systems of degenerate equations we refer to [7]). We also recall [28] and [27] for other type of control problems associated to parabolic operators with interior degeneracy in divergence and non-divergence form, respectively.

We emphasize the fact that an interior degeneracy does not imply a simple adaptation of previous results and of the techniques used for boundary degeneracy. Indeed, imposing homogeneous Dirichlet boundary conditions, in the latter case one knows a priori that any function in the reference functional space vanishes exactly at the degeneracy point. Now, since the degeneracy point is in the interior of the spatial domain, such information is not valid anymore, and we cannot take advantage of this fact.

For this reason, the present paper is devoted to study the operator defined in (1.1), that couples a general degenerate diffusion coefficient with a general singular potential with degeneracy and singularity at the interior of the space domain. In particular, under suitable conditions on all the parameters of the operator, we establish Carleman estimates and, as a consequence, null controllability for the associated generalized heat problem. Clearly, this result generalizes the one obtained in [29, 30]. In fact, if λ=0 (that is, if we consider the purely degenerate case), we recover the main contributions therein. See also [26] for the problem in non-divergence form for both Dirichlet and Neumann boundary conditions.

We also remark the fact that, though we have in mind prototypes as power functions for the degeneracy and the singularity, we don not limit our investigation to these functions, which are analytic out of their zero. Indeed, in this paper, pure powers singularities and degeneracies are considered only as a by-product of our main results, which are valid for non-smooth general coefficients. This is quite a new view-point when dealing with Carleman estimates, since in this framework it is natural to assume that all the coefficients in force are quite regular. However, though this strategy has been successful for years, it is clear that also more irregular coefficients can be considered and appear in a natural way (for instance, see [34, 37]). Nevertheless, it will be clear from the proof that Carleman estimates do hold without particular conditions also in the non-smooth setting, while for observability (and thus controllability) another technical condition is needed; however, such a condition is trivially true for the prototypes.

For this reason, for the first time to our best knowledge, in [30] non-smooth degenerate coefficients were treated. Continuing in this direction, here we consider operators which contain both degenerate and singular coefficients, as in [24, 23, 45], but with low regularity.

The classical approach to study singular operators in dimension 1 relies in the validity of the Hardy–Poincaré inequality

(1.4)01u2x2𝑑x401(u)2𝑑x,

which is valid for every uH1(0,1) with u(0)=0. Similar inequalities are the starting point to prove well-posedness of the associated problems in the Sobolev spaces under consideration. In our situation, we prove an inequality related to (1.4), but with a degeneracy coefficient in the gradient term. Such an estimate is valid in a suitable Hilbert space we shall introduce below, and it states the existence of C>0 such that for all u, we have

01u2b𝑑xC01a(u)2𝑑x.

This inequality, which is related to another weighted Hardy–Poincaré inequality (see Proposition 2.7), is the key step for the well-posedness of (1.5). Once this is done, global Carleman estimates follow, provided that an ad hoc choice of the weight functions is made (see Theorem 3.3).

The introduction of the space (which may coincide with the usual Sobolev space in some cases) is another feature of this paper, which is completely new with respect to all the previous approaches. Including the integrability of u2/b in the definition of has the advantage of obtaining immediately some useful functional properties, that in general could be hard to show in the usual Sobolev spaces. Indeed, solutions were already found in suitable function spaces for the “critical” and “supercritical” cases (when λ equals or exceeds the best constant in the classical Hardy–Poincaré inequality) in [47] and [48] for purely singular problems. However, as already done in the purely degenerate case (see [1, 7, 8, 13, 12, 14, 24, 23, 25, 29, 30, 31]), a weighted Sobolev space must be used. For this reason, we believe that it is natural to unify these approaches in the singular/degenerate, as we do.

Now, let us consider the evolution problem

(1.5){ut-(aux)x-λb(x)u=h(t,x)χω(x),(t,x)QT,u(t,0)=u(t,1)=0,t(0,T),u(0,x)=u0(x),x(0,1),

where u0L2(0,1), the control hL2(QT) acts on a non-empty interval ω(0,1) and χω denotes the characteristic function of ω.

As usual, we say that problem (1.5) is null controllable if there exists hL2(QT) such that u(T,x)0 for x[0,1]. A common strategy to show that (1.5) is null controllable is to prove Carleman estimates for any solution v of the adjoint problem of (1.5)

{vt+(avx)x+λb(x)v=0,(t,x)QT,v(t,0)=v(t,1)=0,t(0,T),v(T,x)=vT(x),

and then deduce an observability inequality of the form

(1.6)01v2(0,x)𝑑xCT0Tωv2(t,x)𝑑x𝑑t,

where CT>0 is a universal constant. In the non-degenerate case this has been obtained by a well-established procedure using Carleman and Caccioppoli inequalities. In our singular/degenerate non-smooth situation, we need a new suitable Caccioppoli inequality (see Proposition 4.6), as well as global Carleman estimates in the non-smooth non-degenerate and non-singular case (see Proposition 4.8), which will be used far away from x0 within a localization procedure via cut-off functions. Once these tools are established, we will be able to prove an observability inequality like (1.6), and then controllability results for (1.5). However, we cannot do that in all cases, since we have to exclude that both the degeneracy and the singularity are strong, see condition (SSD) below.

Finally, we remark that our studies with non-smooth coefficients are particularly useful. In fact, though null controllability results could be obtained also in other ways, for example by a localization technique (at least when x0ω), in [30] it is shown that with non-smooth coefficients, even when λ=0, this is not always the case. For this, our approach with observability inequalities is very general and permits to cover more involved situations.

The paper is organized as follows. In Section 2, we study the well-posedness of problem (1.5), giving some general tools that we shall use several times. In Section 3, we provide one of the main results of this paper, i.e., Carleman estimates for the adjoint problem to (1.5). In Section 4, we apply the previous Carleman estimates to prove an observability inequality, which, together with a Caccioppoli type inequality, lets us derive new null controllability results for the associated singular/degenerate problem, also when the degeneracy and the singularity points are inside the control region.

A final comment on the notation: by c or C we shall denote universal positive constants, which are allowed to vary from line to line.

2 Well-posedness

The ways in which a and b degenerate at x0 can be quite different, and for this reason we distinguish four different types of degeneracy. In particular, we consider the following cases.

Hypothesis 2.1

Hypothesis 2.1 (Doubly weakly degenerate case (WWD))

There exists x0(0,1) such that a(x0)=b(x0)=0, a,b>0 on [0,1]{x0}, a,bW1,1(0,1), and there exist K1,K2(0,1) such that (x-x0)aK1a and (x-x0)bK2b a.e. in [0,1].

Hypothesis 2.2

Hypothesis 2.2 (Weakly-strongly degenerate case (WSD))

There exists x0(0,1) such that a(x0)=b(x0)=0, a,b>0 on [0,1]{x0}, aW1,1(0,1), bW1,(0,1), and there exist K1(0,1) and K21 such that (x-x0)aK1a and (x-x0)bK2b a.e. in [0,1].

Hypothesis 2.3

Hypothesis 2.3 (Strongly-weakly degenerate case (SWD))

There exists x0(0,1) such that a(x0)=b(x0)=0, a,b>0 on [0,1]{x0}, aW1,(0,1), bW1,1(0,1), and there exist K11 and K2(0,1) such that (x-x0)aK1a and (x-x0)bK2b a.e. in [0,1].

Hypothesis 2.4

Hypothesis 2.4 (Doubly strongly degenerate case (SSD))

There exists x0(0,1) such that a(x0)=b(x0)=0, a,b>0 on [0,1]{x0}, a,bW1,(0,1), and there exist K1,K21 such that (x-x0)aK1a and (x-x0)bK2b a.e. in [0,1].

Typical examples for the previous degeneracies and singularities are a(x)=|x-x0|K1 and b(x)=|x-x0|K2, with 0<K1,K2<2.

Remark 2.5

The restriction Ki<2 is related to the controllability issue. Indeed, it is clear from the proof of Theorem 2.22 that such a condition is useless, for example, when λ<0. On the other hand, concerning controllability, we will not consider the case Ki2, since if a(x)=|x-x0|K1, K12 and λ=0, by a standard change of variables (see [30]), problem (1.5) may be transformed in a non-degenerate heat equation on an unbounded domain, while the control remains distributed in a bounded domain. This situation is now well-understood, and the lack of null controllability was proved by Micu and Zuazua in [41].

We will use the following result several times; we state it for a, but an analogous one holds for b replacing K1 with K2.

Lemma 2.6

Lemma 2.6 ([29, Lemma 2.1])

Assume that there exists x0(0,1) such that a(x0)=0, a>0 on [0,1]{x0}, and either of the following holds:

  1. aW1,1(0,1) and there exist K1(0,1) such that (x-x0)aK1a a.e. in [0,1],

  2. aW1,(0,1) and there exist K1[1,2) such that (x-x0)aK1a a.e. in [0,1].

Then, the following hold:

  1. For all γK1 the map x|x-x0|γais non-increasing on the left of x=x0 and non-decreasing on the right of x=x0, so that

    limxx0|x-x0|γa=0for all γ>K1.
  2. If K1<1, then 1aL1(0,1).

  3. If K1[1,2), then 1aL1(0,1) and 1aL1(0,1).

For the well-posedness of the problem, we start by introducing the following weighted Hilbert spaces, which are suitable to study all situations, namely the (WWD), (SSD), (WSD) and (SWD) cases:

Ha1(0,1):={uW01,1(0,1):auL2(0,1)},
Ha,b1(0,1):={uHa1(0,1):ubL2(0,1)},

endowed with the inner products

u,vHa1(0,1):=01auv𝑑x+01uv𝑑x,
u,vHa,b1(0,1):=01auv𝑑x+01uv𝑑x+01uvb𝑑x,

respectively.

Note that, if uHa1(0,1), then auL2(0,1), since |au|(max[0,1]a)a|u|.

We recall the following weighted Hardy–Poincaré inequality, see [29, Proposition 2.6].

Proposition 2.7

Assume that pC([0,1]), p>0 on [0,1]{x0}, p(x0)=0, and there exists q>1 such that the following condition is satisfied:

  1. The function xp(x)|x-x0|q is non-increasing on the left of x=x0 and non-decreasing on the right of x=x0.

Then, there exists a constant CHP>0 such that for any function w, locally absolutely continuous on the set [0,x0)(x0,1] and satisfying

w(0)=w(1)=0with 01p(x)|w(x)|2𝑑x<+,

the following inequality holds:

(2.1)01p(x)(x-x0)2w2(x)𝑑xCHP01p(x)|w(x)|2𝑑x.
Remark 2.8

Actually, such a proposition was proved in [29] also requiring q<2. However, as it is clear from the proof, the result is true without such an upper bound on q, that in [29] was used for other estimates.

Moreover, we will also need other types of Hardy inequalities. Let us start with the following crucial one.

Lemma 2.9

If K1+K22 and K2<1, then there exists a constant C>0 such that

(2.2)01u2b𝑑xC01a(u)2𝑑x

for every uHa1(0,1).

Proof.

We set p(x):=(x-x0)2b, so that p satisfies condition (C1) of Proposition 2.7 with q=2-K2>1, by Lemma 2.6. Thus, taken uHa1(0,1), by Proposition 2.7, we get

01u2b𝑑x=01p(x)(x-x0)2u2𝑑xCHP01p(x)|u(x)|2𝑑x.

Now, by Lemma 2.6,

p(x)=(x-x0)2-K1-K2a(x)(x-x0)K1a(x)(x-x0)K2b(x)ca(x)

for some c>0, and the claim follows. ∎

Remark 2.10

A similar proof shows that, when K1+2K22 and K2<1/2, then

01u2b2𝑑xC01a(u)2𝑑x

for every uHa1(0,1).

Lemma 2.9 implies that Ha1(0,1)=Ha,b1(0,1) when K1+K22 and K2<1. However, inequality (2.2) holds in other cases, see Proposition 2.14 below. In order to prove such a proposition, we need a preliminary result.

Lemma 2.11

If K21, then u(x0)=0 for every uHa,b1(0,1).

Proof.

Since uW01,1(0,1), there exists limxx0u(x)=L. If L0, then |u(x)|L2 in a neighborhood of x0, that is

|u(x)|2bL24bL1(0,1),

by Lemma 2.6, and thus L=0. ∎

We also need the following result, whose proof, with the aid of Lemma 2.11, is a simple adaptation of the one given in [31, Lemma 3.2].

Lemma 2.12

If K21, then

Hc1(0,1):={uH01(0,1):suppu(0,1){x0}}

is dense in Ha,b1(0,1).

In the spirit of [18, Lemma 5.3.1], we are now ready for the following “classical” Hardy inequality in the space Ha,b1(0,1) for a(x)=|x-x0|α and b(x)=|x-x0|2-α. However, note that our inequality is more interesting than the classical one, since we admit a singularity inside the interval.

Lemma 2.13

For every αR, the inequality

(1-α)2401u2|x-x0|2-α𝑑x01|x-x0|α(u)2𝑑x

holds true for every uH|x-x0|α,|x-x0|2-α1(0,1).

Proof.

The case α=1 is trivial. So, take β=(1-α)/20 and ε(0,1-x0). First case: β<0 (α>1). In this case we have

x0+ε1(x-x0)α(u)2𝑑x=x0+ε1(x-x0)α((x-x0)β((x-x0)-βu)+β(x-x0)-1u)2𝑑x
β2x0+ε1(x-x0)α-2u2𝑑x+2βx0+ε1(x-x0)α+β-1u((x-x0)-βu)𝑑x
=β2x0+ε1(x-x0)α-2u2𝑑x+β((x-x0)-βu)2|x0+ε1 (since α+β-1=-β)
β2x0+ε1(x-x0)α-2u2𝑑x.

Letting ε0+, we get that

(2.3)x01(x-x0)α(u)2𝑑xβ2x01(x-x0)α-2u2𝑑x.

Second case: β>0. In this situation we have 2-α>1. Thus, in view of Lemma 2.12 with K2=2-α, we will prove (2.3) first if uHc1(0,1) and then, by density, if uH|x-x0|α,|x-x0|2-α1(0,1). Thus, take uHc1(0,1); proceeding as above, we get

x0+ε1(x-x0)α(u)2𝑑xβ2x0+ε1(x-x0)α-2u2𝑑x+β((x-x0)-αu)2|x0+ε1
β2x0+ε1(x-x0)α-2u2𝑑x,

since u(x0+ε)=0 for ε small enough. Passing to the limit as ε0+, and using Lemma 2.12, we get that (2.3) holds true for every uH|x-x0|α,|x-x0|2-α1(0,1).

Operating in a symmetric way on the left of x0, we get the conclusion. ∎

As a corollary of the previous result, we get the following improvement of Lemma 2.9.

Proposition 2.14

If one among Hypotheses 2.12.3 holds with K1+K22, then we have that (2.2) holds for every uHa,b1(0,1).

Proof.

By Lemma 2.6 and Lemma 2.13 with α=2-K2, we immediately get that for every uHa,b1(0,1),

01u2bdxc01u2|x-x0|K2dxc01|x-x0|2-K2(u)2dxc01|x-x0|K1(u)2dxc01a(u)2dx.
Remark 2.15

It is well known that when K1=K2=1, an inequality of the form (2.2) does not hold (see [42]). Being such an inequality fundamental for the observability inequality (see Lemma 4.9), it is no surprise if with our techniques we cannot handle this case in Section 4.

The fundamental space in which we will work is clearly the one where the Hardy–Poincaré-type inequality (2.2) holds. In view of Proposition, it is clear that such a space is

:=Ha,b1(0,1).
Remark 2.16

Under the assumptions of Proposition 2.14, the standard norm 2 is equivalent to

u2:=01a(u)2𝑑x

for all u. Indeed, for all u, we have

01u2𝑑x=01bu2b𝑑xc01a(u)2𝑑x,

and this is enough to conclude the assertion.

Moreover, when λ<0, an equivalent norm is given by

u2:=01a(u)2𝑑x-λ01u2b𝑑x.

This is particularly useful if Hypothesis 2.4 holds (see the proof of Theorem 2.22).

First, let us call C* the best constant of (2.2) in . From now on, we make the following assumptions on a, b and λ.

Hypothesis 2.17

Either of the following holds:

  1. One among Hypotheses 2.12.3 holds true with K1+K22, and we assume that

    (2.4)λ(0,1C*).

  2. Hypotheses 2.1, 2.2, 2.3 or 2.4 hold with λ<0.

Observe that the assumption λ0 is not restrictive, since the case λ=0 was already considered in [29] and in [30].

Using the previous lemmas one can prove the next inequality.

Proposition 2.18

Assume Hypothesis 2.17. Then, there exists Λ(0,1] such that for all uH,

01a(u)2𝑑x-λ01u2b𝑑xΛ01a(u)2𝑑x.

Proof.

If λ<0, the result is obvious taking Λ=1. Now, assume that λ(0,1C*). Then,

01a(u)2𝑑x-λ01u2b𝑑x01a(u)2𝑑x-λC*01a(u)2𝑑xΛ01a(u)2𝑑x.

We recall the following definition.

Definition 2.19

Let u0L2(0,1) and hL2(QT). A function u is said to be a (weak) solution of (1.5) if

uL2(0,T;)H1([0,T];*),

and it satisfies (1.5) in the sense of *-valued distributions.

Note that, by [43, Lemma 11.4], any solution belongs to C([0,T];L2(0,1)).

Finally, we introduce the Hilbert space

Ha,b2(0,1):={uHa1(0,1):auH1(0,1) and AuL2(0,1)},

where

Au:=(au)+λbuwith D(A)=Ha,b2(0,1).
Remark 2.20

Observe that if uD(A), then ub and ubL2(0,1), so that uHa,b1(0,1) and inequality (2.2) holds.

We also recall the following integration by parts with functions in the reference spaces.

Lemma 2.21

Lemma 2.21 (Green’s formula, [31, Lemma 2.3])

Assume one among the Hypotheses 2.12.4. Then, for all (u,v)Ha,b2(0,1)×Ha1(0,1) the following identity holds:

(2.5)01(au)v𝑑x=-01auv𝑑x.

Observe that in the non-degenerate case, it is well known that the heat operator with an inverse-square singular potential

ut-Δu-λu|x|2v

gives rise to well-posed Cauchy–Dirichlet problems if and only if λ is not larger than the best Hardy inequality (see [5, 11, 48]). For this reason, it is not strange that we require an analogous condition for problem (1.5), by invoking Hypothesis 2.17. As a consequence, using the standard semigroup theory, we have that (1.5) is well-posed.

Theorem 2.22

Assume Hypothesis 2.17. For every u0L2(0,1) and hL2(QT), there exists a unique solution of problem (1.5). In particular, the operator A:D(A)L2(0,1) is non-positive and self-adjoint in L2(0,1) and it generates an analytic contraction semigroup of angle π/2. Moreover, let u0D(A). Then,

hW1,1(0,T;L2(0,1))uC1(0,T;L2(0,1))C([0,T];D(A)),
hL2(QT)uH1(0,T;L2(0,1)).

Proof.

Observe that D(A) is dense in L2(0,1). The existence of the unique solution follows in a standard way by a Faedo–Galerkin procedure, see, e.g., [43, Theorem 11.3] or [39, Theorem 3.4.1 and Remark 3.4.3]. Let us prove the other facts.

A is non-positive. By Proposition 2.18, Remark 2.16 and Lemma 2.21, for all uD(A), we have

-Au,uL2(0,1)=-01((au)+λbu)u𝑑x=01a(u)2𝑑x-λ01u2b𝑑xCu2.

A is self-adjoint. Let T:L2(0,1)L2(0,1) be the mapping defined in the following usual way: To each hL2(0,1) associate the weak solution u=T(h) of

01(auv-λuvb)𝑑x=01hv𝑑x

for every v. Note that T is well defined by the Lax–Milgram Lemma via Proposition 2.18, which also implies that T is continuous. Now, it is easy to see that T is injective and symmetric. Thus, it is self-adjoint. As a consequence, A=T-1:D(A)L2(0,1) is self-adjoint (for example, see [44, Proposition A.8.2]).

A is m-dissipative. Being A non-positive and self-adjoint, this is a straightforward consequence of [17, Corollary 2.4.8]. Then, (A,D(A)) generates a cosine family and an analytic contractive semigroup of angle π/2 on L2(0,1) (see, for instance, [3, Examples 3.14.16 and 3.7.5]).

The additional regularity is a consequence of [17, Lemma 4.1.5 and Proposition 4.1.6] in the first case, and of [2, 6.2.2 and 6.2.4] in the second one. ∎

3 Carleman estimates for singular/degenerate problems

In this section we prove one of the main result of this paper, i.e., a new Carleman estimate with boundary terms for solutions of the singular/degenerate problem

(3.1){vt+(avx)x+λb(x)v=h(t,x)=h,(t,x)QT,v(t,0)=v(t,1)=0,t(0,T),v(T,x)=vT(x),

which is the adjoint of problem (1.5).

On the degenerate function a we make the following assumption.

Hypothesis 3.1

Hypothesis 2.17 holds. Moreover, if K1>4/3, then there exists a constant θ(0,K1] such that the following condition is satisfied:

  1. The function xa(x)|x-x0|θ is non-increasing on the left of x=x0 and non-decreasing on the right of x=x0.

In addition, when K1>3/2 the function in (C2) is bounded below away from 0, and there exists a constant Σ>0 such that

(3.2)|a(x)|Σ|x-x0|2θ-3for a.e. x[0,1].

Moreover, if λ<0, we require that

(3.3)(x-x0)b(x)0in [0,1].
Remark 3.2

If a(x)=|x-x0|K1, then condition (C2) is clearly satisfied with θ=K1. Moreover, the additional requirements for the sub-case K1>3/2 are technical ones and were introduced in [30] to guarantee the convergence of some integrals (see [30, Appendix]). Of course, the prototype a(x)=|x-x0|K1 satisfies again such conditions with θ=K1. Finally, (3.3) is clearly satisfied by the prototype b(x)=|x-x0|K2.

To prove Carleman estimate, let us introduce the function φ:=Θψ, where

(3.4)Θ(t):=1[t(T-t)]4andψ(x):=c1[x0xy-x0a(y)𝑑y-c2]

with c2>sup[0,1]x0xy-x0a(y)𝑑y and c1>0 (for the observability inequality, c1 will be taken sufficiently large, see Lemma 4.7). Observe that Θ(t)+ as t0+,T-, and clearly -c1c2ψ<0.

The main result of this section is the following

Theorem 3.3

Assume Hypothesis 3.1. Then, there exist two positive constants C and s0, such that every solution v of (3.1) in

(3.5)𝒱:=L2(0,T;Ha,b2(0,1))H1(0,T;)

satisfies, for all ss0,

QT(sΘa(vx)2+s3Θ3(x-x0)2av2)e2sφ𝑑x𝑑tC(QTh2e2sφ𝑑x𝑑t+sc10T[aΘe2sφ(t,x)(x-x0)(vx)2dt]x=0x=1).
Remark 3.4

In [46] Vancostenoble and Zuazua proved a related Carleman inequality for the non-degenerate singular one-dimensional problem

(3.6){vt+vxx+μx2+λxβv=h,(t,x)QT,v(t,0)=v(t,1)=0,t(0,T),v(T,x)=vT(x),x(0,1),

where β[0,2). When μ=0 and x0=0, such an inequality reads as follows:

QT(s3Θ3x2v2+s2Θv2x2+s2Θv2x2/3)e2sΨ𝑑x𝑑t12QTh2e2sΨ𝑑x𝑑t,

where Ψ(x)=x22-1<0 in [0,1]. Actually, it is proved for solutions v such that

(3.7)v(t,x)=0for all (t,x)(0,T)×(1-η,1) and for some η(0,1).

However, in [46, Remark 3.5] Vancostenoble and Zuazua say that Carleman estimates can be proved also for all solutions of (3.6) not satisfying (3.7). We think that this latter situation is much more interesting, since by the Carleman estimates, if h=0, then v0 even if (3.7) does not hold.

The proof of Theorem 3.3 is quite long, and several intermediate lemmas will be used. First, for s>0, define the function

w(t,x):=esφ(t,x)v(t,x),

where v is any solution of (3.1) in 𝒱. Observe that, since v𝒱 and φ<0, w𝒱 and satisfies

(3.8){(e-sφw)t+(a(e-sφw)x)x+λe-sφwb=h,(t,x)(0,T)×(0,1),w(t,0)=w(t,1)=0,t(0,T),w(T,x)=w(0,x)=0,x(0,1).

As usual, we rewrite the previous problem as follows. Setting

Lv:=vt+(avx)x+λvbandLsw=esφL(e-sφw),

then (3.8) becomes

{Lsw=esφh,w(t,0)=w(t,1)=0,t(0,T),w(T,x)=w(0,x)=0,x(0,1).

Computing Lsw, one has

Lsw=Ls+w+Ls-w,

where

Ls+w:=(awx)x+λwb-sφtw+s2aφx2w,
Ls-w:=wt-2saφxwx-s(aφx)xw.

Clearly,

2Ls+w,Ls-w2Ls+w,Ls-w+Ls+wL2(QT)2+Ls-wL2(QT)2
(3.9)=LswL2(QT)2=hesφL2(QT)2,

where , denotes the scalar product in L2(QT). As usual, we will separate the scalar product Ls+w,Ls-w in distributed terms and boundary terms.

Lemma 3.5

The following identity holds:

(3.10)Ls+w,Ls-w=DT+BT,

where

DT=s2QTφttw2𝑑x𝑑t-2s2QTaφxφtxw2𝑑x𝑑t+sQT(2a2φxx+aaφx)(wx)2𝑑x𝑑t
+s3QT(2aφxx+aφx)a(φx)2w2𝑑x𝑑t-sλQTaφxbb2w2𝑑x𝑑t,
BT=0T[awxwt]x=0x=1𝑑t-s201[w2φt]t=0t=T𝑑x+s2201[a(φx)2w2]t=0t=T𝑑t
+0T[-sφx(awx)2+s2aφtφxw2-s3a2(φx)3w2-sλaφxbw2]x=0x=1𝑑t
+0T[-sa(aφx)xwwx]x=0x=1𝑑t-1201[a(wx)2-λ12bw2]t=0t=T𝑑x,

i.e., DT and BT are the distributed and boundary terms, respectively.

Proof.

Computing Ls+w,Ls-w, one has that

Ls+w,Ls-w=I1+I2+I3+I4,

where

I1:=QT((awx)x-sφtw+s2a(φx)2w)wt𝑑x𝑑t,
I2:=QT((awx)x-sφtw+s2a(φx)2w)(-2saφxwx)𝑑x𝑑t,
I3:=QT((awx)x-sφtw+s2a(φx)2w)(-s(aφx)xw)𝑑x𝑑t,
I4:=λQTwb(wt-2saφxwx-s(aφx)xw)𝑑x𝑑t.

By several integrations by parts in space and in time (see [1, Lemma 3.4], [29, Lemma 3.1] or [30, Lemma 3.1]) and by observing that QTa(aφx)xxwwx𝑑x𝑑t=0 (by the very definition of φ), we get

I1+I2+I3=s2QTφttw2𝑑x𝑑t-2s2QTaφxφtxw2𝑑x𝑑t+sQT(2a2φxx+aaφx)(wx)2𝑑x𝑑t
+s3QT(2aφxx+aφx)a(φx)2w2𝑑x𝑑t+0T[awxwt]x=0x=1𝑑t-s201[w2φt]t=0t=T𝑑x
+s2201[a(φx)2w2]t=0t=T𝑑t+0T[-sφx(awx)2+s2aφtφxw2-s3a2(φx)3w2]x=0x=1𝑑t
(3.11)+0T[-sa(aφx)xwwx]x=0x=1𝑑t-1201[a(wx)2]t=0t=T𝑑x.

Next, we compute I4:

I4=λ(QT12b(w2)t𝑑x𝑑t-2sQTabφxwxw𝑑x𝑑t-sQT(aφx)xbw2𝑑x𝑑t)
=λ(0112b[w2]t=0t=T𝑑x-sQTabφx(w2)x𝑑x𝑑t-sQT(aφx)xbw2𝑑x𝑑t)
=λ(0112b[w2]t=0t=T𝑑x-s0T[abφxw2]x=0x=1𝑑t+sQT(aφxb)xw2𝑑x𝑑t-sQT(aφx)xbw2𝑑x𝑑t)
(3.12)=λ(0112b[w2]t=0t=T𝑑x-s0T[aφxbw2]x=0x=1𝑑t-sQTaφxbb2w2𝑑x𝑑t).

By adding (3.11) and (3.12), (3.10) follows immediately. ∎

For the boundary terms in (3.10), we have the following lemma.

Lemma 3.6

The boundary terms in (3.10) reduce to

-s0T[Θ(awx)2ψ]x=0x=1𝑑t.

Proof.

As in [29] or [30], using the definition of φ and the boundary conditions on w, one has that

0T[awxwt]x=0x=1𝑑t-s201[w2φt]t=0t=T𝑑x+s2201[a(φx)2w2]t=0t=T𝑑t
+0T[-sφx(awx)2+s2aφtφxw2-s3a2(φx)3w2]x=0x=1𝑑t
(3.13)+0T[-sa(aφx)xwwx]x=0x=1𝑑t-1201[a(wx)2]t=0t=T𝑑x=-s0T[Θ(awx)2ψ]x=0x=1𝑑t.

Moreover, since w𝒱, we have that wC([0,T];); then w(0,x) and w(T,x) are well defined. From the boundary conditions of w, we get that

01[12bw2]t=0t=T𝑑x=0.

Now, consider the last boundary term

sλ0T[aφxbw2]x=0x=1𝑑t.

Using the definition of φ, this term becomes

λ0T[Θaψbw2]x=0x=1𝑑t.

By the definition of ψ, the function Θaψbw2 is bounded in (0,T). Thus, by the boundary conditions on w, one has

sλ0T[Θaψbw2]x=0x=1𝑑t=0.

Now, the crucial step is to prove the following estimate.

Lemma 3.7

Assume Hypothesis 3.1. Then, there exist two positive constants s0 and C such that for all ss0, the distributed terms of (3.10) satisfy the estimate

s2QTφttw2𝑑x𝑑t-2s2QTaφxφtxw2𝑑x𝑑t+sQT(2a2φxx+aaφx)(wx)2𝑑x𝑑t
+s30T01(2aφxx+aφx)a(φx)2w2𝑑x𝑑t-sλQTaφxbb2w2𝑑x𝑑t
   C2sQTΘa(wx)2dxdt+C32s3QTΘ3(x-x0)2aw2dxdt.

Proof.

Proceeding as in [29, Lemma 3.2] or in [30, Lemma 4.1], one can prove that, for s large enough,

s2QTφttw2𝑑x𝑑t-2s2QTaφxφtxw2𝑑x𝑑t
+sQT(2a2φxx+aaφx)(wx)2𝑑x𝑑t+s3QT(2aφxx+aφx)a(φx)2w2𝑑x𝑑t
   3C4sQTΘa(wx)2dxdt+C32s3QTΘ3(x-x0)2aw2dxdt,

where C is a positive constant. Let us remark that one can assume C as large as desired, provided that s0 increases as well. Indeed, taken k>0, from

Cs𝒜1+C3s3𝒜2=kCsk𝒜1+k3C3s3k3𝒜2,

we can choose s0=ks0 and C=kC large as needed.

Now, we estimate the term

-sλQTaφxbb2w2𝑑x𝑑t.

If λ<0, the thesis follows immediately by the previous inequality and by (3.3). Otherwise, if λ>0, by the definition of φ and the assumption on b, one has

-sλQTaφxbb2w2𝑑x𝑑t=-sλQTΘaψbb2w2𝑑x𝑑t=-sλc1QTΘ(x-x0)bb2w2𝑑x𝑑t-sλc1K2QTΘbw2𝑑x𝑑t.

Since w(t,) for every t[0,1], for w𝒱, by (2.2) we get

QTΘbw2𝑑x𝑑tC*QTΘa(wx)2𝑑x𝑑t.

Hence,

-sλQTaφxbb2w2𝑑x𝑑t-sλc1K2C*QTΘa(wx)2𝑑x𝑑t,

and we can assume, in view of what remarked above, that this last quantity is greater than

-sC4QTΘa(wx)2𝑑x𝑑t.

Summing up, the distributed terms of QTLs+wLs-w𝑑x𝑑t can be estimated as

DTC2sQTΘa(wx)2𝑑x𝑑t+C32s3QTΘ3(x-x0)2aw2𝑑x𝑑t

for s large enough and C>0. ∎

From Lemma 3.5, Lemma 3.6 and Lemma 3.7, we deduce immediately that there exist two positive constants C and s0 such that for all ss0,

(3.14)QTLs+wLs-w𝑑x𝑑tCsQTΘa(wx)2𝑑x𝑑t+Cs3QTΘ3(x-x0)2aw2𝑑x𝑑t-s0T[Θa2wx2ψ]x=0x=1𝑑t.

Thus, a straightforward consequence of (3.9) and of (3.14) is the next result.

Lemma 3.8

Assume Hypothesis 3.1. Then, there exist two positive constants C and s0 such that for all ss0,

(3.15)sQTΘa(wx)2𝑑x𝑑t+s3QTΘ3(x-x0)2aw2𝑑x𝑑tC(QTh2e2sφ(t,x)𝑑x𝑑t+s0T[Θa2(wx)2ψ]x=0x=1𝑑t).

Recalling the definition of w, we have v=e-sφw and vx=-sΘψe-sφw+e-sφwx. Thus, substituting in (3.15), Theorem 3.3 follows.

4 Observability results and application to null controllability

In this section we shall apply the just established Carleman inequalities to observability and controllability issues. For this, we assume that the control set ω satisfies the following assumption.

Hypothesis 4.1

The subset ω is such that either of the following holds:

  1. it is an interval which contains the degeneracy point, i.e.,

    (4.1)ω=(α,β)(0,1)andx0ω,
  2. it is an interval lying on one side of the degeneracy point, i.e.,

    (4.2)ω=(α,β)(0,1)andx0ω¯.

On the coefficients a and b we essentially start with the assumptions made so far, with the exception of Hypothesis 2.4, and we add another technical one. We summarize all of them in the following.

Hypothesis 4.2

  1. Assume one among Hypotheses 2.12.3 with K1+K22 and λ<1/C*.

  2. If λ<0, (3.3) holds.

  3. If K1>4/3, condition (C2) holds, and if K1>3/2, (3.2) is satisfied.

  4. If Hypothesis 2.1 or 2.2 holds, there exist two functions 𝔤Lloc([0,1]{x0}), 𝔥Wloc1,([0,1]{x0}) and two strictly positive constants 𝔤0, 𝔥0 such that 𝔤(x)𝔤0 for a.e. x in [0,1] and

    (4.3)-a(x)2a(x)(xB𝔤(t)𝑑t+𝔥0)+a(x)𝔤(x)=𝔥(x,B)

    for a.e. x,B[0,1] with x<B<x0 or x0<x<B.

Remark 4.3

Since we require identity (4.3) far from x0, once a is given, it is easy to find 𝔤,𝔥,𝔤0 and 𝔥0 with the desired properties. For example, if a(x):=|x-x0|α,α(0,1), we can take 𝔤(x)𝔤0=𝔥0=1 and

𝔥(x,B)=|x-x0|α2-1[α2sign(x-x0)(B+1-x)+|x-x0|]

for all x and B[0,1] with x<B<x0 or x0<x<B. Clearly,

𝔤Lloc([0,1]{x0})and 𝔥Wloc1,([0,1]{x0};L(0,1)).

Now, we associate to problem (1.5) the homogeneous adjoint problem

(4.4){vt+(avx)x+λb(x)v=0,(t,x)QT,v(t,0)=v(t,1)=0,t(0,T),v(T,x)=vT(x),

where T>0 is given and vT(x)L2(0,1). By the Carleman estimate in Theorem 3.3, we will deduce the following observability inequality for all the degenerate cases.

Proposition 4.4

Assume Hypotheses 4.1 and 4.2. Then, there exists a positive constant CT such that every solution vC([0,T];L2(0,1))L2(0,T;H) of (4.4) satisfies

(4.5)01v2(0,x)𝑑xCT0Tωv2(t,x)𝑑x𝑑t.

Using the observability inequality (4.5) and a standard technique (e.g., see [38, Section 7.4]), one can prove the following null controllability result for the linear degenerate problem (1.5).

Theorem 4.5

Assume Hypotheses 4.1 and 4.2. Then, given u0L2(0,1), there exists hL2(QT) such that the solution u of (1.5) satisfies

u(T,x)=0for every x[0,1].

Moreover,

QTh2𝑑x𝑑tC01u02(x)𝑑x

for some positive constant C.

4.1 Proof of Proposition 4.4

In this subsection we will prove, as a consequence of the Carleman estimate proved in Section 3, the observability inequality (4.5). For this purpose, we will give some preliminary results. As a first step, consider the adjoint problem

(4.6){vt+Av=0,(t,x)QT,v(t,0)=v(t,1)=0,t(0,T),v(T,x)=vT(x)D(A2),

where

D(A2)={uD(A):AuD(A)}

and

Au:=(aux)x+λub.

Observe that D(A2) is densely defined in D(A) for the graph norm (see, for example, [9, Lemma 7.2]) and hence in L2(0,1). As in [13, 12, 25, 29], define the following class of functions:

𝒲:={v is a solution of (4.6)}.

Obviously (see, for example, [9, Theorem 7.5]),

𝒲C1([0,T];Ha,b2(0,1))𝒱𝒰,

where 𝒱 is defined in (3.5) and

(4.7)𝒰:=C([0,T];L2(0,1))L2(0,T;).

We start with the following proposition.

Proposition 4.6

Proposition 4.6 (Caccioppoli’s inequality)

Assume Hypothesis 2.17. Let ω and ω be two open subintervals of (0,1) such that ωω(0,1) and x0ω¯. Let φ(t,x)=Θ(t)Υ(x), where Θ is defined in (3.4) and

ΥC([0,1],(-,0))C1([0,1]{x0},(-,0))

is such that

(4.8)|Υx|cain [0,1]{x0}

for some c>0. Then, there exist two positive constants C and s0 such that every solution vW of the adjoint problem (4.6) satisfies

(4.9)0Tω(vx)2e2sφ𝑑x𝑑tC0Tωv2𝑑x𝑑t

for all ss0.

Of course, our prototype for Υ is the function ψ defined in (3.4), since

|ψ(x)|=c1|x-x0|2a(x)1a(x)c1a(x),

by Lemma 2.6.

Proof of Proposition 4.6.

The proof follows the one of [29, Proposition 4.2], but it is different due to the presence of the singular term.

Let us consider a smooth function ξ:[0,1] such that

0ξ(x)1for all x[0,1],
ξ(x)=1,xω,
ξ(x)=0,x[0,1]ω.

Since v solves (4.6), we have

0=0Tddt(01ξ2e2sφv2𝑑x)𝑑t
=QT2sξ2φte2sφv2+2ξ2e2sφvvtdxdt
=2QTξ2sφte2sφv2𝑑x𝑑t+2QTξ2e2sφv(-λvb-(avx)x)𝑑x𝑑t
(4.10)=2QTξ2sφte2sφv2𝑑x𝑑t-2λQTξ2e2sφv2b𝑑x𝑑t+2QT(ξ2e2sφv)xavx𝑑x𝑑t.

If λ<0, then, differentiating the last term in (4.10), we get

2QTξ2e2sφa(vx)2𝑑x𝑑t=2λQTξ2e2sφv2b𝑑x𝑑t-2QTξ2sφte2sφv2𝑑x𝑑t-2QT(ξ2e2sφ)xavvx𝑑x𝑑t
-2QTξ2sφte2sφv2𝑑x𝑑t-2QT(ξ2e2sφ)xavvx𝑑x𝑑t,

and then one can proceed as in the proof of [29, Proposition 4.2], obtaining the claim.

Otherwise, if λ>0 and ε>0 is fixed, then by the Cauchy–Schwarz inequality, for w=ξesφv, we have

01ξ2e2sφv2b𝑑xC*01a(wx)2𝑑x
Cε01a[(ξesφ)x]2v2𝑑x+ε01ξ2e2sφa(vx)2𝑑xfor some Cε>0.

Moreover,

[(ξesφ)x]2Cχω(e2sφ+s2(φx)2esφ)
Cχω(1+1a)

for some positive constant C. Indeed, e2sφ<1, while s2(φx)2e2sφ can be estimated with

c(-maxΥ)2(Υx)2ca,

by (4.8), for some constants c>0. Thus,

2λQTξ2e2sφv2b𝑑x𝑑t2λCεQTa[(ξesφ)x]2v2𝑑x𝑑t+2λεQTξ2e2sφa(vx)2𝑑x𝑑t
(4.11)C0Tωv2𝑑x𝑑t+2λεQTξ2e2sφa(vx)2𝑑x𝑑t

for a positive constant C depending on ε. Hence, differentiating the last term in (4.10) and using (4.11), we get

2QTξ2e2sφa(vx)2𝑑x𝑑t=2λQTξ2e2sφv2b𝑑x𝑑t-2QTξ2sφte2sφv2𝑑x𝑑t-2QT(ξ2e2sφ)xavvx𝑑x𝑑t
C0Tωv2𝑑x𝑑t+2λεQTξ2e2sφa(vx)2𝑑x𝑑t
-2QTξ2sφte2sφv2𝑑x𝑑t-2QT(ξ2e2sφ)xavvx𝑑x𝑑t.

Thus, applying again the Cauchy–Schwarz inequality, we get

(2-2λε)QTξ2e2sφa(vx)2𝑑x𝑑tC0Tωv2𝑑x𝑑t-2QTξ2sφte2sφv2𝑑x𝑑t-2QT(ξ2e2sφ)xavvx𝑑x𝑑t
C0Tωv2𝑑x𝑑t-20Tωξ2sφte2sφv2𝑑x𝑑t
+2ε0Tω(aξesφvx)2𝑑x𝑑t+Dε0Tω(a(ξ2e2sφ)xξesφv)2𝑑x𝑑t
=C0Tωv2𝑑x𝑑t-20Tωξ2sφte2sφv2𝑑x𝑑t
+2ε0Tωξ2e2sφa(vx)2𝑑x𝑑t+Dε0Tω[(ξ2e2sφ)x]2ξ2e2sφav2𝑑x𝑑t

for some Dε>0. Hence,

2(1-ε-λε)0Tωξ2e2sφa(vx)2𝑑x𝑑t
C0Tωv2𝑑x𝑑t-20Tωξ2sφte2sφv2𝑑x𝑑t+Dε0Tω[(ξ2e2sφ)x]2ξ2e2sφav2𝑑x𝑑t.

Since x0ω¯, we have

2(1-ε-λε)infωa(x)0Tωe2sφ(vx)2𝑑x𝑑t2(1-ε-λε)0Tω¯ξ2e2sφa(vx)2𝑑x𝑑t
2(1-ε-λε)0Tωξ2e2sφa(vx)2𝑑x𝑑t
C0Tωv2𝑑x𝑑t-20Tωξ2sφte2sφv2𝑑x𝑑t
+Dε0Tω[(ξ2e2sφ)x]2ξ2e2sφav2𝑑x𝑑t.

Finally, we show that there exists a positive constant C (still depending on ε) such that

-20Tωξ2sφte2sφv2𝑑x𝑑t+Dε0Tω[(ξ2e2sφ)x]2ξ2e2sφav2𝑑x𝑑tC0Tωv2𝑑x𝑑t,

so that the claim will follow. Indeed,

|sφte2sφ|c1s01/4(-maxΥ)1/4,

|Θ˙|cΘ5/4 and

|sφte2sφ|cs(-Υ)Θ5/4e2sφc(s(-Υ))5/4

for some constants c>0 which may vary at every step.

On the other hand,

[(ξ2e2sφ)x]2ξ2e2sφ

can be estimated by

C(e2sφ+s2(φx)2e2sφ)χω,

and proceeding as before, we get the claim, choosing ε small enough, namely ε<(1+λ)-1. ∎

We shall also use the following lemma.

Lemma 4.7

Assume Hypotheses 4.1 and 4.2. Then, there exist two positive constants C and s0 such that every solution vW of (4.6) satisfies, for all ss0,

QT(sΘa(vx)2+s3Θ3(x-x0)2av2)e2sφ𝑑x𝑑tC0Tωv2𝑑x𝑑t.

Here Θ and φ are as in (3.4) with c1 sufficiently large.

Using the following non-degenerate classical Carleman estimate, one has that the proof of the previous lemma is a simple adaptation of the proof of [30, Lemma 5.1 and 5.2], to which we refer, also to explain why c1 must be large.

Proposition 4.8

Proposition 4.8 (Non-degenerate non-singular Carleman estimate)

Let z be the solution of

(4.12){zt+(azx)x+λzb=hL2((0,T)×(A,B)),z(t,A)=z(t,B)=0,t(0,T),

where bC([A,B]) is such that bb0>0 in [A,B] and a satisfies either of the following:

  1. aW1,1(A,B), aa0>0 in (A,B) and there exist two functions 𝔤L1(A,B), 𝔥W1,(A,B) and two strictly positive constants 𝔤0, 𝔥0 such that 𝔤(x)𝔤0 for a.e. x in [A,B] and

    -a(x)2a(x)(xB𝔤(t)𝑑t+𝔥0)+a(x)𝔤(x)=𝔥(x)for a.e. x[A,B].
  2. aW1,(A,B) and aa0>0 in (A,B).

Then, for all λR, there exist three positive constants C, r and s0 such that for any s>s0,

(4.13)0TAB(sΘ(zx)2+s3Θ3z2)e2sΦ𝑑x𝑑tC(0TABh2e2sΦ𝑑x𝑑t-BT),

where

BT={sr0T[a3/2e2sΦΘ(xB𝔤(τ)𝑑τ+𝔥0)(zx)2]x=Ax=B𝑑tif (a1) holds,sr0T[ae2sΦΘerζ(vx)2]x=Ax=B𝑑tif (a2) holds.

Here the function Φ is defined as Φ(t,x):=Θ(t)ρ(x), where Θ is as in (3.4),

(4.14)ρ(x):={-r[Ax1a(t)tB𝔤(s)𝑑s𝑑t+Ax𝔥0a(t)𝑑t]-𝔠if (a1) holds,erζ(x)-𝔠if (a2) holds,

and

ζ(x)=𝔡xB1a(t)𝑑t.

Here d=aL(A,B) and c>0 is chosen, in the second case, in such a way that max[A,B]ρ<0.

Proof.

Rewrite the equation of (4.12) as zt+(azx)x=h¯, where h¯:=h-λzb. Then, applying [30, Theorem 3.1], there exist two positive constants C and s0>0, such that

(4.15)0TAB(sΘ(zx)2+s3Θ3z2)e2sΦ𝑑x𝑑tC(0TABh¯2e2sΦ𝑑x𝑑t-BT)

for all ss0. Using the definition of h¯, the term 0TABe2sΦh¯2𝑑x𝑑t can be estimated in the following way:

(4.16)0TABh¯2e2sΦ𝑑x𝑑t20TABh2e2sΦ𝑑x𝑑t+2λ20TABz2b2e2sΦ𝑑x𝑑t.

Applying the classical Poincaré inequality to w(t,x):=esΦz(t,x) and observing that 0<infΘΘcΘ2, one has

2λ20TABz2b2e2sΦ𝑑x𝑑t=2λ20TABw2b2𝑑x𝑑t2λ2b02C0TAB(wx)2𝑑x𝑑t
C0TAB(s2Θ2z2+(zx)2)e2sΦ𝑑x𝑑t
0TABs2Θ(zx)2e2sΦ𝑑x𝑑t+0TABs32Θ3z2e2sΦ𝑑x𝑑t

for s large enough. Using this last inequality in (4.16), we have

(4.17)0TABh¯2e2sΦ𝑑x𝑑t20TABe2sΦh2𝑑x𝑑t+0TABs2Θ(zx)2e2sΦ𝑑x𝑑t+0TABs32Θ3z2e2sΦ𝑑x𝑑t.

Using this inequality in (4.15), (4.13) follows immediately. ∎

In order to prove Proposition 4.4, the last result that we need is the following.

Lemma 4.9

Assume Hypotheses 4.1 and 4.2. Then, there exists a positive constant CT such that every solution vW of (4.6) satisfies

01v2(0,x)𝑑xCT0Tωv2(t,x)𝑑x𝑑t.

Proof.

Multiplying the equation of (4.6) by vt and integrating by parts over (0,1), one has

0=01(vt+(avx)x+λvb)vt𝑑x
=01(vt2+(avx)xvt+λvvtb)𝑑x
=01vt2𝑑x+[avxvt]x=0x=1-01avxvtx𝑑x+λ2ddt01v2b𝑑x
=01vt2𝑑x-12ddt01a(vx)2+λ2ddt01v2b𝑑x
-12ddt01a(vx)2𝑑x+λ2ddt01v2b𝑑x.

Thus, the function

t01a(vx)2𝑑x-λ01v2b𝑑x

is non-decreasing for all t[0,T]. In particular,

01a(vx)2(0,x)𝑑x-λ01v2(0,x)b(x)𝑑x01a(vx)2(t,x)𝑑x-λ01v2(t,x)b(x)𝑑x(1+|λ|C*)01a(vx)2(t,x)𝑑x,

by Proposition 2.14. Integrating the previous inequality over [T/4,3T/4], Θ being bounded therein, we find

01a(x)(vx)2(0,x)𝑑x-λ01v2(0,x)b(x)𝑑x2T(1+|λ|C*)T/43T/401a(vx)2𝑑x𝑑t
(4.18)CTT/43T/401sΘa(vx)2e2sφ𝑑x𝑑t.

Hence, from the previous inequality and Lemma 4.7, if λ<0, then

01a(vx)2(0,x)𝑑x01a(vx)2(0,x)𝑑x-λ01v2(0,x)b(x)𝑑xC0Tωv2𝑑x𝑑t

for some positive constant C>0.

If λ>0, using again Lemma 4.7 and (4.18), one has

(4.19)01a(vx)2(0,x)𝑑x-λ01v2(0,x)b(x)𝑑xC0Tωv2𝑑x𝑑t.

Hence, by (2.2) and (4.19), we have

01a(vx)2(0,x)𝑑xλ01v2(0,x)b(x)𝑑x+C0Tωv2𝑑x𝑑tλC*01a(vx)2(0,x)𝑑x+C0Tωv2𝑑x𝑑t.

Thus,

(1-λC*)01a(vx)2(0,x)𝑑xC0Tωv2𝑑x𝑑t

for a positive constant C. In every case, there exists C>0 such that

(4.20)01a(vx)2(0,x)𝑑xC0Tωv2𝑑x𝑑t.

The Hardy–Poincaré inequality (see Proposition 2.7) and (4.20) imply that

01(a(x-x0)2)1/3v2(0,x)𝑑x01p(x-x0)2v2(0,x)𝑑x
CHP01p(vx)2(0,x)𝑑x
cCHP01a(vx)2(0,x)𝑑x
C0Tωv2𝑑x𝑑t

for a positive constant C. Here p(x)=(a(x)|x-x0|4)1/3 if K1>4/3, while p(x)=|x-x0|4/3max[0,1]a1/3 otherwise, and c,C are obtained by Lemma 2.6.

Again by Lemma 2.6, we have

(a(x)(x-x0)2)1/3C3:=min{(a(1)(1-x0)2)1/3,(a(0)x02)1/3}>0.

Hence,

C301v(0,x)2𝑑xC0Tωv2𝑑x𝑑t

and the claim follows. ∎

Proof of Proposition 4.4.

It follows by a density argument as for the proof of [29, Proposition 4.1]. ∎

Funding statement: The authors are members of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM), and supported by the GNAMPA project 2014 Systems with irregular operators. Dimitri Mugnai was also supported by the M.I.U.R. project Variational and perturbative aspects of nonlinear differential problems.

References

[1] Alabau-Boussouira F., Cannarsa P. and Fragnelli G., Carleman estimates for degenerate parabolic operators with applications to null controllability, J. Evol. Equ. 6 (2006), 161–204. 10.1007/s00028-006-0222-6Search in Google Scholar

[2] Arendt W., Semigroups and evolution equations: Functional calculus, regularity and kernel Estimates, Handbook of Differential Equations: Evolutionary equations, Volume I, North-Holland, Amsterdam (2004), 1–85. 10.1016/S1874-5717(04)80003-3Search in Google Scholar

[3] Arendt W., Batty C. J. K., Hieber M. and Neubrander F., Vector-valued Laplace Transforms and Cauchy Problems, Monogr. Math. 96, Birkhäuser, Basel, 2001. 10.1007/978-3-0348-5075-9Search in Google Scholar

[4] Baras P. and Goldstein J., Remarks on the inverse square potential in quantum mechanics, Differential Equations, North-Holland Math. Stud. 92, North-Holland, Amsterdam (1984), 31–35. 10.1016/S0304-0208(08)73675-2Search in Google Scholar

[5] Baras P. and Goldstein J., The heat equation with a singular potential, Trans. Amer. Math. Soc. 284 (1984), 121–139. 10.1090/S0002-9947-1984-0742415-3Search in Google Scholar

[6] Bebernes J. and Eberly D., Mathematical Problems from Combustion Theory, Appl. Math. Sci. 83, Springer, New York, 1989. 10.1007/978-1-4612-4546-9Search in Google Scholar

[7] Boutaayamou I., Fragnelli G. and Maniar L., Lipschitz stability for linear cascade parabolic systems with interior degeneracy, Electron. J. Differential Equations 2014 (2014), 1–26. Search in Google Scholar

[8] Boutaayamou I., Fragnelli G. and Maniar L., Carleman estimates for parabolic equations with interior degeneracy and Neumann boundary conditions, J. Anal. Math., to appear. 10.1007/s11854-018-0030-2Search in Google Scholar

[9] Brezis H., Functional Analysis, Sobolev Spaces and Partial Differential Equations, Springer, New York, 2011. 10.1007/978-0-387-70914-7Search in Google Scholar

[10] Brezis H. and Vazquez J. L., Blow-up solutions of some nonlinear elliptic equations, Rev. Mat. Complut. 10 (1997), 443–469. 10.5209/rev_REMA.1997.v10.n2.17459Search in Google Scholar

[11] Cabré X. and Martel Y., Existence versus explosion instantanée pour des équations de la chaleur linéaires avec potentiel singulier, C. R. Math. Acad. Sci. Paris 329 (1999), 973–978. 10.1016/S0764-4442(00)88588-2Search in Google Scholar

[12] Cannarsa P., Fragnelli G. and Rocchetti D., Null controllability of degenerate parabolic operators with drift, Netw. Heterog. Media 2 (2007), 693–713. 10.3934/nhm.2007.2.695Search in Google Scholar

[13] Cannarsa P., Fragnelli G. and Rocchetti D., Controllability results for a class of one-dimensional degenerate parabolic problems in nondivergence form, J. Evol. Equ. 8 (2008), 583–616. 10.1007/s00028-008-0353-34Search in Google Scholar

[14] Cannarsa P., Martinez P. and Vancostenoble J., Null controllability of the degenerate heat equations, Adv. Differ. Equ. 10 (2005), 153–190. 10.57262/ade/1355867887Search in Google Scholar

[15] Cannarsa P., Martinez P. and Vancostenoble J., Carleman estimates for a class of degenerate parabolic operators, SIAM J. Control Optim. 47 (2008), 1–19. 10.1137/04062062XSearch in Google Scholar

[16] Cazacu C., Controllability of the heat equation with an inverse-square potential localized on the boundary, SIAM J. Control Optim. 52 (2014), 2055–2089. 10.1137/120862557Search in Google Scholar

[17] Cazenave T. and Haraux A., An Introduction to Semilinear Evolution Equations, Clarendon Press, Oxford, 1998. 10.1093/oso/9780198502777.001.0001Search in Google Scholar

[18] Davies E. B., Spectral Theory and Differential Operators, Cambridge Stud. Adv. Math. 42, Cambridge University Press, Cambridge, 1995. 10.1017/CBO9780511623721Search in Google Scholar

[19] De Castro A. S., Bound states of the Dirac equation for a class of effective quadratic plus inversely quadratic potentials, Ann. Phys. 311 (2004), 170–181. 10.1016/j.aop.2003.12.007Search in Google Scholar

[20] Dold J. W., Galaktionov V. A., Lacey A. A. and Vazquez J. L., Rate of approach to a singular steady state in quasilinear reaction-diffusion equations, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 26 (1998), 663–687. Search in Google Scholar

[21] Epstein C. L. and Mazzeo R., Degenerate Diffusion Operators Arising in Population Biology, Ann. of Math. Stud. 185, Princeton University Press, Pinceton, 2013. 10.1515/9781400846108Search in Google Scholar

[22] Ervedoza S., Control and stabilization properties for a singular heat equation with an inverse-square potential, Comm. Partial Differential Equations 33 (2008), 1996–2019. 10.1080/03605300802402633Search in Google Scholar

[23] Fotouhi M. and Salimi L., Null controllability of degenerate/singular parabolic equations, J. Dyn. Control Syst. 18 (2012), 573–602. 10.1007/s10883-012-9160-5Search in Google Scholar

[24] Fotouhi M. and Salimi L., Controllability results for a class of one-dimensional degenerate/singular parabolic equations, Commun. Pure Appl. Anal. 12 (2013), 1415–1430. 10.3934/cpaa.2013.12.1415Search in Google Scholar

[25] Fragnelli G., Null controllability of degenerate parabolic equations in non-divergence form via Carleman estimates, Discrete Contin. Dyn. Syst. Ser. S 6 (2013), 687–701. 10.3934/dcdss.2013.6.687Search in Google Scholar

[26] Fragnelli G., Interior degenerate/singular parabolic equations in nondivergence form: well-posedness and Carleman estimates, J. Differential Equations 260 (2016), 1314–1371. 10.1016/j.jde.2015.09.019Search in Google Scholar

[27] Fragnelli G., Marinoschi G., Mininni R. M. and Romanelli S., A control approach for an identification problem associated to a strongly degenerate parabolic system with interior degeneracy, New Prospects in Direct, Inverse and Control Problems for Evolution Equations, Springer INdAM Ser. 10, Springer, Berlin (2014), 121–139. 10.1007/978-3-319-11406-4_7Search in Google Scholar

[28] Fragnelli G., Marinoschi G., Mininni R. M. and Romanelli S., Identification of a diffusion coefficient in strongly degenerate parabolic equations with interior degeneracy, J. Evol. Equ. 15 (2015), 27–51. 10.1007/s00028-014-0247-1Search in Google Scholar

[29] Fragnelli G. and Mugnai D., Carleman estimates and observability inequalities for parabolic equations with interior degeneracy, Adv. Nonlinear Anal. 2 (2013), 339–378. 10.1515/anona-2013-0015Search in Google Scholar

[30] Fragnelli G. and Mugnai D., Carleman estimates, observability inequalities and null controllability for interior degenerate non-smooth parabolic equations, Mem. Amer. Math. Soc., to appear. 10.1090/memo/1146Search in Google Scholar

[31] Fragnelli G., Goldstein G. Ruiz Goldstein J. A. and Romanelli S., Generators with interior degeneracy on spaces of L2 type, Electron. J. Differential Equations 2012 (2012), 1–30. Search in Google Scholar

[32] Fursikov A. V. and Imanuvilov O. Y., Controllability of Evolution Equations, Lect. Notes Ser. 34, Seoul National University, Seoul, 1996. Search in Google Scholar

[33] Galaktionov V. and Vazquez J. L., Continuation of blow-up solutions of nonlinear heat equations in several space dimensions, Comm. Pure Appl. Math. 50 (1997), 1–67. 10.1002/(SICI)1097-0312(199701)50:1<1::AID-CPA1>3.0.CO;2-HSearch in Google Scholar

[34] Garofalo N., Unique continuation for a class of elliptic operators which degenerate on a manifold of arbitrary codimension, J. Differential Equations 104 (1993), 117–146. 10.1006/jdeq.1993.1065Search in Google Scholar

[35] Hagan P. and Woodward D., Equivalent Black volatilities, Appl. Math. Finance 6 (1999), 147–159. 10.1080/135048699334500Search in Google Scholar

[36] Karachalios N. I. and Zographopoulos N. B., On the dynamics of a degenerate parabolic equation: Global bifurcation of stationary states and convergence, Calc. Var. Partial Differential Equations 25 (2006), 361–393. 10.1007/s00526-005-0347-4Search in Google Scholar

[37] Koch H. and Tataru D., Carleman estimates and unique continuation for second order parabolic equations with nonsmooth coefficients, Comm. Partial Differential Equations 34 (2009), 305–366. 10.1080/03605300902740395Search in Google Scholar

[38] Le Rousseau J. and Lebeau G., On Carleman estimates for elliptic and parabolic operators. Applications to unique continuation and control of parabolic equations, ESAIM Control Optim. Calc. Var. 18 (2012), 712–747. 10.1051/cocv/2011168Search in Google Scholar

[39] Lions J.-L. and Magenes E., Non-Homogeneous Boundary Value Problems and Applications, Volume I, Grundlehren Math. Wiss. 181, Springer, New York, 1972. 10.1007/978-3-642-65161-8Search in Google Scholar

[40] Martinez P. and Vancostenoble J., Carleman estimates for one-dimensional degenerate heat equations, J. Evol. Equ. 6 (2006), 325–362. 10.1007/s00028-006-0214-6Search in Google Scholar

[41] Micu S. and Zuazua E., On the lack of null controllability of the heat equation on the half-line, Trans. Amer. Math. Soc. 353 (2001), 1635–1659. 10.1090/S0002-9947-00-02665-9Search in Google Scholar

[42] Muckenhoupt B., Hardy’s inequality with weights, Studia Math. 44 (1972), 31–38. 10.4064/sm-44-1-31-38Search in Google Scholar

[43] Renardy M. and Rogers R. C., An Introduction to Partial Differential Equations, 2nd ed., Texts Appl. Math. 13, Springer, New York, 2004. Search in Google Scholar

[44] Taylor M. E., Partial Differential Equations I: Basic Theory, 2nd ed., Appl. Math. Sci. 115, Springer, New York, 2011. 10.1007/978-1-4419-7055-8Search in Google Scholar

[45] Vancostenoble J., Improved Hardy–Poincaré inequalities and sharp Carleman estimates for degenerate/singular parabolic problems, Discrete Contin. Dyn. Syst. Ser. S 4 (2011), 761–790. 10.3934/dcdss.2011.4.761Search in Google Scholar

[46] Vancostenoble J. and Zuazua E., Null controllability for the heat equation with singular inverse-square potentials, J. Funct. Anal. 254 (2008), 1864–1902. 10.1016/j.jfa.2007.12.015Search in Google Scholar

[47] Vancostenoble J. and Zuazua E., Hardy inequalities, observability, and control for the wave and Schrödinger equations with singular potentials, SIAM J. Math. Anal. 41 (2009), 1508–1532. 10.1137/080731396Search in Google Scholar

[48] Vázquez J. L. and Zuazua E., The Hardy inequality and the asymptotic behaviour of the heat equation with an inverse-square potential, J. Funct. Anal. 173 (2000), 103–153. 10.1006/jfan.1999.3556Search in Google Scholar

Received: 2015-11-18
Accepted: 2016-1-14
Published Online: 2016-2-16
Published in Print: 2017-2-1

© 2017 by De Gruyter

Downloaded on 5.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2015-0163/html
Scroll to top button